Download as pdf or txt
Download as pdf or txt
You are on page 1of 84

Wind Farm Integration in British

Columbia Stages 1 & 2: Planning and


Interconnection Criteria

Issued:
Revised:
Revised:

January 21st, 2005


March 4th, 2005
March 28th, 2005

Prepared for British Columbia Transmission


Corporation

Report Number: 2005-10988-2.R01.3

SUBMITTED BY:

Electric Systems Consulting


ABB Inc.
940 Main Campus Drive, Suite 300
Raleigh, NC 27606

Legal Notice
This document, prepared by ABB Inc., is an account of work sponsored by British Columbia
Transmission Corporation (BCTC). Neither BCTC nor ABB Inc., nor any person or persons acting
on behalf of either party: (i) makes any warranty or representation, expressed or implied, with respect
to the use of any information contained in this report, or that the use of any information, apparatus,
method, or process disclosed in this report may not infringe privately owned rights, or (ii) assumes
any liabilities with respect to the use of or for damages resulting from the use of any information,
apparatus, method, or process disclosed in this document.

Electric Systems Consulting

Technical Report

ABB Inc.
Title: Wind Farm Integration in British Columbia Dept.
Stages 1 & 2: Planning and Interconnection Consulting
Criteria

Date

Pages

3/28/2005

79

Author(s):

Reviewed by:

Approved by:

Pouyan Pourbeik
Executive Summary:

Willie Wong

Willie Wong

The British Columbia Transmission Corporation (BCTC) commissioned this study to seek recommendations on
establishing planning and interconnection criteria related to the incorporation of large amounts of wind generation
into the BCTC system. In addition, to review the potential operational impact of wind generation. This report
constitutes stages 1 and 2 of the study namely, the planning and interconnection aspects of the work.
This document presents a summary of the experience with wind generation within the North American continent
and elsewhere in the world. In addition, thorough discussions are given on various wind turbine generator
technologies. Based on this review and ABBs experience in performing detailed studies related to wind farms,
recommendations are given on the best practice emerging standards for interconnecting large wind farms to utility
grids.
The key points of interest are:
1. Interest in wind generation has been growing rapidly in the past few years and with estimates that wind
power could potentially supply 20% of the US energy demand it is expected that this trend will continue
for the immediate future.
2. Until recently, the only mature planning and interconnection standards related to wind generation were
those applied in Denmark and Germany. Increasingly, other utilities around the world are adopting
similar standards such as Ireland, Spain and even here in North America with the recent standard put out
by the Alberta Electric System Operator (AESO), Hydro Qubec and a proposal by the American Wind
Energy Associations (AWEA) that was submitted to FERC, and FERCs response to that proposal.
3. In the early days of wind turbine generator development, most (if not all) wind turbines were designed to
disconnect from the system immediately following a system fault. This was because most farms were
small and connected at distribution levels. With wind generation increasing in size, and being connected
to the bulk transmission systems, this is no longer an acceptable response. In Europe the practice (by the
Danish, German and Spanish power authorities) has been to demand adherence to a low-voltage ridethrough requirement. Similar low-voltage ride-through capabilities are now required by Hydro Qubec,
AESO, ESB National Grid in Ireland, FERC in the US etc. This has been documented here, and an
appropriate recommendation in this regard has been made for BCTC.
4. The other major concern is related to fluctuations in the wind farm output (MW output due to wind
speed variations) and their impact on system reserve. This is will be discussed in the companion report
on operations impact.
5. Other issues related to voltage flicker, harmonics, controls interaction and coordination of controls and
protection are discussed. Simple methodologies are provided to assess if and when such factors need to
be studied in detail.
In conclusion, the intent of this report is to objectively discuss the issues and concerns related to the integration of
wind generation into the BCTC System. The outcome has been a series of recommendations, all technically
feasible and viable, for possible incorporation into a new standard for interconnecting wind farms to the BCTC
system. The goal of any such standard will no doubt be to establish minimum technical requirements for
connecting wind farms to the transmission system. Such a standard should make no distinction as to the preferred
or not preferred technologies or types of wind generation equipment, but rather rely on the ingenuity of the wind
farm developers, wind turbine generator manufacturers and other power equipment manufacturers to come up
with the most cost effective means of meeting and/or exceeding the technical standards established. In the end,

ii

most of the issues surrounding integration of wind generation into the bulk transmission system are commercial
issues and limitations and not technical limitations. The technical problems can be addressed regardless of the
type or vendor of the wind generation system, provided proper analysis and design is performed upfront. As wind
generation technology advances they are becoming more and more economically viable and quite competitive
compared to other more traditional energy sources this ultimately will help to ensure greater utilization of our
natural renewable resources helping to safe guard our global environment.

iii

TABLE OF CONTENTS
TERMINOLOGY AND ACRONYMS .................................................................................................. 1
1

INTRODUCTION ............................................................................................................................ 2

2 WIND ENERGY BACK GROUND ON THE GLOBAL PENETRATION OF WIND


ENERGY, FUTURE TRENDS AND THE MAJOR TECHNOLOGIES.......................................... 4
2.1
WIND ENERGY CONVERSION ...................................................................................................... 4
2.2
WORLDWIDE PENETRATION OF WIND GENERATION AND EXPECTED FUTURE TRENDS ............ 6
2.2.1
North America .................................................................................................................... 6
2.2.2
Europe................................................................................................................................. 7
2.2.3
Australasia and Other Continents...................................................................................... 7
2.3
WIND TURBINE GENERATOR TECHNOLOGIES ............................................................................ 8
2.3.1
Wind Turbine Control Philosophies .................................................................................. 9
2.3.2
Conventional Induction Generators ................................................................................ 16
2.3.3
Doubly-Fed Induction Generators................................................................................... 18
2.3.4
Other Designs ................................................................................................................... 19
3

A SURVEY OF PLANNING AND INTERCONNECTION STANDARDS WORLDWIDE23


3.1
NORTH AMERICAN CONTINENT ................................................................................................ 23
3.1.1
Alberta .............................................................................................................................. 24
3.1.2
American Wind Energy Association Proposal to the Federal Energy Regulatory
Commission....................................................................................................................................... 26
3.1.3
California.......................................................................................................................... 27
3.1.4
Federal Energy Regulatory Commission Proposal......................................................... 28
3.1.5
Hydro-Qubec .................................................................................................................. 28
3.1.6
Midwest............................................................................................................................. 30
3.1.7
New Mexico ...................................................................................................................... 30
3.1.8
New York........................................................................................................................... 30
3.1.9
Texas ................................................................................................................................. 31
3.1.10 Western Electricity Coordinating Council (WECC) ....................................................... 31
3.2
EUROPE...................................................................................................................................... 32
3.2.1
Denmark ........................................................................................................................... 32
3.2.2
Germany ........................................................................................................................... 34
3.2.3
Ireland............................................................................................................................... 36
3.2.4
Spain ................................................................................................................................. 39
3.3
REST OF THE WORLD................................................................................................................. 41
3.4
GENERAL OBSERVATIONS AND EXPERIENCE WITH WIND GENERATION STUDIES IN THE US.. 41

PLANNING CRITERIA AND STUDY APPROACH ............................................................... 44


4.1
PLANNING.................................................................................................................................. 44
4.1.1
Modeling of WTG and Wind Farms for Steady-State and Dynamic Studies:................. 44
4.1.2
Reactive Capability: ......................................................................................................... 52
4.1.3
Line Ampacity ................................................................................................................... 53
4.1.4
Other Planning Issues:..................................................................................................... 54

INTERCONNECTION STANDARDS AND CRITERIA......................................................... 68


5.1
SUMMARY AND COMMENTS ON WIND STANDARDS WORLD-WIDE ......................................... 68
5.2
WIND GENERATION FACILITIES WITH AGGREGATED CAPACITY OF LESS THAN 10 MVA ...... 70
5.3
PROPOSED WIND INTERCONNECTION STANDARD OUTLINE FOR WIND FARMS WITH AN
AGGREGATED CAPACITY GREATER THAN 10 MVA............................................................................. 71

CONCLUSIONS & RECOMMENDATIONS ............................................................................ 76

REFERENCES........................................................................................................................................ 77

iv

Terminology and Acronyms


AVR
DFIG
GW
HVDC
LVRT
MVA

- Automatic Voltage Regulator


- Doubly-fed induction generator
- Gigawatts (units for measuring active power). 1 GW = 1000 MW.
- High-voltage dc transmission system
Low-voltage ride-through
- Megavolt-amps (unit for measuring total combined real and reactive
power)
MW
- Megawatts (units for measuring active power)
PLC
Programmable Logic Control
STATCOM Static compensators based on voltage-source converter technologies.
SVC
Static VAr Compensator. This term is used typically to refer to thyristor
based technology for dynamic shunt compensation.
VAr
- Volt-amps reactive (unit for measuring reactive power)
WTG
Wind Turbine Generator
WindVAr A term used by various manufacturers of wind turbine manufacturers to
refer to a system of providing voltage regulation. For doubly-fed
induction generator manufacturers, this refers to a centralized
proportional-integral regulator that measures the voltage at the collector
substation and based on the error between the measured and reference
voltage sends a command signal to all the WTGs in the farm to adjust their
reactive power output in order to regulate the collector substation voltage.
For some conventional induction generator manufacturers, this refers to
thryristor switched capacitor banks (essentially a static-VAr compensator)
installed together with their conventional induction generators, thus giving
dynamic VAr capability at each WTG and allowing for voltage regulation.

1 Introduction
The Kyoto Protocol, signed by 160 nations1 on 11th December 1997, is a legally binding
protocol signed by industrialized nations aimed at reducing six greenhouse gases (CO2,
CH4, N2O, HFCs, PFCs, SF6) by year 2008 to 2012. The original protocol aimed at a
collective reduction among these nations of 5.2% in greenhouse gas emissions. Since the
emissions by some countries actually increased after the signing of the protocol, when
compared to year 2000 emission levels the actual required reduction is roughly 10%.
Canadas share means a reduction of 6%1. This among other reasons is one of the
primary driving forces behind an increase in renewable energy generation globally.
Wind energy is one of the most mature of the various renewable energy technologies2 and
has recently gained much favor in Canada, the USA, Europe and other parts of the world.
Wind energy resources have dramatically increased over the past decade. At the end of
2003 the total installed capacity of wind generation in Europe was up to 28.7 GW
(www.ewae.org). Presently, there is an estimated 6.4 GW of installed capacity of wind
generation in the USA (www.awea.org) and 439 MW in Canada (www.canwea.ca). It is
estimated that wind energy could supply up to 20% of the US national energy demand
[1]. In Canada, the Canadian Wind Energy Association supports a goal of achieving 10
GW of installed capacity by 2010; to supply 5% of the electricity demand in Canada. It
is hence clear that wind will continue to be a significant portion of newly proposed
generating facilities in both the North American and European continents.
A few important characteristics of wind generation that stand out as compared to other
conventional generation technologies are,
1. Wind farms tend to be composed of large numbers of turbines spread out across a
relatively large geographical area, as opposed to conventional thermal plants
where a single turbine is capable of generating the same order of magnitude of
power.
2. Remote location. Particularly in Europe, many of the new facilities are aimed at
offshore sites where wind profiles are more constant, thereby ensuring higher
capacity factors for the wind farm. It is interesting to note that British Columbia
also has significant offshore potential.
3. Intermittent nature of the power generated.
4. The significantly different wind turbine generator technologies.
In almost all regions of the North American Continent, each (sometimes a few together)
proposed wind farm is usually evaluated through a System Impact Study i.e. comparing
the performance of the system with and with-out the proposed wind farm. A typical
system impact study consists of four major parts, namely, i) thermal evaluation ii) impact
1

See http://www.iitap.iastate.edu/gcp/kyoto/finalagree.html
Here we are referring to modern renewable technologies such as wind, photovoltaics, etc. If one were to
consider systems such as hydro-power as renewable (which they indeed are), then one could say that hydro
power technologies are far more mature than wind energy technologies.
2

on transfer capability of key interfaces iii) effect on stability and iv) impact on shortcircuit levels. As the penetration and size of wind farms increases, its impact on the
transmission grid requires a much broader analysis. Adopting suitable reliability criteria
for evaluating the steady-state and dynamic performance of these technologies on the grid
is very important.
The British Columbia Transmission Company (BCTC) has received inquires for wind
farm interconnection in the province of several hundred MW. This study constitutes an
initiative by BCTC to investigate the potential impact of wind generation on the British
Columbia electric transmission system. This report constitutes the first two stages of this
work. The documentation here is primarily qualitative and is an overview of existing
planning and interconnection criteria established by US and European electric authorities.
In addition, ABB has performed both detailed and system impact studies related to wind
farm interconnections in the US and Europe. Thus, using these resources this document
addresses the following topics:

Section 2 presents an overview of wind energy and various wind turbine


technologies. This section establishes the background upon which some of
the materials in the following sections are developed.
Section 3 presents an overview of planning standards and interconnection
requirements for utilities in the US, Europe and Australasia3, and how, if at
all, wind farm interconnections are specifically addressed in these standards.
Section 4 presents recommendations based on the literature review presented
in section 3 and our experience with wind farm studies, on the approach
BCTC may take towards planning criteria and studies for incorporating wind
farms into the BCTC system.
Section 5 presents the proposed approach BCTC may take towards
interconnection standards and criteria for incorporating wind farms into the
BCTC system.
Section 6 presents the conclusions and recommendations of this work.

The noun Australasia is commonly used to refer to the region of the world consisting of Australia, New
Zealand and the nearby islands.

2 Wind Energy Back Ground on the Global Penetration of


Wind Energy, Future Trends and the Major Technologies
2.1

Wind Energy Conversion

Wind energy has been in use for centuries. The concept is simple, to capture the kinetic
energy of the wind to force the rotation of a turbine and to then use this mechanical
power to perform useful work. Originally, wind turbines (or wind mills) were used for
pumping water, grinding grain and other such agricultural activities. The first known
windmills were developed for the tasks of grain-grinding and water-pumping the
earliest designs were of a vertical axis system developed in Persia (Iran) about 500-900
A.D [2]. The first windmills to appear in Europe were of a horizontal design and the
Dutch set out in the 1390s A.D. to refine this design. It is thus no surprise that Dutch
wind turbine manufacturers today are the leading manufacturers since they have had a
long affinity for wind power generation.
In recent years technological advancements have made it possible to utilize wind energy
for the production of electricity. Given that the fuel source (wind) is inexhaustible and
free, the urge to utilize this resource is clear. Never-the-less, there are significant
challenges, both from an equipment design stand point and a system design standpoint, to
effectively exploit the wind energy reserve.
Figure 2-1 shows a diagrammatic representation of a wind turbine.

Figure 2-1: Wind turbine.

As shown in Figure 2-1, a wind turbine consists of the rotor blades, the nacelle, a gearbox
and a generator. In addition, there will be electronic controls and other ancillary
equipment, such as a step up transformer, associated with the unit. These controls may
be mounted either in the nacelle or at the base of the tower. Most modern turbines use a
three blade design that point upwind. As the wind blows over each blade it causes lift
much like on an airplane wing and thus causes the rotor to rotate. The electrical
generator extracts this mechanical power and converts it to useful electrical power. The
gearbox is the mechanical transition between the rotor blades, often rotating at ten to
twenty rounds per minute and the generator rotating at seventy to eighty times faster
some modern wind turbine designs are gearless.
The theoretical maximum efficiency of a wind turbine is given by Betzs law [3]. The
law states that a lifting rotor can at most extract 60%4 of the energy from the air stream.
In practice, modern designs aim at efficiencies in the order of 40%. Moreover, for a wind
turbine there is no single efficiency since the efficiency of the turbine is a function of the
wind speed. Thus, often performance coefficients are quoted as a function of wind speed;
that is, the ratio of power extracted to power available in the wind at a given wind speed.
In the early 1980s a typical wind turbine had a rotor diameter of 10 meters and would
generate in the order of 25 kW. Modern wind turbines such as the NEG Micon NM825,
GE 1.5MW and Vestas V90 have rotor diameters of 70 to 90 meters and generate
between 1.5 to 3 MW. In addition, wind turbines designed primarily for offshore
applications, where winds are more prevalent, presently have reached ratings of 4.5 MW.
When deployed in a wind farm, the typical spacing between adjacent wind turbines is
between 3 to 5 rotor diameters (depending on the actual farm layout). Thus, the modern
wind farm, which may consist of 50 to 100 turbines, will span several square kilometers
of land (or sea).
An important concept is the expected energy output of a typical wind turbine (or farm)
over an annual period. This is often expressed as the capacity factor of the wind turbine
(or farm). The capacity factor is defined as:
capacity factor =

Actual annual energy produced


Energy produced if wind turbine (farm) was at full capacity for the entire year

Clearly the capacity factor of a wind farm depends on two things:


1. The design and performance of the wind turbines
2. The wind profile at the site
A reasonably economic capacity factor may range between 0.25 to 0.3. Anything above
0.3 would be a very good site. For land wind farms, it is rare to find sites with a capacity
factor much higher than 0.3 to 0.35. Offshore sites, on the other hand, tend to have
higher capacity factors and typically range between 0.35 to 0.45 [4, 5].
4

59.3% to be exact.
NEG Micon and Vestas merged in 2004 to become Vestas Wind Systems. Thus, commercially speaking
these units, or similar units, would now be supplied by Vestas.

It is interesting to note that the typical time frame between commencement of work on
construction of a wind farm and full operation of the farm is 12 months [1]. In
comparison, the development of conventional power plants and/or transmission assets
may typically take several years. Thus, without proper and advanced planning, wind
generation assets may grow in a system so rapidly as to not allow adequate transmission
reinforcements to be implemented in time to facilitate their interconnection.
2.2

Worldwide Penetration of Wind Generation and Expected Future


Trends

As of January, 2004 the total estimated global wind generation installed capacity is 37
GW [6]. This section gives a brief description of the distribution of this capacity on each
continent and the trends in the industry.
2.2.1 North America
The North American Continent presently has an estimated 6.8 GW of installed wind
generation. Of this only 439 MW are in Canada, the rest resides in the USA. A list of
current wind farm projects (and proposed projects) in the US may be found at the
American Wind Energy Association website (www.awea.org). The leading regions for
wind installations are California, Texas and the Midwest (particularly Minnesota and
Iowa), respectively.
Since the 1980s the price of wind generation has dropped from 38 cents/kWh to roughly
4 to 7 cents/kWh (in US currency), and continues to decline [6]. Thus the industry is
continuing to grow, as the price of wind energy is quite comparable today with the
traditional fossil fuel resources. The consumption of renewable energies in the United
States increased by 11% in 2002, as compared to 2001. The largest increase was in hydro
power, primarily in the installed base, while the second largest increase was in wind
energy [7].
It is estimated that wind energy could supply up to 20% of the US national energy
demand [1, 6]. It is also interesting to note that of the top twenty states with greatest
wind energy potential, North Dakota, Texas, Kansas, South Dakota and Montana are the
top five with estimated potential of greater than 1 billion kWhs each, while California
(the leading state with installed capacity) is seventeenth on the list [1]. In 2003 alone,
ABB Electric Systems Consulting was engaged in system impact studies for close to 1
GW of proposed wind farms in the Midwest. In 2003-2004, we performed a study for the
Alberta Electric System Operator looking at wind generation interconnection issues.
There is some 1200 MW of proposed wind generation in Alberta alone. It is therefore
fair to assume that wind generation will continue to grow in the North American
continent, both in Canada and the US.
One of the methods used by the Federal government in the USA to promote wind
generation development has been the production tax-credit (PTC). The PTC was first
enacted in 1992 (http://www.worldlinkinsurance.com/windpro/PTC.htm). The PTC

applies to all forms of renewable energy and essentially offers $0.019 $/kWh (2003
numbers) produced by the facility for the first tens years of the facilities life. In contrast,
in European countries such as Germany customers pay a certain additional amount per
month in their utility bill for the use of renewable energy.
2.2.2 Europe
Europe has hitherto been the leader in utilizing wind energy. A map showing the total
installed capacity of wind generation in each region of Europe may be found on the
European Wind Energy Association website (www.ewea.org).
Germany, Spain and Denmark are the leading countries with wind farm installed
capacity. In Denmark, wind energy supplies nearly 18% of the national energy needs [8]
the installed capacity is 60% of the nations peak load. Because of this relatively high
level of penetration, Denmark has started looking to slow down the growth of wind
capacity in the country as well as the potential for developing storage technologies for
enhancing operation of the fluctuating wind power [9]. Presently, Eltra has postponed its
work on developing energy storage technologies, waiting to see the outcome of a few
pilot projects [10]. It is generally accepted that if the total wind generation in a control
area is less than 10% of demand then there is minimal impact on reserve and operations
[1]. This subject is discussed further in the companion report.
Of all the utilities world wide, the German utility E.ON and the Danish utility Eltra
arguably have the most mature planning and interconnection standards with respect to
wind turbine generation. To add to this list, recently ESB National Grid (soon to be
know as EirGrid), which is the electric system operator in Ireland, also recently came out
with comprehensive reports and grid codes related to wind generation interconnection, in
2004. As such much of the noteworthy material presented in Section 3 is based on these
three sources.
Proposals for wind farms continue to grow in Europe, with projected realistic potential
for wind energy in Europe being some 343 TWh/annum [11].
2.2.3 Australasia and Other Continents
The installed capacity of wind generation in Australia nearly doubled from 198 MW in
2003 to 380 MW by the end of 2004 [12]. There is presently an additional 5.6 GW of
wind generation proposed for installation in Australia. One of the signs of an expected
insurgence of wind farms down under is the recent establishment of wind turbine
manufacturing facility by Vestas in Tasmania (www.auswea.org). It is also worth noting
that the proposed amount of wind generation has also almost doubled since this time last
year with the primary focus being the Southern States (www.auswea.org). For one state,
the proposed amount of wind generation capacity is comparable in magnitude to the
regions peak load.

New Zealand is also seeing a surge in wind generation. In 2004 the Te Apiti project was
installed, which is presently the largest wind farm in the Southern hemisphere [13]. The
site has fifty-five NEG Micon NM72 wind turbines (total capacity of 90 MW) and has a
capacity factor of more than 50% - this is perhaps the largest on land capacity factor for
any site in the world and is certainly not typical of land wind farms.
2.3

Wind Turbine Generator Technologies

The largest wind turbine manufacturers worldwide are:


Vestas Danish Company (merger of Vestas and NEG Micon)
Enercon German Company
GE Wind (formerly Enron) US Company
Gamesa Spanish Company
Bonus Danish Company (now owned by Siemens)
In 2004, some significant business ventures were forged. The largest and third largest
manufacturers merged to become Vestas Wind Systems (that is Vestas and NEG Micon),
making Vestas by far the largest manufacturer in the world (based on [14] the combined
estimated market share of Vestas and NEG Micon was 38% in 2002). Also, Siemens
entered into the wind generation business by buying out Bonus. It is not surprising that
among the top are Danish, German and Spanish companies, where the majority of
worldwide wind turbines installations presently exist. The other smaller manufacturers
trying to obtain a significant share of the market are Nordex (German), Made (Spanish),
Mitsubishi (Japanese), Repower (German) and Suzlon (Indian).
There are essentially three major types of wind turbine designs:
1. Constant speed turbines (conventional induction generators)
2. Variable speed turbines (doubly-fed induction generators, variable rotor resistance
induction generators)
3. Gear-less turbines (slow speed conventional generators connected to the grid by a
back-to-back frequency converter)
More recently, a new concept in the electrical machine design is also being proposed, that
of a permanent magnetic generator, which requires no excitation. These fall into the last
of the three categories above.
Figures 2-3 to 2-6 show the electrical lay out for these major designs. Note: other
designs also exist, such as permanent magnetic generator designs with a gearbox;
however, the four figures here represent the major concepts. The figures were provided
by ABB Motors & Drives, Finland, one of the leading suppliers of electrical generators to
the wind turbine manufacturing industry. Presently, ABB supplies 30% of the electrical
generators to the major wind turbine manufacturers worldwide.

Presently, the majority of wind turbines sold in the European market (with the exception
of Germany) have been fixed-speed conventional induction generators. Variable-speed
machines of the doubly-fed and full converter style are certainly starting to gain market
share. In the USA, due to patent right issues [15], the doubly-fed units have thus far been
supplied by one vendor. Most of the gear-less, full converter designed units are inservice in Germany, since the dominant manufacturer that supplies this design is the
German manufacturer Enercon.
2.3.1 Wind Turbine Control Philosophies
The power extracted from wind by a wind turbine may be given by the equation
1
P = v 3 r 2 C p ( )
2
where
P = mechanical power
= air density
v = wind speed
r = wind turbine rotor radius
Cp = coefficient of efficiency
= tip-speed ratio (i.e. the ratio of blade tip speed to wind speed)

(1)

Based on (1), at any given time one does not have any control over the air density, the
speed of the air or the radius of the wind turbine rotor. Thus, to maximize energy output
the only parameter that can be controlled is Cp. It can be shown that the theoretical
maximum value of Cp is 0.593 (this is Betzs Law)6. In practice one can achieve Cp
values close to 0.4. For a given blade pitch and rotation speed, Cp is a non-linear
function of wind speed and will peak at a given tip-speed ratio and drop off again to zero
at higher tip-speed ratios. A typical site for good wind energy production may have
average wind speeds ranging around 7 to 10 m/s. Very high wind speeds are quite rare
and also tend to put significant stress on the turbine. Thus, the general approach is to
design wind turbines to extract the maximum amount of wind energy possible at wind
speeds between 10 to say 15 m/s, and to start to spill away some of the power at wind
speeds in excess of 15 m/s until they shut-down at relatively high wind speed typically,
in excess of 20 to 25 m/s. To achieve this necessitates a form of power control on the
turbine. There are fundamentally two different ways of achieving wind turbine power
control (i) fixed-speed designs that typically use stall, and (ii) variable speed designs that
typically use dynamic blade pitch-control.
2.3.1.1 Stall and Active-Stall for Fixed Speed Wind Turbines
Fixed speed wind turbines, which typically used conventional induction generators (see
section 2.3.2), are in general controlled by stall or active-stall design. For turbines that
have a capacity of around one megawatt or less stall is used. For the larger turbines,
6

See www.windpower.org for a proof of Betzs Law.

active-stall is used since the higher power levels require some pitch control to prevent
excessive turbine stress.
In the stall design, the blades of the wind turbine are essentially bolted to the hub at a
fixed angle. The blades are aerodynamically designed so that as the wind speed increases
beyond a certain point the blade shape gradually begins to yield turbulence in the wind
and thus eventually results in the blades stalling (much like an airplane that tries to climb
too quickly at too sharp an angle of attack and thus stalls). Considering equation (1),
essentially Cp is designed to peak around the expected average wind speed for the site
and then gradually drops off until the unit stalls at high wind speeds. The advantage of
this design is that it avoids mechanical moving parts and some of the controls and other
complexities associated with pitch control. Thus, these units are typically cheaper than
an equivalent variable speed system. In contracts, there are significant challenges to
perfecting the aerodynamic design of a stall system. Also, sudden changes in wind speed
(such as a gust) will translate into sudden changes in turbine torque and generator output.
This requires more robust design of the drive train. In addition, if such units are used in
very weak grid connections they can result in more significant voltage flicker.
For larger turbines (1 MW and up) a variation of this concept, called active-stall, is used.
In active-stall the blades of the turbine do have a pitch control mechanism. At low wind
speeds pitch control of the blades is used similar to that of pitch controlled systems.
Even though the unit is rotating at constant speed, by changing the blade pitch at lower
wind speeds Cp can be changed and thus turbine efficiency improved somewhat. At high
wind speed, when the machine has reached its rated output, active-stall allows much
better control of the turbine than passive stall. If the wind speed increases suddenly,
active-stall systems will pitch the blades in the opposite direction than pitch-controlled
machines that is, they increase the angle of attack into the wind in order to force stalling
of the blades much quicker.
2.3.1.2 Pitch-controlled Turbines
Based on equation (1), if we were able to keep Cp as high as possible over the operating
range of the turbine then we would be able to capture the maximum amount of wind
energy. Theoretically, this can be achieved by maintaining essentially a fixed tip-speed
ratio [16]. This can be achieved several ways such as adjusting the turbine speed as a
function of wind speed or programming the turbine output as a function of turbine speed
based on a preprogrammed power-speed curve that leads to maintaining a fixed tip-speed
ratio. In practice, it is often difficult to obtain a good and reliable measurement of wind
speed (since the anemometer sits behind the rotor on the nacelle and thus the wind speed
being measured by it is distorted [17]). Thus, a typical design used is shown in Figure 22. Based on a measurement of the rotor speed, the power converter is controlled to adjust
the electrical power output of the unit to follow a predefined power-speed curve. This
preprogrammed curve is based on manufacturer calculations that will essentially maintain
a fixed tip-speed ratio and thus optimize energy conversion.

10

All variable speed turbines such as double-fed induction generators and full converter
units (see sections 2.3.3 and 2.3.4) use pitch-control. The other advantage of these
systems is that they produce greater power quality with less likelihood of flicker. This is
because during a short lived wind gust the rotor speed increases and thus absorbs some of
the fluctuation in power in the form of stored mechanical energy rather than directly
translating the power fluctuations onto the grid with fixed-speed systems.
2.3.1.3 Fixed-speed versus Variable Speed Turbines
As discussed above, one of the motivations for variable-speed pitch-control systems is
the ability to capture greater wind energy. In Reference [18], through calculations, the
authors show that a variable speed system can capture more energy over time than a fixed
speed design. In answer to such claims some fixed speed designs actually employ active
stall (explained above) or dual-speed, dual-wound machines that is, the stator of the
generator can be switched between a smaller rated 6-pole stator winding for low wind
speeds and a higher rated 4-pole winding for higher wind speeds. In these ways the
machines efficiency is somewhat improved at lower wind speeds. In the end the cost of
electricity production is determined by many factors and is not only a function of the
energy captured by the wind turbines. Energy production costs depend also on initial
capital cost, projected life span of the wind farm, reliability of the machines, operating
and maintenance costs and other factors. Thus, it has in general been hard to truly verify
if one design is superior to the other over the lifetime of a wind farm [19], and so all of
the above designs have persisted in the market. In the long run, however, as power
electronic equipment become more reliable and less expensive it is expected that variable
speed systems may start to dominate the market [17].

Figure 2-2: Control strategy for variable-speed turbines.

11

generator switchgear
(power switch)

main circuit breaker

f = constant
n = constant
brake

Getriebe 1:50

rotor
bearing

10...24 kV, f = 50 Hz
or 60 Hz

gearbox

start up
equipment
asynchronous generator
with squirrel cage rotor
and two windings

line coupling
transformer
medium voltage
switchgear

wind turbine
control

(Source ABB Motors&Drives, Finland )

Figure 2-3: Conventional induction generator. Constant speed drive.

12

main circuit breaker

gearbox
10...24 kV, f = 50 Hz
or 60 Hz
brake

asynchronous
generator
with slip rings

generator
side
converter

grid side
converter

line coupling
transformer
medium voltage
switchgear

rotor bearing

pitch
drive

frequency
converter

converter control

wind turbine control

(Source ABB Motors&Drives, Finland )

Figure 2-4: Doubly-fed induction generator. Variable speed drive.

13

frequency
converter

line side
converter

rectifier

main circuit
breaker

DC

excitation
converter
10...24 kV, f = 50 Hz
converter
control

rotor
bearing

line coupling
transformer
medium voltage
switchgear

brake
pitch
drive

synchronous
generator

wind turbine control

(Source ABB Motors&Drives, Finland )


Figure 2-5: Gear-less synchronous generator with back-to-back frequency converter.

14

frequency
converter

generator side
converter
DC

main circuit
breaker

line side
converter

10...24 kV, f = 50 Hz

converter
control

line coupling
transformer
medium voltage
switchgear
brake
rotor bearing

synchronous
generator

pitch
drive

wind turbine control

(Source ABB Motors&Drives, Finland )

Figure 2-6: Advanced gear-less design with permanent magnet generator.

15

2.3.2 Conventional Induction Generators


The conventional induction generator is essentially a constant speed system. Though
the electrical generator speed may vary by a fraction of a percent to a percent as the
machine goes from low load to high load, from a mechanical stand point the machine
essentially operates at a constant speed. Figure 2-3 shows an example of this system.
The soft-start thyristor-controlled converter is used to start the machine with minimal
impact on system voltage. Once the unit is connected to the grid, it runs essentially as
a super-synchronous induction generator (see Figure 2-7).

Figure 2-7: Induction machine torque speed curve.

Conventional induction generators, similar to induction motors, absorb reactive power


from the system. This reactive power essentially sustains the rotating magnetic field
in the air gap between the cage rotor and the stator windings. Typically,
manufacturers supply switched capacitor banks at the turbine that are switched based
on the machine kilowatt output. Thus, the effective power factor of the machine is
kept at or very close to unity, when operating at or above rated voltage. Moreover,
some manufacturers provide the added option of placing a thyristor based dynamic
VAr device at the turbine for additional and smoother VAr control [20].
One of the issues that will be discussed in some detail in the following section is that
of low-voltage ride-through. Namely, the ability of the wind turbine to ride-through a
grid disturbance. Since in the past most wind turbines were connected at the
distribution level, it was common practice to disconnect the turbine from the system
following a system fault. Thus, until a few years ago most (if not all) wind turbine

16

manufacturers had some form of protection to disconnect the turbine from the system
following a system fault. As an example, for conventional induction generators it is
typical for the units to disconnect from the system if the voltage at the terminals of the
machine falls below 70 to 80% for more than 100 ms. On most transmission systems
this would mean a high probability that the wind turbine would disconnect for any
nearby system fault.
In the case of conventional induction generators, the reason for this tripping is that
upon fault inception the electrical power falls to a low value (zero for a bolted 3-phase
fault). This results in the turbine speeding-up. Thus, if the turbine is not disconnected
after a certain time it may exceed its pullout torque and become unstable (similar to
an induction motor stalling, except here the units speed increases uncontrollably since
it is running as a generator). This is depicted graphically in Figure 2-8. We start at
the steady-state operating point A. Upon inception of the fault/disturbance, due to the
effective change in the units electrical torque speed curve (as a result of voltage
depression) we move to point B. Now the unit starts to accelerate because mechanical
torque is greater than electrical torque, and thus may go to point C at which point the
unit could go unstable, if the electrical torque upon fault clearing is less than the
turbines mechanical torque (and is on the left side of the torque-speed curve beyond
the break-down or sometime called pull-out torque the peak of the curve).
Therefore, under voltage protection is installed to disconnect the unit from the system
for low-voltage conditions.

Figure 2-8: Behavior of a conventional induction generator during a disturbance.

Most wind turbine manufacturers now offer a low voltage ride-through package. This
was originally driven by European requirements (discussed in section 3.2). This is
achieved by a combination of modified blade pitch control algorithms (that help to
remove the mechanical power following fault inception) and applying uninterruptible
power supplies (UPS) at the turbine to keep the controls running during the fault. In
addition, dynamic reactive power sources such as an SVC or STATCOM may

17

sometimes be required to provide additional support upon fault clearing to ensure


proper voltage recovery.
The largest suppliers of conventional wind turbine generators are Vestas, Bonus (now
Siemens), Made, Nordex and Mitsubishi.
2.3.3 Doubly-Fed Induction Generators
One of the most common variable speed turbines is the doubly-fed induction
generator (DFIG). One design is shown in Figure 2-4. This design employs a series
voltage-source converter to feed the wound rotor of the machine. By operating the
rotor circuit at a variable AC frequency one is able to control the mechanical speed of
the machine. In this design the net power out of the machine is a combination of the
power coming out of the machines stator and that from the rotor (through the
converter) into the system. When the unit is operating at supersynchronous speeds
real power is injected from the rotor, through the converter, into the system. When
the unit is operating at subsynchronous speeds real power is absorbed from the
system, through the converter, by the rotor. At synchronous speed, the voltage on the
rotor is essentially dc and there is no significant net power interchange between the
rotor and the system. Most designs tend to supply reactive power to the system
through the machines stator by effectively changing the d-axis excitation on the rotor.
A vector control strategy is used, where the rotor current is split into d-axis (flux
producing) and q-axis (torque producing) components. Each component is then
controlled separately. The d-axis component is controlled in order to regulate the
machine power factor (i.e. effectively controlling the reactive output of the machine).
The q-axis component is controlled in order to keep the electrical torque of the
machine constant. It is possible, however, to also provide reactive power through the
converter with a 4-quadrant voltage source converter design, and this is presently
done by some manufacturers. The reason for not typically utilizing this feature is that
the back-to-back converter is typically rated at roughly 25 to 30% of the full rating of
the machine (i.e. approximately = maximum slip x machine rated power). Thus, by
utilizing this feature one can only achieve a fraction of the total rating of the machine
in reactive output. None-the-less, by employing this feature in the line side converter
the wind turbine generator line side converter can essentially act as a STATCOM and
supply or absorb reactive power to or from the system even when the actual wind
turbine generator is not running and disconnected from the system. Providing this
feature, though, will typically mean additional cost. The fact that currents are tightly
controlled (with loop speeds typically ranging well into the kilohertz), means that, for
example, controls have the ability to, within limits, hold electrical torque constant (as
opposed to the relation between torque and angle in synchronous machines). Thus,
rapid fluctuations in mechanical power can be temporarily stored as kinetic energy,
thus improving power quality (eventually, however, outer control loops will modify
current orders so as to restore speeds to their optimum setting).
As in the case of conventional induction generation, older designs of DFIGs would
disconnect from the system during a close in fault. In fact, in the case of earlier DFIG
designs, one might say they were more sensitive to system fault and would disconnect
from the system in a much shorter time frame than conventional induction generators
(within milliseconds if the system voltage dropped below 70%). Unlike the case of
conventional induction generation, however, the process leading to separation might

18

not be readily apparent from dynamic simulation results. The concern in DFIG is
usually the fact that large disturbances will, as in the case of synchronous generation,
lead to large initial fault currents, both at the stator, and, due to the laws of flux
conservation, at the rotor as well. These high initial currents will, of course, flow
through the rotor-side converter, which is a concern. Furthermore, this initial surge
following the fault includes a rush of power from the rotor terminals towards the
converter. Due to low voltages at machine terminals during a disturbance, the statorside converter is limited in its ability to pass power to the grid. Consequently, the
additional energy goes into charging the dc bus capacitor and thus dc bus voltage rises
rapidly, depending on the design of the converter controls. This may give rise to
protection acting to short-circuit the capacitor (via a crowbar) in order to protect the
converter power electronic components (see Figure 2-9) [21]. In the past, when the
crowbar circuit fired, the unit would be disconnected from the system.
The newer generations of doubly-fed induction generators are now supplying lowvoltage ride-through [22], achieved through changing the control and protection
philosophy of the voltage-source converter. ABB has successfully demonstrated the
ride-through capability with a 2 MW doubly-fed test set up for the European market
[23]. Some of the subtleties of the low-voltage ride-through strategies are discussed
in section 4.1.1, in the context of modelling.
The largest manufacturers of DFIG units are Vestas, GE, Gamesa and Nordex.
Although, as explained in [15], due to patent right issues only GE supplies DFIG units
in the USA market.

Fault

Rotor
Side
Converter

Stator
Side
Converter

Figure 2-9: Tripping of old design DFIG.

2.3.4 Other Designs


2.3.4.1 Full Converter Units (Synchronous generators)
Presently, the major supplier of full converter units is Enercon. This concept is shown
in Figures 2-5 and 2-6. Other manufacturers are starting to pursue units of this
design. These units have dominated the German market, with perhaps three quarters
of the wind turbine generators in Germany being of this design from Enercon.

19

The concept in this case is to generate power using either a conventional generator7
with a dc field or a permanent magnet generator. This has two basic advantages:
1. It allows for a gearless design. This avoids the mechanical complexity of
gears and hydraulics. That is, the generator can be directly coupled with the
turbine and may spin at whatever rotational velocity as required. The
frequency of the electrical output of the generator is then converted by a backto-back frequency converter to the grid frequency (50 or 60 Hz). However,
the gearless design typically means that the generator has a significantly larger
diameter to accommodate a large number of pole pairs and thus requires a
more spacious nacelle (e.g. the Enercon E112, which is a 4.5 MW unit, has a
generator with 84 poles [24]).
2. Through the use of a frequency converter the full electrical output of the
generator can be converted from a wide range of frequencies to grid
frequency. This means that the wind turbine generator may operate at a wide
range of speeds thus once again providing the benefits of a variable speed
drive unit (see section 2.3.1.3).
In addition, with the used of a voltage-source converter the grid side converter (or
sometimes referred to as inverter) can independently control real and reactive power.
In this way the electrical grid and the generator are decoupled. The features allows
for greater flexibility and easier control for providing:
1) low-voltage ride through, and
2) voltage regulation and reactive power control at each turbine.
To achieve low-voltage ride-through the line side converter (or inverter) can stop
gating the IGBTs if the voltage falls to excessively low levels and be essentially on
stand-by to re-start once the fault clears. On the generator side, the converter (or
rectifier) can be by-passed and the stator made to feed into a braking resistor to
prevent excessive over speed. In addition, since the generator does not directly see
the low network voltages during such an event there is no large transient rotor or
stator currents produce in the machine. Voltage regulation is easily achieved with a
voltage-source converter through controlling the relative phase and magnitude of the
voltage phasor produced by the voltage-source converter as compared to the grid
voltage phasor. This concept is no different than that already used in STATCOM and
voltage-source HVDC systems.
2.3.4.2 The Vestas Opti-Slip Design
Another variable speed design concept is one that is primarily produced by one
manufacturer, Vestas. Vestas refers to this design as Opti-Slip. In this design the
generating unit again is a conventional induction generator with one modification, the
rotor is fitted with what surmounts to essentially a variable external resistance. This
resistance is controlled using power electronics. By varying the rotor resistance one
can essentially flatten out the torque-speed curve of the unit and thus allow stable
operation of the conventional induction generator for a great range of speeds. Figure
2-10 shows the potential change in the units torque-speed curve. As shown in the
7

The commonly used term is to say that a synchronous generator is used since the generator design
is similar to conventional fossil fuel plants. That is, the rotating magnetic field generated by
mechanically rotating a dc field on the rotor induces current in the stator that can be extracted as useful
electrical power. However, in this report we will refer to these units as conventional generators since
the term synchronous generator is a misnomer for it implies that the generator is rotating in
synchronism with the system frequency, which is not true.

20

figure, at 1 per unit electrical torque the unit can operate between roughly 2% to 10%
slip. This is quite a larger range than a standard convention induction generator,
which typically sees a change in speed of only a fraction of a percent from no-load to
full load. However, the speed range is still smaller than other variable speed designs
such as the doubly-fed unit and the conventional (synchronous generator or PMG)
with a full-converter. These latter units can have operating speed ranges from -30 to
+30% slip. Typically, the turbine is operated between 2% to 5% slip. Then in the
event of a short wind gust, the turbine can still increase its speed up to say 10% slip
and thus absorb some of the additional energy in the form of stored energy in the
shaft. In this way this design, much like other variable speed designs, can offer better
power quality with lesser short term power fluctuations that may lead to flicker.

Figure 2-10: Variation in the torque-speed curve of a variable rotor-resistance induction


generator.

These variable rotor-resistance units are still essentially an induction generator with
no field excitation. Thus, they absorb reactive power from the system and required
additional shunt compensation in the form of switched capacitor banks and/or static
var systems to compensated for the reactive power consumption at full-load. These
units can also be supplied with low-voltage ride-through. Similar to conventional
induction generators this is achieved through a combination of modified blade pitch
control algorithms (that help to remove the mechanical power following fault
inception) and applying uninterruptible power supplies (UPS) at the turbine to keep
the controls running during the fault. In addition, dynamic reactive power sources
such as an SVC or STATCOM may sometimes be required to provide additional

21

support upon fault clearing to ensure proper voltage recovery. In addition, the rotor
current transient during severe voltage dips is often controlled to protect the power
electronics in the rotor circuit by switching the rotor resistance to its maximum value.
This has the added benefit of flattening out the machines torque-speed curve and thus
making it less likely to go unstable during a disturbance. However, capturing this
phenomenon in a model can be challenging (see section 4.1.1 for further discussion).

22

3 A Survey of Planning and Interconnection Standards


Worldwide
The bulk of the installed capacity and experience with wind turbine generation has
been primarily in Europe and the US. In addition, three European nations Denmark,
Germany and Spain together account for almost 80% of the European installed
capacity of wind generation. It is thus not surprising that Denmark and Germany
have led the way in terms of setting standards for planning and interconnection of
wind farms to utility grid systems. Ireland recently has followed this trend with
putting out a number of comprehensive documents on wind interconnection standards,
in 2004. In North America, the Alberta Electric Systems Operator (AESO) in Canada
was the first region to adopt an explicit wind generation interconnection standard that
was published in November, 2004. The AESO standard was developed based on
input from stakeholders and many months of review. One of the inputs to the process
was a study, similar to this, performed by ABB for the AESO.
Though there are quite well developed planning and interconnection standards in the
North American continent, to our knowledge, only a few (such as AESO and Hydro
Qubec) explicitly address the interconnection of wind generation. The American
Wind Energy Associations (AWEA) also recently made a proposal to FERC
(discussed further below) to implement a national standard for wind generation
interconnection. Based on this proposal from AWEA, FERC has proposed a set of
interconnection standards for wind generation [74]. This document is discussed
below. The North American Electric Reliability Council (NERC) has also recently
commissioned a new wind generator task force to look at reliability issues specific to
wind generation. The task force was started in February 2005 and has the goal of
making recommendations some time by the end of 2005.
The International Electrotechnical Committee (IEC) has also published a number of
standards for design and specification of wind turbines [25, 26 and 27]. These
standards, however, are more aimed at the turbine manufacturers for specifying
design standards and techniques for measuring and demonstrating performance
criteria. The Institute of Electrical and Electronic Engineers (IEEE) recently
published a standard (IEEE Std 1547) on the interconnection of distributed resource.
This standard is for generating facilities with an aggregated total capacity of 10 MVA
or less. Thus, this document may be applied to small wind farms.
The following subsections will provide a summary of these various sources with a
particular focus on issues relating to planning for and interconnection of wind energy
systems into a transmission system.
3.1

North American Continent

The various regions of the North American continent with significant wind generation
penetration are listed below in alphabetical order and standards that exist in each
region discussed. (Note: AWEA is not a region of the Continent but rather a major
wind industry organization in the North American continent).

23

3.1.1 Alberta
The AESO standard [28] was developed based on input from stakeholders and many
months of review. One of the inputs to the process was a study, similar to this. The
key features of this standard are summarized in the following paragraphs.
1. Voltage Ride Through Requirements: Wind farms are required to be able to
operate continuously between 90% to 110% of nominal voltage at the point of
connection8. Furthermore, the wind farm should be able to stay connected to
the system for voltage dips or post-transient voltage rises shown in Figure 3-1.
The low-voltage ride-through capability depicted here is the same as the E.ON
standard (discussed in section 3.2.2). This requirement is applicable to all
wind farms with an aggregated megawatt capacity of 5 MW. The standard
also indicates that the AESO will consider on a case by case basis wind farms
that may not fully meet the voltage ride-through requirement until January 1,
2006. After that date all further wind farms must fully comply with the
standard. Clearly, the wind farm is allowed to trip for faults internal to the
farm and transmission faults on radial transmission lines connecting the farm
to the grid.
2. Voltage Control and Reactive Power Capability: The AESO standard gives
quite specific details on the reactive capability of a wind farm and its ability to
regulate voltage. Reactive capability is measured at the low voltage side of
the wind farm substation. At this point, when at full load, the farm should be
capable of 0.90 pf lagging (capacitive) and 0.95 pf leading (inductive). Also,
the wind farm is required to have an automatic voltage regulation system that
is able to regulate the voltage at this point to a specified set-point to within +/0.5%. The voltage regulation system should have adjustable gain, droop and a
reference set-point. Each wind farm should have a wind farm operator
available 24/7 for communication with the Transmission System Operator
(TSO). The TSO may from time to time instruct the wind farm operator to
adjust the wind farm voltage reference set-point and or droop. Finally, the
standard defines an expected portion of this reactive power capability (+0.95
to -0.985 pf) to be dynamic and continuously variable, while the remaining
portion may be achieved using discrete elements such as shunt capacitors.
The standard also states that short term reactive power capability for 1 second
can qualify for the dynamic range.
3. Off Nominal Frequency Operating Requirements: The wind farm is
required to meet the following off-nominal frequency operating conditions.
>61.7 Hz
0 seconds
61.6 Hz to 61.7 Hz
30 seconds
60.6 Hz to <61.6 Hz
3 minutes
>59.4 Hz to <60.6 Hz Continuous Operation
>58.4 Hz to 59.4 Hz
3 minutes
>57.8 Hz to 58.4 Hz
30 seconds
8

In the AESO document point of connection is defined as the highest voltage point at which electric
energy is transferred between the customers facility and the Alberta Transmission System. A Point of
Connection may be a Point of Supply (POS), Point of Delivery (POD) or both.

24

4.

5.

6.

7.

8.

9.

>57.3 Hz to 57.8 Hz
7.5 seconds
>57.0 Hz to 57.3 Hz
45 cycles
57.0 Hz or Less
0 seconds
Supplemental Over Frequency Control and Other Operating
Requirements: The AESO has temporarily removed the requirement for
some of these aspects, pending the outcome of a wind variability study the
results of which are to be published near July 2005. These requirements
concern ramp rates on the wind turbine generators and other functionalities
such as the ability to curtail power automatically for over-frequency grid
conditions (similar to governor action, however, only for over-frequency
conditions).
Modeling and Model Validation: The AESO requires that models, in PSS/E
format (user written if necessary), be submitted to AESO and the AESO must
be permitted to share the models with the WECC. The wind power facility is
then required to provide the AESO with power system studies that
demonstrate the facility will meet the requirements of the standard with
respect to voltage ride-through and voltage regulation.
Commissioning Tests: The standard specifies a set of tests to be performed
during commissioning of the wind farm to illustrate the performance of the
equipment and that it meets the specifications of the standard. For example,
tests such as a step-response to the wind farm voltage regulation system (or
SVC/STATCOM), etc.
Power Quality, Harmonic and Resonance Issues: The standard requires
that all equipment meet the applicable IEEE (or IEC) standards on power
quality and harmonics and that the collector system be appropriately designed
to avoid resonance issues such as ferroresonance on unit transformers,
harmonic resonances that may be caused by shunt capacitor application and
self-excitation. All reputable equipment manufacturers will by default meet
the applicable IEEE and IEC standards on harmonics and power quality. As
for resonance issues, this is a matter of prudent system design.
Monitoring Requirements: The standard states that the wind farm may be
required to make provisions for the installation of a system disturbance
monitor that complies with the AESO Requirements for Phasor Measurement
Units (PMU). The AESO will make such determination, and if required,
request the transmission facility owner to coordinate with the wind power
facility owner the installation of the PMU. The PMU, if required, must
monitor 3-phase voltages and currents as well as frequency.
Equipment Standards: A number of specifications are given on preferred
winding configurations on the substation transformers, types of
under/overfrequency relays, grounding, lightning protection, revenue metering
etc. These specifications may vary slightly from utility to utility depending on
preferred approach and adopted design protocols of the utility.

25

Figure 3-1: Voltage ride through capability requirements for AESO.

3.1.2 American Wind Energy Association Proposal to the Federal Energy


Regulatory Commission
In mid 2004, the American Wind Energy Association (AWEA) issued a document
entitled Standardizing Generator Interconnection Agreements and Procedures
Doket No. RM02-1-001 [29]. This document was presented before the FERC on
September 24, 2004 [30]. Various experts and wind industry stakeholders attended
the meeting to comment on the document [30].
One of the primary motivations for the document is best summarized in the
introduction of the paper, which states:
Wind turbine manufacturers cannot meet the projected demand for their product if
they are required to meet a patchwork quilt of inconsistent standards imposed by each
grid operator.[29]
The paper presents the following requirements to be established as a US standard for
interconnection of wind farms:
1. That the requirement for Power System Stabilizers not apply to wind
generators.
2. Low voltage ride-through capability: The document proposes adopting the
E.ON low voltage ride-through standard for non-synchronous wind generators,
with the requirement being applied to voltage measurements at the point of
interconnection. (see section 3.2.2 for the E.ON low voltage ride-through
curve).
3. Telecommunication Equipment:
The document states the remote
supervisory control and data acquisition (SCADA) should be installed to
achieve the following:

26

a. Limitation of maximum plant output during system emergency and


system contingency events,
b. Bi-directional electronic communication between the system operator
and the wind facility to facilitate scheduling and forecasting
information exchange.
The document states, however, that interconnection of wind farms should be
allowed without such SCADA capabilities where this capability is not
necessary to meet reliability standards.
4. Reactive Power: The document states that based on the generator
interconnection study, non-synchronous wind generators shall be required to
maintain a composite power delivery at rated power output at the point of
interconnection at a power factor ranging from 0.95 leading to 0.95 lagging.
However, the document goes further to say that this requirement should not be
enforced if the generator interconnection study shows that a power factor
range of less than 0.95 lagging (capacitive) can meet the applicable reliability
standards. In addition, the document states that for added flexibility of design
the good utility practice shall be used to locate the required reactive support
either at the wind turbine(s) or throughout the wind plant medium voltage
collection system.
5. Models and Self-study: The document asserts that For wind plants, the
turbine selection and the electrical design of the wind plant is an output of, not
an input to, the Feasibility Study. Due to the very fast project development
cycle for wind generation, the particular make and model of wind turbine has
not normally been selected at the time an interconnection request is
submitted. Thus, it is suggested in the document that a wind developer be
allowed to enter the queue, pay the applicable deposit for doing so, then obtain
system data from the utility to perform self-study of the wind farm and then
present to the utility an electrical design and wind turbine/plant models
appropriate for the transmission provider to conduct the system impact study.
On the above recommendations, items 4 and 5 were perhaps the ones most debated
[30]. It was felt by a number of transmission providers at the conference that the
power factor requirements should be enforced and no exceptions made to the rule.
The issue of model availability is also a concern. Presently, there are IEEE and
CIGRE Task Forces working on this particular concern. One other item discussed
during the conference was how these standards are applied. It was generally agreed
that should such standards come into effect, they should not be made retroactive since
some aspects (particularly low voltage ride-through) cannot necessarily be applied to
old installations.
3.1.3 California
Despite the major influx of wind generation in California, there appear to be no direct
publications (that are publicly available) specifically addressing planning and/or
interconnection standards and requirements specific to wind turbine generators. In
principle much of the existing standards in the form of the WECC/NERC planning
standards have been applied [31]. It is interesting to note that almost half of the
installed capacity of wind generation in California to date is composed of wind farms
ranging from less than 1 megawatt to 50 megawatts.

27

One of the major transmission owners in California, Southern California Edison (


SCE), had this to say at a recent Conference held by the Federal Energy Regulatory
Commission (FERC):
.SCE has required that wind generation developers provide external reactive
resources within their wind park facilities to accommodate their own reactive demand
and compensate for losses within their own collector systems, and to be able to
deliver a 0.95 power factor lead at the point of delivery. .. Power factor
requirements should be imposed on the installations that are the same or equivalent to
the requirements for synchronous generator installations and for all generator
installations.[30]
As can be seen, though there is no specific wind interconnection standard in
California, the trend appears to be one of imposing the same or similar standards on
wind generation as is being applied to conventional synchronous generators.
3.1.4 Federal Energy Regulatory Commission Proposal
On January 24, 2005 FERC released a proposal (Appendix G) Interconnection
Requirements for Wind Generators to be included in the Large Generator
Interconnection Agreement [74]. In summary, FERCs proposed standard states the
following:
1. That wind generating facilities shall demonstrate the ability to remain online
for voltage disturbances up to the periods of time associated with a proposed
low-voltage ride-through capability curve. The curve is to be applied at the
point of interconnection, which is defined as the high voltage side of the wind
farm substation transformer. The low-voltage ride-through curve defined is
essentially that proposed by AWEA, which was adopted from the E.ON
standard (Figure 3-3).
2. The wind plant shall provide SCADA capability to transmit data and receive
instructions from the Transmission Provider. The Transmission Provider and
the wind plant Interconnection Customer shall determine what SCADA
information is essential for the proposed wind plant, taking into account the
size of the plant, its characteristics, location, and importance in maintaining
generation resource adequacy and transmission system reliability in its area.
[74]
3. That the wind farm maintains a power factor range of +/- 0.95 as measured at
the high side of the wind farm substation transformer.
FERC indicates that the transmission service provider may waive any or all of the
above requirements for a given wind farm, but this must be done on a comparable
and not unduly discriminatory basis for all wind plants.
3.1.5 Hydro-Qubec
Hydro-Qubec recent also issued a supplement to its already existing interconnection
requirements document that specifically addresses the requirements on
interconnection of wind farms [32]. The following is a summary of the document.

28

1. Voltage Regulation: Wind farms are required to regulate voltage with an


automatic voltage regulating control system. However, a distinction is made
between the various technologies. For synchronous generators (full converter
units) it is required that the unit be able to produce/absorb reactive power
when at rated real power at up to a power factor of +/- 0.95. For asynchronous
units, the same requirement is imposed on doubly-fed induction generators.
However, asynchronous induction generators that do not have an inherent
reactive capability are required only to maintain a power factor of unity at the
point of interconnection (Note: based on the wording of the document, the
other pf requirements apply at the turbine and not the point of
interconnection). Finally, if enough shunt compensation cannot be provided to
establish unity power factor at the point of interconnection for this last
category of units (e.g. to avoid self-excitation or excessive overvoltages on the
collector system) then the addition of other reactive compensation devices at a
more appropriate point on the Hydro-Qubec system needs to be considered.
2. Off-Nominal Voltage and Frequency Protection: Essentially what amounts
to a voltage ride-through specification is given. This is shown in Figure 3-2.
The off-nominal frequency operating regime is given in Table 3-1.
3. Model Requirements: Prior to the integration study, detailed models for the
wind turbines and their controls must be supplied to Hydro Qubec in
PSS/ETM format.

Figure 3-2: Voltage ride through capability requirements for Hydro Qubec.
Table 3-1: Frequency operating point for Hydro Qubec.
Frequency (Hz)
Duration
f < 55.5
Instantaneous trip
55.5 f < 56.5
0.35 s
56.5 f < 57.0
2s
57.0 f < 57.5
10 s
57.5 f < 58.5
1.5 min
58.5 f < 59.4
11 min
59.4 f 60.6
Continuous
60.6 < f 61.5
11 min
61.5 < f < 61.7
1.5 min
f 61.7
Instantaneous trip

29

The interesting contrasting point between this standard and the others discussed below
is that in this case the wind turbine generators (WTG) are required to ride through a
disturbance that results in zero voltage for up to 100 ms, while all other standards
discussed in this report go down to no less than 15% - in addition, the Hydro Qubec
standard is not explicit about where this voltage is measured so one can assume that it
is at the terminals of the WTG. Also, the standard treats power electronic based
WTGs and non-power electronic based WTGs differently.
3.1.6 Midwest
The Midwest is one of the areas in the US with fast growing wind generation assets.
Once again there have been no specific changes or supplements made to the planning
and interconnection standards applied in the Midwest that are specific to wind
generation [30, 33].
One Midwest utility that does presently make comments specific to wind turbine
generation in its interconnection requirements is Xcel Energy. They require that the
wind turbines be able to ride through voltage swings down to 0.7 pu (for up to 0.5
seconds) and swings back up to 1.2 pu. In addition, voltages may stay below 0.9 pu
for up to 2 seconds; the wind turbine must be able to tolerate this also [34].
3.1.7 New Mexico
FPL Energy recently developed one of the largest wind farms in the country in New
Mexico, in the Public Service Company of New Mexico (PNM) transmission system
[35]. This project exemplifies some of the challenges of introducing wind generation
into a utility grid. A prime example is the record time in which the project was
established. With the system studies completed the project transitioned from concept
to reality in February 2003, the substation was built by July 2003 and the wind
turbines placed in service by the end of 2003. The farm has a peak capacity of 204
MW. This shows the typical short lead time required for placing a wind farm inservice as compared to other more conventional generating facilities, not to mention
transmission upgrades which may at times take several years.
Another interesting aspect of this site was the vicinity of the Blackwater HVDC
station. This raised concerns with regards to interactions between the HVDC and the
wind turbine generators. Thus extensive studies were performed by both ABB (the
HVDC manufacturer) and GE (the wind turbine manufacturer) in collaboration with
PNM to assess the potential for negative interactions and to identify mitigation
options. No significant negative interactions were identified.
3.1.8 New York
In September of 2003, the New York State Energy Research and Development
Authority (NYSERDA) in conjunction with the New York ISO (NYISO)
commissioned a study to evaluate the impact of wind generation on the New York
power systems (www.nyserda.org). At the time of the initial submittal of this report,
only the Phase 1 report for this study had been released publicly [36]. The Phase 2
study report was released in February 2005 and can also be found at the NYSERDA
website. The conclusion of the quite detailed seventeen month study is that the NY
30

bulk power system can reliably accommodate at least 10% penetration of wind
generation (3300 MW) with only minor adjustments to its existing planning, operation
and reliability practices. One of the key assumptions of the report is that this is done
with state-of-the-art technology that includes LVRT, reactive power capabilities and
voltage regulation at each wind farm.
3.1.9 Texas
In 2002, the Electric Reliability Council of Texas (ERCOT) commissioned a project
for developing better stability models for wind turbine generation technologies [37].
The intent was to develop more suitable models for modeling wind generation in
planning and other system studies. Presently, there have been no public documents
published on this project. ERCOT, similar to many other US states, do not have an
explicit wind interconnection standard.
3.1.10 Western Electricity Coordinating Council (WECC)
In early 2003 a white paper was circulated within WECC [38]. The main item for
discussion in this paper is the need for a low-voltage ride-through standard much like
those in Europe (the European standards are discussed in the next section). Since then
utilities within WECC, particularly the WECC Modeling Working Group, have been
continually discussing the subject of wind generation interconnection. At present this
discussion has culminated into the formation of the WECC Generator Electrical Grid
Fault Ride Through Capability [39] document, which was posted on the WECC
website (www.wecc.biz) in October 2004 and was open for comment by WECC
members until December 21, 2004. The document states that all generation facilities
10 MVA and above must remain connected to the grid for normal clearing of a
transmission system fault (3-phase, single-line to ground or any other type of multiphase fault), provided clearing of the fault does not island the generation facility. The
key change in the document is that this standard is proposed to be applied at the point
of interconnection. The WECC Modeling Working Group is also presently working
on guidelines for field testing for the purpose of simulation model validation of wind
turbine generators.

31

3.2

Europe

3.2.1 Denmark
Denmark has perhaps the greatest concentration of wind generation in a single
system, in the world. Spot penetrations are as high as 70% [40], while wind
generation supplies nearly 18% of the nations energy needs [8]. It is thus not
surprising that the first technical specification for connecting wind farms to a
transmission grid was developed in Denmark by the Danish Transmission System
Operator, Eltra [41]. The highlights of that document [41], are summarized here:
1. The specification applies to wind farms that are to be connected to the
transmission network, i.e. voltages at and above 100 kV. In addition, a wind
farm is regarded as connected to the transmission network when no customers
are connected to the network between the wind farm and the transmission
network and thus the bulk of the wind farm power is being injected into the
transmission grid.
2. The purpose of the interconnection standard is to ensure certain abilities in
the wind farm for use at the operation of the power system. [41].
3. The document stresses that the rate of expansion and amount of wind energy
penetration in the Danish system is unparalleled by any other in the world.
Thus, the document has been developed solely on the basis of experience
within the Danish system and will continue to be a work in progress as newer
wind generation technologies and techniques are developed and greater
experience is gained.
4. Power and Power Control: The Danish system has wind farms that are both
land-based and offshore [42]. The significance of this is that land-based
turbines tend to have lower kW capacity (per turbine) and are more
geographically wide spread. Therefore, land-based turbines tend rarely to
achieve full production simultaneously or conversely, rarely will they all come
to a stand still at the same time. In contrast, off-shore wind turbines will
operate at full capacity more often and wind gusts at sea can result in sudden
shut-down of large portions if not all turbines in a single wind farm. Thus,
one of the requirements in the Danish standard is that high wind speed must
not cause all turbines in a single farm to stop simultaneously. This can be
ensured either by geographical spread or by control implementations that
ensure staging both during start-up and shut-down.
5. Production Control: Each wind farm is requested to have a specified
production limit, i.e. a certain MW value that it cannot exceed. This limit
shall be controllable by a single central signal. The control shall take place at
the individual turbine, and it shall be possible to control the production so fast
that it can be reduced to below 20% of the maximum in less than two
seconds. [41]
6. A slow MW limit control signal is given to wind farms in the event of
disconnected lines in the network where flows need to be controlled to
observed line thermal limits; such control occurs within the time frame of line
thermal time constants, approximately 15 minutes.
7. Low-voltage Ride-Through: The Danish standard does not have a low-voltage
ride-through capability curve requirement like the E.ON and REE (Spanish)
requirements (see following subsections). However, the standard does make

32

explicit statements with regard to the capability of wind turbines to withstand


and ride through faults by:
a. Automatic fast reduction (order of magnitude of seconds) of the
mechanical power on the turbine in order to maintain stability during
and immediately after a fault. In addition, the standard states that the
control shall take place so that a controlled, complete power reduction
and power increase is completed within about 30 seconds.
b. A limit on the rate of MW increase.
These requirements are essentially in line with the low-voltage ride-through
capability and implementations described in section 2.3.2.
8. Start-up: The wind farm shall communicate signals of the status of the farm,
e.g. whether units have stopped due to lack of wind or too high a wind speed,
due to forced outage, etc. All turbines in a wind farm should not start
simultaneously, but should be staggered either intentionally by controls or due
to geographic spread.
9. Voltage and Frequency Limits: Limits of operation are given for steady-state
operation at off-nominal frequency and voltage levels. Such requirements are
specific to each utility. For example, the European system is a 50 Hz system
while Alberta is 60 Hz.
10. Reactive Power Consumption: The Danish standard requires that the wind
farm power factor, at the point of common coupling (i.e. interconnection
point) be reactive power neutral, that is unity power factor. This is to be
achieved through the action of all reactive control devices including on-load
tap changers on the substation transformer, switchable shunt capacitor banks
at the turbine and on the collector system and/or other VAr control
mechanisms (such as the VAr capability of doubly-fed induction generator or
gearless machines driven by a four-quadrant frequency converter, see section
2.3).
In actuality, the document states that reactive power
consumption/production of the farm may vary up to 10% of the maximum
power output of the farm in steady state; that is, the farms power factor may
be in the range of 0.995 leading to 0.995 lagging.
11. Power Quality: Limits are specified for voltage flicker, Harmonics, telephone
interference and other telecommunication interference. These are all based on
IEC standards.
12. Though no specifics are given, the document does indicate that in the case of
wind farms connected through HVDC-links there may be a risk of
interactions; no such negative interaction have been experienced though.
13. Stability Requirements: Requirements are given for performance of the wind
farm during and after a system fault. The requirements, however, do not apply
to radial connection. (As it is clearly expected that the wind farm would be
isolated and trip for a fault on a radial connection.) The requirements are that
the wind farm be able to withstand at least three faults within a two minute
time frame, without being disconnected. In addition, that this can be achieved
by fast reduction of the MW production from the farm during the fault and
subsequent increase in production to the original value within 30 seconds after
the fault has cleared. This is in essence a low-voltage ride-through
requirement (see sections 2.3).
14. More Extensive Faults: It is recognized in the document that more extensive
faults may result in an inability to maintain connection of the wind farms to

33

the system. Thus, it is required that the wind turbines in the farm trip for
prolonged low voltage conditions where system voltage does not recover.
15. Protection: The document states that the plant owner is responsible for
ensuring that adequate protection is in place to safeguard the turbines from
damage as a result of faults and other systems events. All protection,
however, must be coordinated with transmission protection.
16. Verification and Tests: The plant owner is given the responsibility to provide
the necessary models for system simulations. These models should be updated
upon commissioning if there are any substantially changes in the equipment
settings. Tests should be performed to verify the machines response,
particularly with respect to response to faults. The Danish specifications also
require that fault recording equipment be provided to record the response of
the entire farm and at least one turbine (of each type in the farm) to system
disturbances. The recording equipment should record up to 10 seconds prior
to the disturbance and 60 seconds after the disturbance. The following signals
are to be recorded:
a. For the entire farm
i. Voltage
ii. Active power
iii. Reactive power
iv. Frequency
v. Current
b. For a single turbine of each type in the farm
i. Rotating speed
ii. Active power
iii. Reactive power
iv. Voltage
3.2.2 Germany
The actual E.ON grid code is in the German language. An English translation of the
general code is given in [43]. A good summary of the new supplementary regulations
that specifically apply to wind generation is given in [44]. The main points of this
document are summarized here:
1. In the past connection of wind farms were based on the assumption that in
case of a grid disturbance, e.g. a severe voltage drop, the wind turbines would
disconnect from the system. In the present German system such a paradigm
may result in several thousand megawatts of wind generation tripping off-line,
which is clearly unacceptable.
2. Thus, one of the main new grid connection regulations as they pertain to wind
farms (based on the Germany experience and regulation in Denmark) is the
requirement of low-voltage ride-through. The E.ON requirement is shown in
Figure 3-3.
3. The wind farms must operate within specified frequency and voltage
bandwidths. For example, the turbines must operate between 47.5 Hz to 51.5
Hz, and must trip immediately outside this range. When frequency is between
50.25 to 51.5 Hz, the wind farms must reduce their power output to control
system frequency.
4. Wind farm cut-in: The cut-in power of the wind farm must not exceed 1.2
times the rated wind farm capacity. This essentially is a limitation on the
starting current of the wind turbines, as most wind turbines are started as

34

induction motors. Modern designs incorporated soft-start power electronics


and thus for most plants this should not be an issue.
5. Reactive power: It must be possible to operate the wind farm between 0.975 pf
(inductive) and 0.975 pf (capacitive). This applies to the whole farm, thus
transformer OLTC, shunt capacitor banks at the turbines, on the collector
system and at the substation as well as other devices may be utilized to
achieve this.
6. Active power output: After a loss of voltage, where the wind farm has been
tripped, as wind turbines come on-line the increase in power of the wind farm
must not exceed 10%/minute. This may be achieved by interleaving the wind
turbines so that they do not all connect back to the grid at the same time. In
addition, in exceptional cases (contingencies that may require generation
curtailment) E.ON reserves the right to limit the output of the wind farm or
completely curtail the farm.
Finally, E.ON requires that the various wind turbine manufacturers demonstrate that
they can meet the above requirements. However, it is stated that Many of these
requirements do not necessarily have to be verified at the wind farm itself, but can
also be measured on an individual turbine, for example during type testing. [44].
This is interpreted as meaning factory testing by the manufacturer.

Figure 3-3: German (E.ON) standard low-voltage ride-through capability requirement.

There is, however, a subtlety to the above E.ON low-voltage ride-through standard.
The grid code states that this curve is applicable to low symmetrical short-circuit
generators. That is, if for a normally cleared (cleared in 150 ms or less) 3-phase fault
the symmetrical component of the short-circuit contribution from the generating
facility is higher than twice its rated current for 150 ms, then the unit must ride
through the curve given in Figure 3-4 [43]. Figure 3-3 applies only when this criteria
is not met. Since full converter units (such as the Enercon machines that dominate the
German market) do not contribute significant short-circuit, and it is assumed that
other asynchronous designs will have their short circuit contributions decay to levels
less than twice their rated current within 150 ms, then Figure 3-3 is applied to wind

35

generation. If, however, a WTG contributes significant short-circuit current it would


be expected to ride-through voltage zero at the point of interconnection like all other
conventional generators under the E.ON standard.

Figure 3-4: German (E.ON) standard low-voltage ride-through capability requirement for
synchronous machines with high short-circuit capacity.

3.2.3 Ireland
Because of the rapid increase in proposals for wind farm facilities in Ireland and a
growing concern by the transmission system operator on operating and stability issues
related to wind generation, a moratorium was placed on wind farm interconnection in
Ireland based on advice of the transmission system operator ESB Nation Grid9. At
the time there were 166 MW of wind generation in the system with an additional 1031
MW of applications being processed [45]. The moratorium was put into place to
allow time to [45]:
1. Work on revising the grid and distribution codes to take into account the
special characteristics of wind generation.
2. Prepare a survey of the accepted connection offers to ascertain the realistic
level of offers.
3. Investigate issues regarding constraining wind farms.
4. Align transmission and distribution connection processes.
5. Prepare a plan for modeling of wind generation and its impact on the system to
be produced.
6. Establish a wind steering group chaired by the Commission to oversee the
progress of these issues.
9

Eirgrid is expected to take over operation of the Irish transmission grid from ESB Nation Grid soon
(www.eirgrid.com).

36

This moratorium was lifted by the Commission for Energy Regulation (CER) in July,
2004. The supplement to the grid code to address wind generation integration was
issued in July, 2004 [46]. The code stipulates the following:
1. Fault Ride Through: The wind farm should be able to ride through system
faults per the E.ON standard curve (Figure 3-3). In addition,
a. During the Transmission System Voltage dip the Wind Farm Power
Station shall provide Active Power in proportion to retained Voltage
and maximize reactive current to the Transmission System without
exceeding WTG limits. The maximization of reactive current shall
continue for at least 600 ms or until the Transmission System Voltage
recovers to within the normal operational range of the Transmission
System, whichever is the sooner;
b. The Wind Farm Power Station shall provide at least 90 % of its
maximum Available Active Power as quickly as the technology allows
and in any event within 1 second of the Transmission System Voltage
recovering to the normal operating range. [46]
2. Frequency Control: The standard stipulates the wind farms should have the
ability to control their real power output automatically as a function of system
frequency per Figure 3-5. The point on the figure (A through D) are specified
by the transmission system operator 60 days prior to commissioning. In
addition, the wind farm must remain in operations when:
a. System frequency is in the range of 49.5 Hz to 50.5 Hz.
b. Remain connected for 60 minutes within the range 47.5 Hz to 52.0 Hz.
c. Remain connected for 20 seconds within the range 47.0 Hz to 47.5 Hz
d. Remain connected during a rate of change of frequency of 0.5 Hz/s
e. No additional WTG shall be started while the frequency is 50.2 Hz.

Figure 3-5: Frequency response performance of wind farm per ESB National Grid code.

37

3. Reactive Power Requirements: The reactive power requirements for the


wind farm are specified per the graph below.

Figure 3-6: Reactive power requirements for wind farms.

4. Voltage Regulation: The standard requires that the wind farm be able to
regulate voltage at the point of interconnection (defined as the low voltage
side of the substation transformer). This is to be achieved through an
automatic voltage regulation system similar to an AVR. Specific performance
requirements are specified such as a response time of 1 second to voltage
steps, etc. [46].
5. Harmonics and Power Quality: The wind farms are required to meet the
applicable IEC standards for voltage flicker and power quality requirements.
6. Monitoring and Signal Communication: Details are given on the types of
signals required to be communicated between the wind farm and the
transmission system operator. These are classified into four groups:
a. Wind farm power station signals (such as voltage, real and reactive
power being generated, voltage set-point at point of interconnection).
b. Meteorological data (such as wind speed and direction, ambient
temperature and pressure).
c. Availability data (such as number of wind turbines off-line etc.)
d. Wind farm curtailment data this would be a signal sent by the
transmission system operator to the wind farm to request megawatt
curtailment.
e. Frequency response data these would be settings for the wind farm
frequency control system (settings A, B, C, D etc. in Figure 3-5).
Also, a signal should be available to turn the frequency control feature
of the wind farm on or off.

38

Although not explicitly discussed in the grid code, ESB National Grid is also working
with the wind turbine manufactures to obtain acceptable dynamic models for the
purpose of system studies. These models are being requested in PSS/ETM format. To
date none of the models have been fully validated [47].
Another interesting observation is that ESB Nation Grid surveyed data that it had
collected over the years of actual faults recorded on the system10. As expected, as one
moves towards the transmission system the majority of faults tend to be single-line to
ground (75% at 220 kV). None-the-less, a few faults were found (particularly multiphase) that resulted in voltages down to 5% or less, thus clearly violating the
proposed E.ON fault ride-through standard. As such, ESB Nation Grid has proposed
that presently they will accept the E.ON curve, however, once WTG technologies are
able to ride-through voltages down to lower levels the grid code may be accordingly
changes. This is an interesting observation since this had been pointed out during the
conference with FERC regarding the proposed AWEA standard, namely that faults
such as a close-in 3-phase fault (though rare) will mean that a wind farm meeting the
E.ON standard may still trip [30].
3.2.4 Spain
An English translation of the Spanish standard [49] could not be obtained. However,
the content is very much similar to those developed by Eltra and E.ON. The concerns
in Spain are similar to Denmark and E.ON. Figure 3-6 shows the Spanish equivalent
of the E.ON low-voltage ride-through requirement.

Figure 3-7: Spanish low-voltage ride-through capability requirement.

A recent paper [40] sheds light on the events in Spain with regard to the integration of
wind turbine generation into the Spanish system. Some of the key points are
highlighted here:
10

Discussion Document for the Review of Requirements for Wind Turbine Generators Under System
Fault Conditions Commonly Referred to as Fault Ride Through, www.eirgrid.com.

39

The Spanish grid operator Red Elctrica de Espaa (REE) says that improved
low-voltage ride-through capability (per Figure 3-7) and improved wind
production forecasts would both significantly reduce the amount of spinning
reserve requirements and yield better system security.
REE also argues that the Spanish system cannot tolerate more than
approximately 17% of the system capacity in wind generation by 2011. Wind
proponents argue that this is an under estimation based on other examples in
Europe such as Denmark where spot penetrations of wind generation are near
70% without flicker or other problems. REE argues that the Spanish system is
unlike Denmark in that it has a poor interconnection to its neighbors with only
3% import capability.
REE has also indicated that the Spanish weather system makes it particularly
difficult to predict wind conditions. Anticyclone conditions often last several
days resulting in low winds and cold weather in the winter, and low winds and
hot weather in the summer in both cases meaning minimum wind generation
capacity during peak load demands.
Presently, REE is investing in tools for better forecasting and prediction of wind
generation. In particular, the SIPREOLICO II project will aim at increasing the
resolution of wind forecasting systems to 15 minute time windows [50].

40

3.3

Rest of the World

Other regions in the world are taking an active interest in renewable energy,
particularly wind energy. As described in section 2, the most notable of these regions
is Australasia. Both in Australia [51] and New Zealand [52] some preliminary work
has started on setting standards for wind energy and other renewable sources,
however, there does not appear to be any mature planning or interconnection
standards explicitly applicable to wind generation. Two of the key concerns in
Australia are the intermittent nature of wind generation and the availability of good
models for system studies. For example, the proposed capacity of wind generation for
the state of South Australia is nearly equal to (if not more) than the present peak load
in that state. As such, there are significant concerns related to load following and
reserve management. The second concern relates to availability of good models,
which are not considered proprietary in nature, that can be readily used and shared
among market participants for studying the wind farms.
3.4

General Observations and Experience with Wind Generation


Studies in the US

ABB has performed a large number of studies related to interconnection of wind


turbines generators in the US and Europe. In summary, the key lessons learnt may be
summarized as follows:
Low-voltage ride-through:
This is the concept that has attracted the most attention. In our experience in
performing studies here in the US we have observed the same concerns as highlighted
in the Danish and German standards. That is, where wind generation has a relatively
high level of penetration in a given part of the transmission system, a single
transmission fault may result in the tripping of hundreds of megawatts of wind turbine
generation if not properly fitted with low-voltage ride-through capabilities. As
discussed in section 2.3, many of the manufacturers (certainly the major
manufacturers) have now working solutions for meeting the E.ON low-voltage ridethrough requirements. Therefore, moving forward it seems only prudent to request a
similar capability of all wind turbines when being connected to the transmission
system.
A side note on this subject is that although a low-voltage ride-through scheme will
allow wind turbines in a farm to ride-through a nearby transmission system
disturbance, this may not necessarily in itself ensure system stability. For example,
consider the following hypothetical scenario. We have many hundreds of megawatts
of wind generation serving a remote end of the transmission system. A nearby
transmission fault results in the tripping of one of say two parallel paths to that section
of the system. The wind generation rides through the fault and following the fault
clearing the wind generation increases, over several seconds, back to or near its
original pre-disturbance megawatt value. However, we now have a much weaker link
to that part of the system and if initially there was a significant level of power
interchange between that area and the rest of the transmission network, with the
41

weakened link there may not be enough dynamic VAr reserve in the vicinity of the
wind farms to maintain voltage stability. As an example, such dynamic requirements
might be provided by the application of a static VAr compensator (SVC), to regulate
transmission system voltage immediately after a severe disturbance thereby ensuring
fast and stable voltage recovery [21]. Thus, planning studies are still critical to ensure
that the system can be securely operated even with such ride through strategies.
Some recent studies by ABB have shown the potential for large portions of wind
generation tripping due to a single transmission fault, where low-voltage ride-through
is not applied.
Conversely, other studies conducted by ABB have indicated the potential for tripping
on overvoltage of DFIG-type units following critical disturbances on stability-limited
systems. At issue in those cases is the ability of DFIG voltage controls to quickly
reduce reactive power orders while voltages recover from the significant voltage
backswings (i.e., after faults have cleared) that are typical following such limiting
contingencies. The studies highlight the importance of in-depth understanding of
DFIG reactive power controllers principles of operation, their settings in the field,
and their modeling.
One of the key concerns in Spain relates to the potential for loss of significant
amounts of generation. Presently the installed capacity of wind generation in Spain is
close to 8000 MW for a system that has a peak load of roughly 41 GW. Also, since
most of the WTG installed on the system were put into service before the
development of low-voltage ride-through technology, most of the units trip when
system voltage falls below 0.85 pu (and some power electronic based units, when the
voltage falls below 90%). Thus, under light load conditions when the wind
generation can constitute an even greater portion of the load this poses a concern.
This is quite a significant concern since a fault could potentially result in a few
thousand megawatts of generation tripping and thus lead to severe frequency control
problems. The action being taken to remedy this is enforcing the E.ON like lowvoltage ride-through standard, among other things.
Coordination of capacitor bank switching:
As discussed in section 2.3, conventional induction generator machines will typically
have associated with them, at the turbine, a number of switched shunt capacitor banks
for power factor regulation of the individual machine. In addition, switched shunt
capacitor banks may also be provided on the collector system and/or at the point of
interconnection (in the substation). Typically, the shunt capacitors at the turbines are
switched to control machine power factor, while the shunt capacitors on the collector
system and at the substation level are switched by PLCs to regulate voltage. It is
important to ensure that:
1. The PLC controls associated with the shunt capacitor banks are properly
coordinated with the existing capacitors on the transmission system as well
as the capacitors at the turbines. Appropriate voltage deadband and delays
should be chosen for switching each bank.
2. When designing the collector system due consideration should be given to
the potential for harmonic resonances on the collector system due to the
42

capacitor banks, and the potential for voltage magnification on the shunt
capacitor banks near the wind turbine generators (at e.g. 600 V) when
switching higher voltage capacitor banks (on the collector system or at the
substation level). Thus, during the design of the wind farm electrical
system these and other equipment application issue should be reviewed to
ensure proper design and integrity of the entire wind farm electrical
system.
3. During a disturbance the megawatt output of the wind turbines will

decrease momentarily, due to suppressed system voltage. If the wind


turbine generator rides-through the disturbance the capacitors at the
turbine terminals should not switch out. If the turbine trips, however, so
must the capacitors connected at the turbine, since otherwise they may
cause excessively high temporary overvoltage on the collector system
when system voltage recovers. This again is a matter of proper
coordination of controls and protection.
HVDC and series capacitor interactions:
It is possible to have adverse interactions between transmission equipment such as
HVDC and series capacitors and nearby wind turbine generators [21, 41]. With series
compensation, there is a potential for self-excitation, while with HVDC there is a
potential for detrimental controls interaction between the HVDC and the wind turbine
controls, particularly with variable speed designs (such as doubly-fed induction
generators). These issues are discussed in some more detail in section 4.1.
The substation transformer:
One of the most common designs for wind farms, in the US, is to collect a group of
say ten or so wind turbine generators onto a single loop or radial underground cable
(typically 34.5 kV) and to then have three to four feeders which pick up all these
loops and connect to a common substation. Then a single substation transformer bank
steps up the power collected by this collector system on to the transmission
network. Due to the relatively low capacity factor of a wind farm (0.3 to 0.4) the
substation transformer is typically not rated at the full installed capacity of the farm.
For example, for a wind farm capable of generating a peak capacity of 100 MW, the
substation transformer may be rated at 50/67/90 MVA. Design calculations may
show that based on the expected amount of time the farm may spend at peak load
during the year, the additional 10% overload at peak load would not result in
significant loss of life for the transformer. This, of course, is an example. Actual
designs would be quite dependent on the site, expected capacity factor etc. An
important observation is that an on load tap-changer (OLTC) on the substation
transformer typically provides added flexibility for steady-state voltage regulation.

43

4 Planning Criteria and Study Approach


Based on the material presented in the previous sections, in this section we present a
proposed approach for planning studies related to wind farms.
4.1

Planning

The general approach to planning should be no different to the existing approach that
BCTC takes with applicable NERC/WECC and BCTC specific planning and
reliability criteria. However, when studying wind farms a particular emphasis needs
to be put on looking at voltage control and regulation and the potential for large
portions of wind generation tripping due to under (or over) voltage tripping strategies.
Analysis will be needed to ensure overall system performance and integrity. The
growing trend is to analyze wind generation in groups rather than one at a time, since
issues such as system voltage recovery and reactive support are better addressed when
studied system wide, rather than one plant at a time.
4.1.1 Modeling of WTG and Wind Farms for Steady-State and Dynamic
Studies:
4.1.1.1 Modeling Various Types of Wind Turbine Technologies
There are many types of wind turbine generators (WTG) that are produced by the
various wind turbine manufacturers. This was discussed in section 2.3. In essence
there are four major types, these are shown in Figure 4-1. Each type requires a
different means of modeling.
Modeling WTG for Steady-State Analysis
For power flow analysis ideally conventional and variable rotor resistance induction
generators should be modeled using the equivalent circuit of an induction machine.
However, the most commonly used simulation tools in the North American market
(GE PSLFTM and PTI PSS/ETM) do not have such functionality, so it is reasonable to
model the WTG as a P, Q bus with constant Q equal to the amount being absorbed at
the real power (P) level being studied. For example, typically conventional induction
generators have a power factor of 0.9 pf. Thus, a WTG operating at full-load of say
1.65 MW would be modeled as a P, Q bus with P = 1.65 MW and Q = -0.799 MVar.
In reality the reactive consumption of the units will change some if the terminal
voltage changes from nominal (based on the characteristics of the machine), however,
in the absence of a steady-state equivalent circuit model of the unit a constant Q
representation is perhaps adequate specially since per typical planning criteria the
system voltage should be maintain at +/- 5% of nominal voltage. Once the WTG has
been modeled, the shunt capacitors at the turbine compensating for the WTG reactive
consumption should also be explicitly modeled. Note: in the past some have made
the gross approximation of modeling these units as a P, Q bus at unity power factor.
This being because the unit's shunt compensation provides adequate reactive power to
fully compensate for the unit's reactive consumption. This, however, is not an
appropriate model since this is only true at 1.0 pu system voltage. As the voltage

44

starts to change from 1.0 pu the reactive output of the shunt capacitors will vary as the
square of voltage while that of the WTG will essentially remain constant. Thus,
modeling the unit as a unity power factor P, Q bus can give quite optimistic results
with regard to steady state voltage.
For doubly-fed induction generators (DFIG) and full converter units that have voltage
source converters, the units may be modeled as a P, V (i.e. specified megawatt level
and voltage) bus with appropriate VAr limits. This is because both these types of unit
have reactive capability. It should be noted, however, that the full converter design
becomes a constant current device (rather than a constant power device) when it
reaches its limit. As such, the actual VAr limit will change linearly with voltage once
the unit is at limit. To model this accurately the device needs to be modeled as a
constant current source once at limit in power flow. Such a feature is likely not easily
available in most power flow simulation programs (unless one changes the unit to a
constant current negative load at its limit) and so a P, Q bus representation is likely
adequate (i.e. P, V bus that has reached its Q limit).
WTG Type
Conventional IM

Table 4-1: Summary of power flow models


Model
Shunt
Compensation
P, Q bus; Q = Explicitly
model
const.1
shunt capacitor

Variable Rotor Resistance IM

P,Q bus; Q= const. 1

Doubly-Fed Induction Generator

P,V bus;
Qmin Q Qmax

Explicitly
model
shunt capacitor

Generator
Transformer
Typically 600/34.5
kV at 6% on
transformer rating
Typically 600/34.5
kV at 6% on
transformer rating
Typically 600/34.5
kV at 6% on
transformer rating

Inherent in machine
typical power factor
+/- 0.95 pf
(or
better)
Full Converter
P,V bus;
Typically, inherent Typically 600/34.5
Qmin Q Qmax 2
to inverter capability kV at 6% on
(+/- 0.95 pf or transformer rating
better)
1. The reactive consumption of an actual induction generator will vary as the voltage drops, but to
properly model this one would need an equivalent circuit power flow model of an induction
machine, which is not available in programs such as PSS/ETM.
2. Provided the converter is a voltage-source converter the unit will have reactive capability.
However, at its limit this is a constant current not a constant VAr device. Thus, once at limit a P,
V model is not exactly correct once again this may be a limitation of some simulation tools.

45

Figure 4-1: Summary of various turbine technologies.

Modeling WTG for Dynamic Analysis


There are a number of components that contribute to the dynamic behavior of a WTG.
These are:
Turbine aerodynamics
Turbine mechanical controls (i.e. pitch control or active-stall control)
Shaft dynamics
Generator electric characteristics
Electrical controls (such as converter controls, switching of shunt capacitor
banks etc.)
Protection relay settings
Most simulation programs capture all of these to some extent. Figure 4-2 illustrates
the components of the WTG.

46

Figure 4-2: Components of a WTG model.

The model structure for the aerodynamics, turbine controls and protection systems for
all of the various WTG will essentially be similar. The parameters will clearly be
different from one manufacture to the other. In the case of stall controlled units there
is no turbine blade pitch control. The shaft dynamics may be modeled as a single
equivalent mass or as is often done these days as two masses, one representing the
rotor/blades and the second representing the electrical generator. In the case of a two
mass model care should be taken to properly represent the damping coefficient
between the two masses for otherwise the oscillation between the two masses will be
grossly over estimated. One example is in the case of doubly-fed induction generators
where a feed forward active damping control loop is used in the converter controls to
dampen out the oscillations between the generator and rotor/blades. In some early
models from one software vendor this effective damping was not properly captured,
which resulted in persistent oscillations between the rotor/blades and the generator,
for system disturbances in simulation work. In general, modeling the shaft as a two
mass system is more important for connections to weak systems where the
perturbations in the generator speed will readily translate into voltage fluctuations on
the system and can thus affect system dynamic performance. Finally note that in the
case of the full converter design since the generator and power system are essentially
decoupled by a full back-to-back converter ideally perturbations in the generator
speed will not be seen on the system since the line side converter can be appropriate
controlled in order to maintain a fixed frequency output regardless of variations in the
speed on the generating unit. Thus, in this case modeling the shaft dynamic from the
perspective of power systems analysis becomes less critical.
Some comments are pertinent with regard to modeling the electrical generator:
Conventional Induction Generator: For this type of machine the generator
may be modeled in the same way as an induction motor. Such models are
readily available in power system simulation programs. Typically, the
equivalent circuit parameters provided by manufacturers for use in modeling
these units are appropriate only for modeling the machine using a single-cage

47

model (i.e. transient fluxes only). This can be pessimistic for simulation of
severe fault conditions since such models have a tendency to over estimate the
current drawn by the machine after fault clearing. A two-cage model
(including subtransient behavior and saturation) is preferred [48]. None-theless, given that most power system simulations are still performed with rather
optimistic load models (i.e. static polynomial ZIP models) the use of a slightly
pessimistic machine model might be a reasonable compromise [48]. Another
important aspect of these designs is blade pitching for active-stall designs
during a severe disturbance. Some manufacturers will pitch the blades during
a severe disturbance in order to remove some of the mechanical power off of
the turbine and thus reduce the level of overspeed to help the recovery of the
unit [20]. This does impact the rate of voltage recovery after a fault is cleared
and can indeed improve voltage recovery [53, 54]. However, to our
knowledge most of the commercial programs presently used in North America
do not capture this control feature.
Again this design may be
Variable Rotor Resistance Generator:
represented using induction machine models. In this case, however, it may be
pertinent to model the variable rotor resistance. In practice the machine tends
to operate up to a slip of roughly 5% at peak load. Then in the event of
sudden wind fluctuations it can increase (decrease) its speed roughly up to
another 5% to absorb (release) some of the energy into the shaft and thus
minimize second to second fluctuations in power providing better power
quality. Also, during a severe voltage dip the rotor resistance is switched to its
maximum value to prevent excessively high rotor currents. Increasing the
effective rotor resistance also has the added benefit of flattening out the motor
torque-speed curve thus making its recovery after fault clearing easier.
Doubly-fed induction generator: With these units one needs to capture the
behavior of the generator and converter since both are connected to the
system. Also, when modeling the shaft as a two mass system adequate
damping should be added to the model to account for the active damping loop
often implemented in the converter controls to improve damping of the
torsional mode between the generator and turbine rotor. One important but
seldom captured (in some commercial programs) behavior of the unit is what
happens during and immediately following a disturbance. For relatively close
in faults, as explained previously the rotor side converter will crowbar
(essentially short circuit the rotor) to protect the power electronics from high
transient currents and thus voltages. Once the rotor crowbar is engaged in this
way the unit essentially becomes an induction generator (or motor if it was
operating at a subsynchronous operating point). Even with active crowbar
systems that disengage the crowbar circuit after the fault has cleared, this is
not done immediately. There may be a short period of time between the fault
clearing and the crowbar circuit disengaging. As such, during this time
depending on the initial operating condition of the DFIG the unit may absorb
significant amounts of both real and reactive power from the system. For
example, if the unit were generating a small amount of power and thus
operating at say 0.15 pu slip then immediately after the fault clears and the
crowbar is still engaged the unit will be essentially an induction motor running
at a relatively high slip and thus absorbing megawatts and megavars. On the
other hand if the unit were at peak load then once the fault clears and the
crowbar circuit has not yet disengaged the unit will be an induction generator
48

running at very high slip possibly 0.2 to 0.3 pu and thus may be absorbing a
significant amount of reactive power, depending on the effective resistance of
the crowbar circuit. This phenomenon may have an effect on the system
voltage recovery in weak systems, but to our knowledge is not presently
captured in the widely used commercial programs in North America. It may
not, however, be appropriate to capture this phenomenon using positivesequence simulation tools. Furthermore, the time period between fault
clearing and the crowbar circuit disengaging is likely between five to a few
tens of milliseconds. Thus this may not be significant. However, this point
needs to be clarified by manufacturers in future model development activities.
Full converter units: For full converter units the most important component
from a system perspective is proper emulation of the dynamic behavior of the
line side converter (or sometimes called inverter). Typically, in the event of a
remote disturbance or fault the line side converter will quickly control the
converter current to its rated (or short term) current rating thus, the unit at its
limit becomes a constant current device. For close in faults where the voltage
is severely depressed the converter may block and stop gating the IGBTs (or
IGCTs). During this time the generator side converter can feed into a resistor
to minimize the overspeeding of the WTG. Once the fault clears the
converters will quickly go back into operation.

4.1.1.2 Wind Farm Modeling for Steady-State (Power Flow) Analysis


Table 4-1 in section 4.1.1.1 provides a summary of how to model the various WTG in
power flow analysis. The question is how to extend this to modeling an entire wind
farm. The wind farms being designed and proposed these days are often of the order
of several tens to hundreds of megawatts in capacity. This means the wind farm has
tens to sometimes nearly a hundred WTGs. Each unit has a dedicated step-up
transformer that raises the units terminal voltage (typically 600 to 690 V) up to the
collector system voltage (typically 25 to 34.5 kV). Then groups of WTG are
connected in a loop or radial line and then connected to a feeder. These feeders are
then all collected together at a single (or sometimes a few) substation. At the
substation the voltage is stepped up from the collector system level to the
transmission system level through a substation transformer. For detailed design and
analysis of the collector system it is necessary to model this entire network in some
form, particularly when studying the potential for harmonic or other resonant
phenomenon. For the study of resonant phenomenon a three-phase harmonic power
flow analysis and/or frequency scanning analysis would be necessary.
For power system studies where we are more interested in the effect that the wind
farm may have on the transmission grid, modeling the details of the entire collector
system may be too exorbitant. Thus, lumped models need to be used for power
system studies. In general the approach should be to group WTGs. For example, if
we are to model a 10 MW wind farm which consists of eleven 900 kW units then it
may be appropriate to lump the whole farm into a single lumped model as shown in
Figure 4-3. If the wind farm were 100 MW with five groups of twenty turbines being
fed to the common substation through five feeders, then we could model the whole
farm with five lumped models, etc. For system impact studies, it is typical practice to
model the whole farm by a single equivalent model such as that shown in Figure 4-3,
since we are only interested in the impact of the facility on the system and not in the

49

collector system design. The model should not be reduced any further since we then
risk not capturing properly the effective reactive and real losses through the substation
transformer etc. Also, as mentioned earlier the shunt compensation in the case of
conventional induction generators should be explicitly modeled rather than over
simplifying the model.

Figure 4-3: Model of conventional induction generator WTG. The GSU impedance is assumed
to be 0.06 pu per turbine on the MVA base of the turbine. The substation transformer
impedance is modeled based on manufacturer data; otherwise a typical impedance of 8 to 10 %
on unit oil-air (OA) rating may be used. The effective collector system impedance may be
calculated by creating a Thevenin equivalent of the collector system; typically this is of the order
of 0.025 pu on the total MVA base of the wind farm. Note: if the collector system is primarily
underground cables, then the effective charging should also be modeled and can be significant.

4.1.1.3 Wind Farm Modeling for Transient Stability Time-Domain Analysis


For power systems transient stability simulations the common practice is to model the
wind farm as a single equivalent machine as shown in Figure 4-3. Again to model
many tens to a hundred individual units can be unmanageable for simulation work,
particularly when the details of the collector system are not known during the initial
study stages. Since on a per unit system models may be easily scaled, one need only
scale the MVA rating of a single WTG model to use it to represent the entire farm.
4.1.1.4 Models Available in The Widely Use Commercially Available Software
Packages
Presently the two major programs used in the North American region for power
system simulations are GE PSLFTM and PTI PSS/ETM. Both programs have had new
releases in the last month, which we have not had an opportunity to review yet. Based
on our last dealings with version 14.2 of GE PSLFTM and PTI PSS/ETM version 29.4
the models available are as follows:
GE PSLFTM

50

o The program has a detailed model of the GE DFIG units and based on
previous work we believe is more representative of the units
performance than the PSS/ETM model. The vendors are working
together to resolve the discrepancies as much as possible.
o The program has generic models for the conventional and variable
rotor resistance type units, but these models have not been developed
in conjunction with those manufacturers and thus are not necessarily
validated.
o Also, to our understanding the DFIG models do not emulate the
behavior of the crowbar circuit as discussed in section 4.1.1.1. Also,
the conventional induction generator models do not have models for
explicitly representing the blade pitching performed by active-stall
units during a disturbance, though one could implement this in epcl
one self as a user-written feature.
o There are no explicit models for full converter units presently.
PTI PSS/ETM
o Most of the major wind turbine manufacturers have worked with PTI
to develop user written models of their turbines. Of these the GE
DFIG models and the Vestas V80 models can be downloaded from the
PTI website. However, the other models are usually supplied by the
turbine vendor to individuals who request them for studies and not
made publicly available. This is for reasons of alleged proprietary
nature of the models as discussed previously this has been a concern
for many power authorities who are required to make models publicly
available to all market participants.
o We have used the DFIG and V80 models extensively and noticed some
issues with them, namely:
Early version of the DFIG model resulted in an awkward fuzz
in the terminal voltage of the machine during a fault. This
made it difficult to properly assess the voltage-ride through
capability of the unit. This has been significantly improved but
could perhaps still be further refined.
The DFIG model does not emulated the behavior of the
crowbar circuit as discussed in section 4.1.1.1.
The V80 model attempts to emulate the switching of the rotor
resistance to maximum rotor resistance during severe low
voltage conditions (and back when voltage recovers).
However, in cases where the unit is tied into a relatively weak
network the model can get into a limit cycling mode where it
continuously bounces back and forth and creates relatively
high-frequency oscillations in the terminal voltage.
In
discussions with the turbine vendor we do not believe that this
is actually the real behavior of the unit, but may be a modeling
artifice. This needs further investigation. The turbine
manufacturer has more detailed EMTDC/PSCADTM models
that they can use to better emulate the behavior of the unit,
however, these models are not publicly available.
Also, consider [47] which indicates that the Irish transmission
operator has seen many of the same (and more) issues with

51

these models. Thus, more work remains to improve the various


models.
4.1.1.5 More Detailed Modeling for Other Types of Analysis
For detailed controls interaction studies and studies related to torsional interaction
more detailed three-phase equipment level models are typically required. Such
models need to be developed in close collaboration with the wind turbine
manufacturer and are typically developed in platforms such as EMTDC/PSCADTM,
EMTP or MATLAB SimulinkTM. These models require a detailed representation of
the converter and its controls for DFIG, variable rotor resistance and full converter
units. Similarly, for studies where the intricacies of the fault ride-through system are
being evaluated detailed three-phase equipment level models are required [55].
4.1.2 Reactive Capability:
BCTCs present standards require that all generators be able to regulate voltage
between 0.9 pf lagging to 0.95 pf leading, at the machine terminals [56]. As weve
seen there are in essence two categories of wind turbine generators (i) conventional
induction generators, and (ii) asynchronous machines that are partially (double-fed
induction generators) or fully (e.g. gear less design with a back-to-back frequency
converter) fed by a back-to-back frequency converter. The former will, in the absence
of shunt compensating devices, consume reactive power at all load levels. The latter
design is capable of generating and absorbing reactive power depending on the design
of the frequency converter. What this translates to is that to maintain the power factor
of the entire farm, at the low voltage side of the substation transformer, between 0.9
pf lagging and 0.95 pf leading additional shunt compensation in the form of switched
capacitor banks or dynamic devices (e.g. static var compensators) or both may be
required. Such additional shunt compensation should be controlled, and coordinated
with other control loops in the wind farm, in order to help regulate the wind farm
substation voltage.
This voltage regulation may be achieved through one of the following means:
Utilization of the automatic control of the reactive capability of the wind
turbine generators themselves. This could be in the form of reactive capability
inherently available in doubly-fed machines and other modern designs such as
permanent magnetic machines with full four-quadrant design converters, or in
the form of conventional induction generators together with power electronic
blocks, such as a shunt voltage source converter, providing dynamic Vars.
A combination of dynamic var compensation in the form of an SVC or
STATCOM, together with coordinated mechanically switched shunt capacitor
banks at the wind farm collector substation.
A third alternative is to have a centralized device such as a large SVC on the
transmission grid, which helps to regulate system voltage in the vicinity of a number
of wind farms. Planning studies would need to be performed to determine the most
cost effective solution. The above requirements would mean that shunt compensation
might be required either on the collector system or at the interconnection point.

52

4.1.3 Line Ampacity


As discussed earlier, the typical capacity factor of a wind farm is between 0.25 to
0.35. Most modern wind turbine designs will cut-in at a wind speed of 4 m/s and
cutout roughly at wind speeds in excess of 20 m/s. What does all this mean from a
planning point of view?
First let us consider the question of line ampacity. The calculation of conductor
steady-state current carrying capacity is largely based on the work of House and
Tuttle [57]. Basically, the current carrying capability of a bare overhead conductor is
based on the maximum allowable conductor temperature and the following heatbalance equation:

qc + qr = I 2 r + qs or I =

qc + qr qs
r

(4-1)

where
I
qc
qr
qs
r

= the continuous steady-state current rating of the conductor


= convective heat loss (primarily a function of ambient temperature,
maximum allowable conductor temperature and wind speed)
= radiated heat loss
= heat received from solar radiation
= conductor resistance

For illustrative purposes, if we assume a 795 ASCR Drake conductor, running East to
West at a latitude of 50o North, on a clear day with the sun mid sky (12:00 pm) then
the conductor current rating as a function of wind speed (assuming the wind is
blowing perpendicular to the conductor) is given in Figure 4-4. In addition, we are
assuming that the maximum allowable conductor temperature is 100 oC with an
ambient atmospheric temperature of 35 oC; i.e. a hot summer day. Figure 4-4 shows
that as wind speeds go from what is typically used in calculating line ampacity (1 to 2
m/s) to values at which a wind farm would be at full capacity (12 m/s or higher), the
line ampacity almost doubles. In fact, most wind turbines will not cut-in until winds
are at least 4 m/s or higher. Here we are not considering other limiting factors that
may limit the capacity of a transmission line, such as:
current rating of riser poles or other connected equipment such as
breakers, disconnect switched, bus segments, wave traps etc.
clearance restrictions on a line that limit the tolerable amount of sag
(this may of course be translated into a lower allowed maximum
conductor temperature).
Thus, it may be prudent to consider the effect this assumption would have on the
thermal rating of the conductors when performing steady-state analysis. The are some
caveats to this:
1. As stated above, the conductor thermal limit is not always
necessarily the limiting factor on the thermal limit of a
transmission line.
2. Most modern wind turbines have towers that exceed 50 m in
height, while a typical transmission tower is between 20 to 30
m in height. Thus, there may be a noticeable difference in

53

wind conditions at the wind turbine tower height as compared


to the height of the transmission conductors.
3. The calculation in Figure 4-4 is based on the assumption that
the wind direction is perpendicular to the line. The convective
losses due to the wind will significantly decrease as the angle
between the wind direct and the direction of the line decreases.
4. Certain lines in the system will extend across a larger
geographic region and thus are exposed to different wind
regimes. Thus, it may not be appropriate to apply such
considerations to these long lines.
Steady-State Current Limit (A)

3000.00
2500.00
2000.00
1500.00
1000.00
500.00
0.00
0.00

5.00

10.00

15.00

20.00

25.00

Wind Speed (m/s)

Figure 4-4: Line ampacity as a function of wind speed; wind direction is assumed to be
perpendicular to the conductor. Assumed 795 ACSR Drake conductors, at sea level, running
East to West, at 50o North latitude, at 12:00pm and with ambient air temperature of 35 oC.
Maximum allowable steady-state conductor temperature assumed to be 100 oC.

4.1.4 Other Planning Issues:


4.1.4.1 Controls Interaction:
With respect to wind turbine generators, one might conceive three main potential
scenarios that may give rise to undesirable interactions and thus may require further
study. These are described below in detail.

Subsynchronous Torsional Interaction:


The first possible concern with regard to torsional modes is that of subsynchronous
torsional interactions (SSTI). This phenomenon was first observed for the Square
Butte HVDC project in 1976 [58]. Historically, subsynchronous torsional interactions
(SSTI) have been recognized as being the phenomenon by which controls associated
with transmission equipment, such as SVC [59] or HVDC [60], may introduce
negative damping torques in the frequency range associated with the torsional
mechanical modes of oscillation of nearby thermal turbine-generating units. ABB has

54

unparalleled experience in this area and has shown through more than three decades
of experience in the design and installation of SVC and HVDC systems, that such
concerns can be addressed through analysis and thus proper design and optimization
of control systems. That is, proper control design and tuning can mitigate the
potential for SSTI.
To illustrate the concept of SSTI, consider the diagram in Figure 4-5. A perturbation
in the speed of the turbine-generator shaft would consequently result in a change in
the generator voltage phase position, and also a possible change in its terminal voltage
and current due to a change in flux linkage. This then changes the ac bus voltage at
the nearby HVDC terminal, assumed here to be rectifying. Consequently, this would
affect the dc voltage, current, and power. The HVDC current regulator then acts to
restore the dc current, leading to a change in the electrical power and thus electrical
torque on the turbine-generator shaft. Assuming a strong electrical coupling between
the HVDC and generator, it may be possible to have a phase relationship between the
initial speed perturbation and the consequential perturbation in generator electrical
torque that destabilizes one or more of the mechanical torsional modes. This of
course is true only for those torsional modes in which the electrical generator
participates. A closed loop therefore exists that involves the speed perturbation,
resulting ac voltage perturbation, and the response of the HVDC converter and its
controls to the ac voltage perturbation.
HVDC
Controls

Vac

Idc

GSU
Ig

Vg

HP-IP
Turbine

Generator

LP
Turbine

Rotating
Exciter

Figure 4-5: Subsynchronous torsional interaction with HVDC and conventional generation.

In the case of wind generation, an interaction between the rotor and blade dynamics of
the wind turbine generator and a closely coupled, near-by active transmission system
such as an HVDC converter is possible [21, 41]. ABB recently performed a study of
this exact scenario with respect to doubly-fed induction generators. The scenario
studied was for a relatively large wind farm (200 MW) being connected to the
transmission grid in the vicinity of a back-to-back HVDC system (also rated 200
MW). For this study no adverse interaction was observed.

55

Quantitative Analysis of SSTI:


As described above, SSTI is an interaction between the generator torsional
mechanical modes (or blade modes) and a nearby HVDC terminal. This interaction is
driven by three factors:
1. The electrical vicinity of the HVDC and relative coupling between the HVDC
and the generators under study.
2. The relative size of the HVDC as compared to the generating unit.
3. The total phase lag from a perturbation in the generator speed to the resultant
perturbation in generator electrical torque, including the action of the HVDC
controls.
The third item above requires detailed analysis. This will be discussed briefly below.
Items 1 and 2, however, can be easily incorporated into a relatively simple screening
tool.
Consider the one extreme where the generator being studied is connected radially to
an HVDC converter station of equal rating. In this case, one would expect the
maximum amount of coupling between the generator and HVDC. The interaction
factor in this case is defined as unity. On the other extreme, if the HVDC were
hundreds of miles away in a highly meshed system, then minimal if any interaction
would be expected. In addition, if the MW rating of the HVDC were negligible as
compared to the generating unit, then one would expect little interaction since the
HVDC would not be able to effect a large enough change in power to result in a
significant perturbation in generator electrical torque. These factors are captured in a
quantity called the unit interaction factor (UIF) first introduced in [61] for
conventional HVDC. The UIF is given by:
MVAHVDC
UIF =
MVAGen

SC Gout
1

SC Gin

(4-2)

where MVAHVDC is the rating of the HVDC system, MVAGen is the rating of the
generator under study, and SCGin and SCGout is the system short circuit strength at the
HVDC commutating bus with and without the generator under study being in service,
respectively. Clearly, for a purely radial case SCGout is zero and thus UIF is equal to 1
times the ratio of HVDC to generator rating, while for a distant generator SCGout is
approximately equal to SCGin and thus UIF approaches zero.
Thus, to screen a case to identify if further analysis is warranted a systematic
approach is followed:
1. Determine all transmission contingencies/outages that tend toward a radial or
nearly radial connection between the generator (a wind farm in this case) and
HVDC.
2. Rank (list) the outages by their contribution to system strength, weakest first.

56

3. Determine UIF for the generator under study for various credible
combinations and permutations of unit and line outages based on the lists
established in (1) and (2) above.
Now for the case of a wind farm, since all the wind turbines are identical units they
may be lumped into one equivalent generator to calculate the UIF. Also, one may
calculate the UIF for the extreme case of when only the first wind turbine comes online. However, clearly for this case (only one wind turbine on-line) SCGin will be
most likely negligibly different from SCGout and thus the UIF will in most cases be
very small.
The accepted practice for conventional HVDC is that more detailed analysis is
warranted if system conditions are found that result in a UIF of 0.1 or greater. [61]. It
should be pointed out that a UIF of 0.1 or greater simply gives an indication of
significant coupling between the HVDC converter and the generator under study.
This alone does not necessarily imply a detrimental interaction.
If more detailed analysis is warranted, then using well-established frequency response
techniques, the net electrical damping over the required frequency range is calculated.
First, a detailed model of the wind farm, electrical system and HVDC controls is
established. The wind farm is not necessarily represented at the level of explicitly
representing each wind turbine, however, it may need to be split into a number of
wind turbine generator each representing a group of wind turbine in the farm that are
on a single feeder. This can be done in a suitable software package, such as
MATLAB SimulinkTM. A suitable electrical network equivalent of the nearby system
should be modeled. The network equivalent should be represented using differential
equations or discrete components (inductors and capacitors) to account for the
variations in network impedance as a function of frequency. In addition, the HVDC
system is modeled with great care and detail, representing as closely as possible the
full details of the actual HVDC control strategy. The rotor mechanical systems of all
generating units are then removed (or disabled) from the model. This means that a
stiff-shaft representation of the generator is used, with infinite inertia. This stiff-shaft
representation of the generator is not merely a simplification; it is an essential aspect
that is needed in order to isolate the characteristics of the electrical system and to
measure them properly. A sinusoidal speed perturbation signal is injected into the
machine model () and the resulting perturbation in electrical torque (Te) is
measured. The transfer function from speed to electrical torque is then calculated
(Te/). The real part of this transfer function, which is in phase with the initial
speed perturbation, represents a damping torque:
Te
De = Re

(4-3)

This real part is then extracted. The calculation is repeated over the requested
frequency interval, and the results are depicted as a family of curves showing
electrical damping versus rotor frequency, with a separate curve for each studied
system condition. A pictorial example of such calculation results is presented in
Figure 4-6.

57

System under N-3 condition


HVDC out-of-service

Electrical Damping (pu)

System under N-3 condition (UIF = 0.8)


HVDC in-service with imbedded
supplemental damping control

System under N-3 condition (UIF = 0.8)


HVDC in-service without supplemental
damping control
0

-1
10

20
30
Rotor Frame Frequency (Hz)

40

Figure 4-6: Electrical damping torque calculations (hypothetical case).

Since the interest here is in a small signal response, the sinusoidal speed perturbation
injected into the model is of relatively small magnitude, typically a few percent. The
model, however, is a nonlinear model representing properly all aspects of system
response and control strategies associated with the transmission equipment. What is
produced by this technique, then, is the small-signal characteristics of the system valid
at the operating point of interest.

Control Instability:
The second possible interaction phenomenon, is simply the potential for interactions
between the wind turbine controls and controls of other nearby transmission or
generation equipment. For example, the DFIG based designs often incorporate a farm
wide central control system to regulate voltage at the substation. An example of such
a system is the GE WindVAr system. This is essentially a centralized controller that
regulates the voltage at the substation connecting the farm to the grid. The high side
(or low side) voltage on the substation transformer together with a current
compensation signal are fed into a proportional-integral (PI) regulator, which
compares the measured voltage (plus current compensation) to a reference signal.
The error is fed in to the PI regulator. The regulators output is fed to the power factor
reference input of all wind turbines in the farm. Thus, effectively this centralized
controller adjusts the power factor of all the wind turbines in the farm on a continuous
basis in order to regulate (within the capability of the farm) the bus voltage at the
interconnection point. If there are any other regulating devices nearby (e.g. power
plant, SVC, HVDC etc.) that are attempting to regulate the same point, steps need to
be taken to ensure the two control systems do not hunt or interact with each other.
Often it suffices simply to provide an appropriate level of droop into the dynamic
VAR regulator through the use of the current compensation setting.

Transient Torque Issues:


The third potential for adverse interactions is system phenomena that may expose the
shaft of a wind turbine to repeated and significant transient torque pulsations. For
58

example, nearby cyclic loads such as arc furnaces, or high-speed reclosing on a


transmission line emanating from the wind farm substation, or repeated commutation
failures on a HVDC link connecting the wind farm to the AC system.
If there are nearby equipment that can expose the wind turbine to such repeated
transient torques, as a first step, some simple transient stability analysis may be
performed to estimate the expected step change in the electrical torque on a wind
turbine generator due to the electrical event, e.g. switching of a line. Then the wind
turbine generator manufacturer must be consulted to identify if the observed level of
transient torque is a concern if the wind turbine were exposed to such a recurring
transient torque (e.g. a few times in a short time period due to high-speed reclosing, or
continuously for a nearby arc furnace etc.). Based on consultation with the wind
turbine manufacturer, more detailed analysis may be required to assess if a potential
problem exists and how it may be remedied.
As a matter of interest, in the past if has been widely accepted that for conventional
thermal turbine-generators, a line switching event that results in a transient torque of
0.5 pu or less (on machine base) does not require more detailed analysis and should
not result in significant loss of life on the turbine-generator shaft [62]. Also, reclosing
of a line following a fault is not a concern provided it is done at least ten seconds after
the fault has cleared in order to allow sufficient time for shaft torsional modes to
decay [63].

4.1.4.2 Harmonics:
In general there are two ways in which harmonics can be generated by wind turbine
generators:
(1) due to saturation in electrical machines
(2) due to harmonic injection by power electronic equipment
The first item is no different than that of any other electrical generator. The
manufacturer of the electrical machine must build their units to comply with the
industry standards (IEEE/ANSI in North America and IEC elsewhere in the world).
The second cause may come from one of two sources. Harmonic injection by softstart thyristor based converters typically used in conventional induction generator
designs for starting the unit (see Figure 2-3). The requirement here is simple, the
wind turbine manufacturer should ensure that their design conforms to IEEE Std. 519.
This should not be too difficult to achieve. The second is by variable speed designs
that use frequency converters, such as the doubly-fed induction generator or the
conventional (synchronous) generator connected through a back-to-back frequency
converter. Once again, the requirement would be to ensure that the manufacturers
design complies with IEEE Std. 519. Typically, with the variable speed designs the
frequency converters are voltage-source converter technologies. This means that the
designs are typically based on pulse width modulation (PWM). These converters will
mainly generator high order harmonics (several kHz).
Finally, the power plant developer must ensure that during the design stage of the
wind farm collector system due consideration is given to ensuring that there is no
adverse harmonic resonance. This is usually a concern when interconnecting to a

59

weak node in the system, thus typically this may be a concern when connecting wind
farms to a distribution network. The concern is that with the high charging
capacitance on underground cables (typically used in wind farm collector systems)
and/or fixed or switched capacitor banks on the collector system, harmonic resonance
may occur and thus give rise to significant voltage distortion. In addition, there may
be a potential for voltage magnification on the shunt capacitor banks near the wind
turbine generators (e.g. at 600 V) when switching higher voltage capacitors on the
collector system or at the substation level. The harmonic resonance issue can be
resolved by judicious design and/or application of filters. If voltage magnification is
deemed possible, solutions might be to minimize the switching surge due to the high
voltage capacitor banks by applying synchronously switched breakers, where
possible. Alternatively, surge arresters may be applied at the lower voltage capacitors
banks to protect them. Thus, during the design of the wind farm electrical system
these and other equipment application issues should be reviewed to ensure proper
design and integrity of the entire wind farm electrical system.

4.1.4.3 Power Quality:


The main power quality question related to wind turbine generators is that of voltage
flicker. In the USA and Canada, the IEEE Standards are adopted. The main
documents that deal with this issue are IEEE Standard 519-1992 and IEEE Standard
141-1995. Both these documents, however, present the flicker limits in the form of a
statistical curve. In addition, the work that led to the IEEE flicker curves is based on
data and research conducted over 50 years ago. Thus, this data seems somewhat out
of date.
The IEC Standard 1000-3-7 [64] is a more comprehensive standard on flicker limits,
though this document is presently not enforced in the USA. Furthermore, IEC
Standard 61400-21 [27] describes methods for actually measuring flicker performance
for wind turbines. As an example, [64] provides the voltage change limits shown in
Table 4-2 and the first step screening limits for power fluctuations in Table 4-3.
Table 4-2: Voltage fluctuation limits (number of fluctuations per hour) [64].

Rate/hour
(r)
r<1
1 < r <10
10 < r < 100
100 < r < 1000

V (%)
Medium
High
Voltage
Voltage
4
3
3
2.5
2
1.5
1.25
1

60

Table 4-3: Stage 1 limits on power fluctuations over a one-minute period (number of fluctuations
per minute) [64].

Rate/minute
(r)

r > 200
10 < r < 200
r < 10

S/Ssc
(%)
0.1
0.2
0.4

Ssc is the short circuit MVA at the bus under study

Based on Table 4-2 and 4-3, a simple screening tool may be developed to identify if
and when more detailed analysis is warranted; that is, when a wind farm is likely to
pose a flicker problem.
Let us illustrate by example. Reference [65] presents the results of a study where data
was collected at a 103.5 MW wind farm in Minnesota for a 12 month period. Based
on this data the maximum change in the farm output (observed during the entire
recording period) for various time intervals were as given here in Table 4-4.
Table 4-4: Maximum power fluctuations over a 12 month period [65].

Time Frame

Maximum Change
(% of farm capacity)
7.3
14.0
63

1 second time frame


1 minute time frame
1 hour time frame

Of course, these changes are not step changes (particularly for the 1 minute and 1
hour time frames) but rather gradual in nature. Now let us assume that we are going
to connect this farm to a 240 kV bus with a minimum short circuit capability of 5000
MVA. If we assume that:
1. the changes are instantaneous (this is a gross, assumption)
2. ignoring the action of any nearby voltage regulation devices
then a relatively simple, however pessimistic, estimate is as follows:

V (%) in one second = (103.5 x 0.073) / 5000 = 0.1 %


V (%) in one minute = (103.5 x 0.14) / 5000 = 0.3 %
V (%) in one hour = (103.5 x 0.63) / 5000 = 1.3 %
Comparing these values to those shown in Table 4-2 and 4-3, we see that the
estimated voltage fluctuations are well below the limits. Thus, we conclude that for
this particular fictitious site there is little potential for voltage flicker problems. Note:
such analysis should be performed for the case with minimum system short circuit
levels, which may be the case with a prior outage of generation and/or a major
transmission element.
Data of the nature presented in Table 4-4 would clearly not be available when a wind
farm is in the planning stage. However, as an initial screening tool these numbers
may be applied to any proposed site to identify if a voltage flicker problem is likely.

61

Then if simple calculations of the nature shown above indicate a potential for large
voltage fluctuations, a more detailed study is warranted. Such detailed analysis
should be done in collaboration with the wind turbine manufacturer, who likely would
have representative data of 1 second, 1 minute and other time frame kW fluctuations
of the output of their turbine for a representative site (It is fully realized that site
specific data would be more applicable, however, when in the planning stages of a
study such site specific data may not be available.). Care should be taken to not over
estimate the potential flicker impact. For example, if a single wind turbine were
quoted to have a 1% fluctuation in its output per second, this does not necessarily
translate to a 1% fluctuation in the output of the entire farm, particularly if the farm is
spread over a wide geographic area. This is because short term fluctuations are of a
random nature and thus the larger the farm the more these short term (second to
second) variation will likely cancel each other out since not all turbine outputs will
be fluctuating together.
As a final statement, it is perhaps fair to say that such flicker concerns are more
dominant at the lower voltage levels where the system short circuit level is weaker.
Thus, it is not likely that voltage flicker would be a significant issue when connecting
farms directly to a strong extra high voltage transmission system.

4.1.4.4 Short Circuit Impact:


The short circuit impact of wind generation depends on the type of generator. For
example, full converter designs (i.e. designs where a synchronous generator
connected to the wind turbine is connected to the grid through a back-to-back
frequency converter, see Figure 2-6) will not contribute to system fault current. On
the other hand, all other designs such as conventional induction generators and
doubly-fed induction generators may be treated in the same way as other rotating
machines. That is, the short circuit current may be calculated based on the units
subtransient and transient impedances. Due consideration must be given to the fact
that conventional induction generators have no excitation system, and thus their
contributions to short-circuit levels may decay rapidly (although larger units will
exhibit in general slower decays) and thus depending on the protection clearing time
their short circuit contribution at the time of fault clearing may be quite small.
Doubly-fed induction generators, depending on design, may aggressively bring back
terminal currents to nominal levels (if not tripped in the process) even before the fault
is cleared. It is interesting to note that per US ANSI Std., induction motors (and
hence one could say induction generators) of 50 hp (~37 kW) or larger should be
considered as a source in short circuit calculations for rating of breakers [66].
In general, for conservative results, the short circuit calculations may be dealt with in
a similar way as with other rotating machinery, taking care not to include wind
turbine generators with full converter designs. Full converter designs essentially
control the current output of the line side converter (inverter) and thus even during a
fault the current output from the inverter is not significantly higher than its rated
current output (though some designs may have higher short-term current ratings to
allow for reactive boosting). To calculate the fault current at the point of
interconnection, both the generator step-up transformer and substation transformer
winding configurations need to be taken into consideration (i.e. Y delta transition).

62

4.1.4.5 Protective Relaying Impact:


The main concern with protective relaying is one of coordination. One example is the
application of switch shunt capacitor banks in the wind farm or at the substation. The
PLC controls associated with the capacitor banks should be properly coordinated with
the existing capacitors on the transmission system as well as the capacitors at the
turbines. Furthermore, if the wind turbine generator rides-through nearby system
disturbances the capacitors at the turbine terminals should not switch out (these are
typically controlled to maintain the generator power factor for conventional induction
generators). If the turbine trips, however, so must the capacitors connected at the
turbine, since otherwise they may cause excessively high temporary overvoltage on
the collector system when system voltage recovers. Likewise, if shunt capacitors are
applied on the collector system or at the substation, these too may have to be designed
to come off-line in a timely manner when the wind turbine generators are
disconnected from the system by a prolonged disturbance. Appropriate voltage
deadband and delays should be chosen for switching each bank. These settings are a
matter of proper coordination of controls and protection.
Where under voltage relays and/or low-voltage ride-through schemes are applied on
the wind turbines, these settings should be properly coordinated with the transmission
system protective relaying. That is, for example, normal clearing in zone 1 on a
transmission line protective relay should have the opportunity to clear a line well in
advance of a turbine tripping. In some cases such coordination may mean upgrading
local transmission protective devices. For example, it is not uncommon to find on
some subtransmission systems in areas where there is not a significant amount of
generation that Zone 1 relaying may be based on distance relays while Zone 2 is
based on overcurrent. Thus, for a close in internal fault on a line the nearby end may
trip in several cycles while the remote end, clearing on Zone 2, may take many tens of
cycles to clear. Thus, the fault essentially remains on the system for perhaps half a
second or more. Under such circumstances if a wind farm were to be installed on this
system it may experience low-voltage conditions for a prolonged duration due to the
slow clearing of the Zone 2 relaying. As such, rather than requiring what may end up
being an exorbitant requisite on the low-voltage ride-through capability for the wind
farm, it would be far more prudent and cost effective to upgrade the line protection to
allow transfer-tripping on the remote end of the line and thus clearing the fault from
the system in a far shorter time period.

4.1.4.6 Self-Excitation:
Self-excitation, as it applied to induction motors or generators, is the condition by
which an electrical resonance occurs between the inductance of the induction machine
and a series capacitance [67].
Such self-excitation may be caused by two sources:
1. The application of series capacitors on a radial transmission line(s) feeding a
wind farm with induction generators (note: even doubly-fed induction
generators are susceptible).
2. Islanding that leads to a substantial amount of induction generator wind
generation remaining in the island feeding primarily capacitive load (such as
shunt capacitors used for power factor correction and voltage support).

63

In the case of series compensation, the potential for self-excitation exists primarily
during start up when the wind turbines are being started as induction motors. Thus,
they may hit a resonant frequency as the turbine tries to come up to speed. For
doubly-fed induction generators, since they are variable speed machines, a potential
for self-excitation may exist even under normal operation when the generator is
operating at a significant slip frequency. Such analysis needs to be performed using
system and machine models that capture the frequency dependence of network
parameters, that is programs that represent network elements with differential
equations rather than a constant impedance matrix. If a wind farm is to be fed by a
series compensated line, then analysis should be performed to identify the potential
for self-excitation or other possible resonances (e.g. SSR).
For distribution connected wind farms a transmission fault may result in a wind farm
operating in an isolated region of the distribution system (e.g. outage of a distribution
substation transformer). Such conditions should generally be avoided, by transfer
tripping of the wind turbines. This is because (a) the islanded condition may result in
self-excitation and (b) most asynchronous wind generators cannot maintain stable
operation in a system without a synchronous source (There are of course exceptions to
this, for example, properly designed vector controlled doubly-fed machines or WTGs
connected to the load through a full voltage-source converter.). This is perhaps one of
the other reasons why in the past where most, if not all, wind farms were connected to
the distribution system the turbines were allowed to trip quickly following a network
disturbance. This should be kept in mind simply to realize that though low-voltage
ride-through capability is a desirable feature for large wind farms connected to the
transmission grid, it may not necessarily be desirable for small (say 5 MW or so) wind
farms buried in the distribution system.

4.1.4.7 Transmission Planning for Intermittent Resources


One of the main challenges with transmission planning as it relates to wind power
generation is how to account for the fact that a wind farm spends perhaps only a few
hours of the year at its peak capacity. Furthermore, when studying a system with
numerous wind farms, particularly land based and spread out over a relatively large
area, it is quite unlikely that all wind farms will be at the peak capacity at any given
point in time. None-the-less, most system impact studies in the USA have performed
power flow analysis using a peak load cases married with peak capacity on the
generation under study. Though, this assumption may be quite valid for conventional
plants (e.g. fossil fuel, nuclear and even hydro in some cases) it is quite unlikely when
considering wind generation. There is, however, some merit to this sort of analysis
since it helps to identify fatal-flaws in the transmission system. For example, if there
is adequate transmission to be able to simply inject the peak capacity of the farm into
the system.
Presently, the standard process in North America for performing studies related to the
interconnection of power plants is a four step process:
1. A fatal-flaw analysis to quickly identify if any major limitation existing at the
point of interconnection. If this screening analysis is favorable then one
moves to the next step.
2. At this stage data is submitted by the plant developer to the utility/ISO and an
interconnection system impact study is initiated. This study takes a more

64

detailed look at steady-state (including short-circuit analysis) related issues


and also looks at dynamic performance of the system with the plant addition.
3. At the next level a facility study is performed which is yet another more
detailed study in order to identify the exact system upgrades required and what
the interconnection costs will be to facilitate these required upgrades. At this
point an interconnection agreement is signed.
4. Finally a transmission service study is performed to identify if the generation
facility can transfer its power to the load centers it wishes to sell into.
Sometimes, step 4 in the above list can be performed in parallel with steps 1 through
3.
One of the concerns with the above approach is that for wind farms representing the
farm at peak load throughout the studies is not necessarily the best approach. When
performing power flow analysis related to wind farm interconnection and planning
there should be two objectives:
1. Fatal Flaw Analysis: Can the megawatts from the wind farm be injected into
the system and reliably transmitted to the nearest major transmission
substation?
2. Planning Analysis: Does the introduction of the wind farm(s) create or
further exasperate transfer capabilities and major transmission paths on the
bulk transmission grid?
To answer the first question one can perform analysis similar to what is typically done
at present. Namely, assume peak output from the farm and identify with some limited
N-1 contingency analysis if the peak megawatt output of the wind farm can be
reliably carried to the transmission grid. The simplest case is when the wind farm is
connected to a major transmission substation through a radial line. Obviously in this
case the radial line should have a continuous thermal rating equal to or greater than
the peak capacity of the wind farm. In addition, the remaining lines emanating from
the major substation should ideally not be thermally overloaded under an N-1
contingency when the wind farm is at peak load.
To answer the second question, however, some more detailed analysis is required.
One approach is as follows:
1. Start by defining the size (number and type of wind turbines) and geographical
location of the wind farm(s) to be studied.
2. Then by using a detail meteorological model as well as wind turbine data (i.e.
hub height, power curve, etc.) perform simulations to determine the megawatts
generated by each wind farm for 8760 hours in the year. This can be done
both for a historic year, where recorded wind data is available, and then using
a meteorological model forecasted for a future year. This type of work can be
done by a number of wind engineering experts such as AWS True Wind or
3TIER Environmental Forecast Group Inc., etc.
3. Now by using a market simulation tool such as ABBs Gridview or GEs
MAPS, one can perform a security constraint, economic dispatch for each
hour of the year assuming that all available wind power is purchased and
injected into the system. This analysis can be performed for both the historic
year as a benchmark and for future years. As an input to the simulation tool
one would need to enter the daily load curve throughout the year as well as
fuel costs, outage rates etc. for all other generating units.

65

Based on the results of item 3 above one can identify what transmission paths are
limiting and for how many hours of the year. This analysis can then help to identify
what transmission bottleneck may exist that require upgrading or otherwise enhancing
to facilitate the full utilization of the wind resources.

4.1.4.8 Capacity Credit


Another question that is often asked is whether or not a capacity credit can be given to
intermittent resources such as wind and if so how much? Some methods use rather
simplified approaches such as using the capacity factor of the wind farm. These
approaches can be quite rough and present risk to both the wind farm operator and the
utility. Generally, the best way of calculating capacity credit is to use similar tools as
discussed above. However, for capacity credit calculations it is essential that the
simulation be performed for several years to fully capture a good sample for
probability calculations (since wind variability may change from year to year [68]).
The calculation technique would be similar to standard Loss of Load Expectation
(LOLE) analysis.
First we would develop a model of the existing generating facilities taking into
account outage rates etc. Then we would take simulated megawatt output from the
wind farm on an hour by hour basis for the time period under study (developed based
on a meteorological forecasting model or available recorded data). Also, hourly load
curves would be needed. Then two sets of Monte Carlo simulations would be
performed. The first set, as a benchmark, would calculate the hours when capacity
does not meet demand (load) using the existing generation on the system. This
calculation would be done for a range of installed base; that is, for example first with
all generators assumed to be in-service, then with say 100 MW assumed to be
unavailable during the entire period etc. In this way we can calculate LOLE and plot
it versus installed MW capacity as shown in Figure 4-6. Now we can go back and run
the same cases with the wind generating facility in-service (and generating megawatts
based on the hourly values calculated from our meteorological model). The
difference between the two curves, at the desired LOLE level, is the capacity credit
that can be offered to the wind farm (see Figure 4-6). This is also referred to as the
Effective Load Carrying Capability (ELCC).
Note that in general todays wind turbine technologies are quite reliable and so in the
case of wind farms it is the unavailability of the energy source (adequate wind) that
mainly determines a forced outage in the above analysis. A typical modern WTG
may have a forced outage rate of say only 2 to 5%, due to mechanical failures. In
contrast, for a conventional fossil fuel plant the fuel source is extremely reliable and it
is mechanical failures that result in forced outages.
Finally, we should note that other simpler techniques are being applied in the industry
to calculate the effective load carrying capability of wind generation. A summary of
some of these techniques may be found in [68]. The methodology presented here is
perhaps the classical approach that is data intensive and uses a probabilistic technique.

66

0.35

LOLE (days/yr)

0.3
0.25
0.2

Without Wind
With Wind

0.15
0.1

Capacity is 160 MW

0.05
0
7800

8000

8200

8400

8600

8800

9000

Total Installed Conventional Generation


Figure 4-6: Capacity credit calculation for a wind farm.

67

5 Interconnection Standards and Criteria


Based on the discussions in the previous sections, the following recommendations are
made as a starting point for interconnection requirements specific to wind generation.
5.1

Summary and Comments on Wind Standards World-Wide

Let us now take a critical look at some of the more highly debated aspects of all the
various standards and practices discussed in section 3:
PSS: Power system stabilizers (PSS), in the traditional sense, do not apply to
wind turbine generators since they are essentially asynchronous units. Even in
the case of so-called synchronous WTG, these are typically connected to the
grid through a full four-quadrant back-to-back converter. As such, these units
are also essentially decoupled from the system and do not really participate in
electromechanical modes of rotor oscillation. The fundamental concept
behind PSS tuning is to introduce a component of damping torque, of
electromagnetic origin, on the shaft of a generator in the range of frequencies
associated with electromechanical modes of rotor oscillation [69]. The
concept was first introduced in [70]. Based on the generalized theory of
induced damping and synchronizing torques [71, 72], one can show that even
in the case of stabilizers applied to FACTS devices the mechanism is one of
induced damping torque on the shaft of nearby synchronous generators, for
these modes are essentially modal oscillations between a single or small group
of units and the rest of the system (local modes) or among large groups of
units system wide (inter-area modes) [73]. Thus, asynchronous units do not
participate in these modes in the conventional sense. In the case of a
conventional induction generator there is no excitation system with which to
effect control. None-the-less, based on the concepts in [71, 72] it may be
possible to design stabilizers for WTGs with converters or on an SVC at a
wind farm substation to improve the damping of inter-area modes. This,
however, would be contingent on the location of the wind farm and thus the
controllability of the mode from that location. Also, even if this was achieved,
to be fully effective all WTG need to be on-line to reap the full benefit of all
stabilizers. Thus, in the case of a wind farm a more effective means of
stabilizing electromechanical modes may be to place a stabilizer at a
centralized location such as on an SVC dedicated to the farm. In short,
although based on decades of research and field experience it is clear that all
conventional synchronous generators do participate in electromechanical
modes of rotor oscillation (whether local, inter-area or both) and can benefit
from the application of PSS this same general conclusions cannot be easily
extended to WTGs there has been little research in this area. However, a
wind farm can certainly have an effect on the damping of inter-area modes of
rotor oscillation simply through the effect that it has on inter-tie power flows
but one cannot necessarily generalize to say whether such an impact would be
positive or negative. Furthermore, should it degrade small-signal stability the
best course of action may be PSS applications on nearby conventional
synchronous generators that do not have a PSS or supplemental damping

68

controls on other devices such as SVCs etc. Thus, at this time we recommend
that PSSs should not be required on WTGs.
AVR and Frequency Governing: The concept of automatic voltage
regulation can be easily extended to wind farms. For technologies such as
DFIG or the full converter units both the unit itself and on a farm level,
automatic regulators may be applied to regulate the reactive output of all the
units to maintain a certain voltage at say the point of interconnection between
the farm and grid. In the case of conventional or variable rotor resistance
units, again voltage regulation can be effected, however, in this case it would
likely come from a supervisory control system regulating the reactive
resources throughout the farm (e.g. an SVC or STATCOM at the substation
coordinated together with a number of mechanically switched shunt devices).
As for frequency control, this is discussed in more detail in the companion
report. In general the concept may be extended to WTGs and already (as seen
in the discussions in section 3) some power authorities are requesting this
feature. However, we feel that asking wind farms to actively participate in
primary frequency control during under-frequency events may not be as
technically feasible or reliable as employing conventional generation
particularly in a system rich with hydro generation.
Low Voltage Rid-Through: There is a range of requirements. Some, most
notably Hydro Qubec and Denmark, require that the wind farm ride-through
for voltages down to zero for the duration of a normally cleared (Zone 1) fault.
While others German, Spain, Ireland and Alberta have settled for voltages
down to 15 to 20% at the point of interconnection. Since wind farms are being
proposed in all these localities certainly both standards can be met, at least by
some manufactures. However, based on discussions at FERC [30] and in the
Irish experience, there is still a tendency to indicate that the technology is
rapidly improving but presently the cost of achieving ride-through down to
zero voltage is significant. As such our recommendation is to adopt the Irish
approach, namely to accepted the E.ON standard at present which can be met
by all the reputable manufactures and to then revise the standard to a more
rigorous level as the technology improves. Clearly such an approach will
mean that once a more rigorous standard has been put into place it cannot be
applied retroactively to older installations since it may not be physically
possible to upgrade some of the older technologies to meet newer standards,
without possibly replacing the WTGs.
Reactive Power and Power Factor: Some standards are beginning to request
that reactive power compensation be provided whether or not the wind
generation is on-line (or at no-load). Although in principle this may sound
good, in practice one needs to be aware of the potential implications. To
optimize the collector design, much of the reactive compensation may be
placed at various locations on the collector system. In addition, in the case of
DFIG and full converter units, the reactive capability of the unit is primarily at
the turbine. Although it is quite possible to (through proper design and
functionality) make this reactive power available even when the turbines are
not running (specially so for the full converter units since their full reactive
capability is in the line side converter), this could be problematic if such
reactive power is for the purpose of controlling the bulk transmission system
voltage. Under extreme light loading conditions in the wind farm, if the
reactive power injection at the turbines is increased for the purpose of
69

supplying reactive power to the grid, stead-state voltages on the collector


system (and at the turbines) may rise to excessive levels thereby
compromising insulation integrity within the farm and causing potential safety
hazards. In the end we recommend that the voltage regulation be provided to
adequately control voltage at the point of interconnection throughout the range
of power delivery of the farm. This functionality should be properly
coordinated with other voltage regulating devices on the transmission system.
In cases where a good portion of reactive compensation is provided at the
wind farm substation in the form of an SVC or STATCOM, such coordination
is probably more easily achieved since it is done through a centralized control.
Moreover, the device may be more easily used for voltage regulation even
when the wind farm is disconnected or supplying no power since it is not on
the collector system but directly connected to the substation. Finally, based on
general consensus we believe a power factor requirement of +/- 0.95 at the
point of interconnection is a reasonable requirement.
5.2

Wind Generation Facilities with Aggregated Capacity of Less than


10 MVA

The benefit of the recommendations in this section are quite clear when applied to
large wind farms (e.g. 50 to 100 MW and larger) connected to the transmission
system. However, for quite small wind farms (e.g. 5 to 10 MW or less) that are
buried in the distribution system and have no net megawatt injection into the
transmission network, the cost of implementing some of the recommended controls
described below may out weigh the benefits, since such units may not significantly
contribute to the overall system performance. Thus, BCTC should retain discretional
rights to override the requirements generally discussed below, where such
requirements may not render any significant benefit. As a caution, consider an
example. If say twenty 10 MW wind farms were to be connected to the distribution
system in a region of the system that is close to a transmission node, and thus a
significant portion of this power was on a regular basis being transferred onto the bulk
transmission grid, then although we have twenty individually small farms, in essence
it is as if we have a 200 MW wind farm connected to the transmission system. Thus,
in a case such as this it would be prudent to require the types of controls and strategies
being recommended below for each farm since not doing so could expose the system
to a sudden loss of 200 MW of generation. It is therefore evident that the distinction
between distribution connected and transmission connected wind generation may be
somewhat subjective. An easy way to make this distinction is to discuss the net
amount of megawatt transfer. That is, if the megawatt output from a small
distribution-connected wind farm is being consumed locally at the distribution level
such that a system disturbance that disconnects the wind farm from the system also
disconnects the load (i.e. no net imbalance in system load generation) then, for
example, the requirements for low-voltage ride-through and power
limitation/curtailment may be moot for such a farm. However, if megawatts are being
transported over the transmission system to remote loads then the loss of the
generation following a disturbance would constitute load/generation imbalance and
thus would not be acceptable in large amounts. In general we recommend that BCTC
adopt the IEEE Std. 1547TM Standard for Interconnection Distributed Resources with
Electric Power Systems, and its applicable guides, for farms with an aggregated
combined power output of 10 MVA or less this standard was recently developed for

70

distributed generation facilities of 10 MVA or less. The IEEE is presently working on


extending this activity through the development of Std. 1547.5 for facilities with an
aggregate capacity of more than 10 MVA this activity has just recently started.
5.3

Proposed Wind Interconnection Standard Outline For Wind Farms


with an Aggregated Capacity Greater than 10 MVA

Production Limits and Control:


The maximum output of each wind farm must be specified, and the wind farm
15 minute average megawatt output must not at any time exceed this value.
The transmission system operator, through telemetry (SCADA), should have a
means at all times to monitor the megawatt output of a wind farm and whether
or not turbines are on or off-line due to high winds, lack of wind or cut-out
due to a disturbance or other reason.
The transmission system operator, through telemetry (SCADA) or other means
of contact with the wind farm operator, must have a means of issuing a
directive to reduce the megawatt output of the wind farm, or curtail the wind
farm in the event of an emergency condition where the wind farm output is to
be reduced to respect thermal limits on nearby transmission corridors or when
otherwise generation far exceeds demand. This can be achieved through a
centralized controller at the wind farm that feeds a signal to the reference input
of each turbine in the farm. The reference input of such a controller may be
biased in order to effect a reduction in the power output of the wind farm.
Such a reference bias may be issued either automatically through telemetry by
the system operator or may be communicated by the system operator to the
plant operator who in turn imposes the required bias on the control reference.
Alternatively, the wind farm operator may intentionally disconnect blocks of
wind turbines in the farm to effect a reduction in generation. In either case,
this should be done in a controlled fashion and at the specified rate by the
operator (e.g. certain number of MW/minute).
All turbines in a single wind farm must not start or stop simultaneously when
cutting-in or out due to wind fluctuations. This can be done through
staggering the start-up and shut-down controls of the wind turbines in a farm
such that they do not start simultaneously.
Low-voltage ride-through (LVRT):

All wind turbines should be able to ride through a normally cleared single or
multi-phase fault at the high-side (transmission voltage level) of the substation
transformer. Clearly, this requirement does not apply to radially fed wind
farms, where an internal fault on the line would in effect disconnect the wind
farm from the system. In such cases, the turbine in the farm should be able to
ride through any normally cleared single or multi-phase faults external to the
radial line feeding the farm. In addition, WTGs in the farm are not expected
to ride-through faults that are internal to the wind farm (i.e. on the wind farm
collector system). Note: there are possible caveats to the above condition on a
radially fed wind farm. Consider a case where a wind farm is fed by a radial
line. Now if a second wind farm were to be connected somewhere along this
line in the future through a ring-bus arrangement, it would now be possible to
71

isolate the segment of line between the ring-bus and each of the two wind
farms separately. Therefore, ideally it would be desirable that for a fault that
trips the line segment between this new substation and one of the wind farms,
the second wind farm does not also trip.
Based on presently applied deterministic planning standards generally adopted
throughout the North American continent, the above requirement would
translate to the wind farming being required to ride-through a 3-phase fault
close to the point of interconnection (but external to the wind farm). This
would mean being able to tolerate at the point of interconnection a duration of
between 3 to 6 cycles of zero voltage. Depending on the wind turbine
technology this does not necessarily translate to zero voltage at the turbine
terminals. We believe that this can be achieved (certainly for conventional
induction generators). In fact based on requirement by Hydro Qubec [20]
this is being met by manufacturers. None-the-less, discussions at FERC [30]
suggested that this may still be a relatively new development and still requires
further development. Thus, based on the most common accepted strategy, we
recommend that the E.ON fault ride-through curve (Figure 3-3) be adopted at
the point of interconnection (the high voltage side of the wind farm substation
transformer) this is the same as that proposed by AWEA. This does mean
that a close in bolted 3-phase fault (the most common cause of which is
closing breakers into a workmans ground being inadvertently left connected
to a bus segment after maintenance work) would result in tripping the farm.
However, it is unlikely that this would result in multiple wind farms/generator
tripping. So if the wind farm in question has a total installed capacity of less
than 500 MW (the largest unit on the BC system), this is likely a tolerable
event. What cannot be tolerated is to have a single disturbance result in the
loss of multiple generating plants. In the future as the technology improves
(and wind penetration increases) then a fault ride-through strategy similar to
Hydro Qubec (Figure 3-2) or E.ONs standard for high short-circuit units
(Figure 3-4) may be adopted. Note: some jurisdictions have asked for an
ability to tolerate zero voltage for up to 600 ms this is unreasonable since not
even conventional fossil fuel plants would typically be able to tolerate this.
What we are recommending is to slowly transition to requiring that the wind
farm be able to ride-through any (even a bolted 3-phase fault) fault external to
the farm for the time it takes to clear the fault in Zone 1 (typically 3 to 6 cycles
at transmission voltages).
For more prolonged disturbances, where the system voltage does not recover
for an extended period of time (significantly beyond normal Zone 1 and Zone
2 clearing times), the wind turbines in the farm must be tripped by under
voltage protection. In the case of conventional induction generators, the shunt
capacitors at the turbine used for power factor correction must also be
disconnected from the system if and when the turbine is disconnected to
prevent overvoltage conditions when the system voltage recovers.

Voltage/Frequency Operating Limits:


For transient overvoltage conditions and for off-nominal frequency operation, WTG
should be expected to meet BCTCs existing criteria under section 8.4.6 [56]. Note:
for low-voltage ride-through the specifications in the above subsection should
override those in Table 5, section 9.4 of [56]. Clearly, all wind farms must operate
72

continuously at and between +/- 10% of nominal system voltage and +/- 0.5 HZ of
nominal system frequency.
Power Quality:
WTG should be expected to meet BCTCs existing requirements under section 8.2 of
[56]. This is inline with IEEE Std 519.
As an added note, the wind farm owners should be responsible for performing studies
where necessary to ensure that they avoid self-excitation and other low-order
harmonic resonance phenomenon on their collector system due to shunt capacitors (or
cable charging).
Control Interactions:
Where necessary, the wind farm owner is responsible for performing studies to ensure
that automatic controls on the WTG (or employed farm wide) do not adversely
interact with other automatically control transmission devices such as nearby SVC,
HVDC etc.
Reactive Power Requirements:
Based on standard industry practice [31] and recent recommendations by AWEA [29],
and FERC [74] it is recommended that wind farm facilities be able to operate between
+/- 0.95 power factor when at its peak megawatt capacity, as measured at the point of
interconnection. This point is defined as the high voltage side of the substation
transformer. This reactive capability should also be in the appropriate ratio of discrete
(mechanically switched shunt capacitors) and dynamic (rotating machine VAr
capability, SVC, STATCOM etc.) to ensure both steady-state and transient stability
during and after a disturbance. Studies may of course determine that the actual
reactive needs are slightly higher or lower depending on the size and location of the
wind farm.
Protection:

The wind farm developer is given the responsibility to put into place adequate
protection to safeguard all wind turbines and other equipment within the wind
farm. Such protection, however, must be properly coordinated with controls
and the protection on the BCTC transmission system.
To design and implement such protection, in special cases, the plant developer
may be required to perform specialized analysis such as:
o transient torque analysis, if and when high-speed reclosing is
exercised on lines emanating from the wind farm
o controls interaction studies, if the wind farm is being placed in-service
in close vicinity to an existing power plant or HVDC system, etc.

Modeling Requirements and Verification Tests:

The wind farm developer is given the responsibility to ensure that adequate
models of the wind farm are supplied to BCTC for system studies.
73

Such models must be updated upon commissioning of the wind farm, if any
model structures or parameters are substantially changed between the planning
and commissioning phase of the farm.
Verification tests or studies must be performed to demonstrate that the wind
farm meets all of the requirements set forth by BCTC. In particular, the lowvoltage ride-through capability of the wind turbine must be demonstrated
either through detailed (3-phase type) modeling of the wind turbine controls,
or by field or factory tests demonstrating the capability of the wind turbines to
ride through the prescribed fault conditions. Also, of importance is
verification (by tests or detailed modeling) of the performance and response of
the reactive power controller (such as dynamic VAR control systems on
doubly-fed machines) to ensure proper coordination between controls and
protection.

Power System Stabilizers and Automatic Voltage Regulators:


WTG should not be required to have power system stabilizers fitted to them. The
concept of automatic voltage regulators (AVR) is not necessarily directly applicable
to WTG technologies. However, it may be stipulated that the voltage at the point of
interconnection between the wind farm and the transmission gird should be
automatically regulated. How exactly this is achieved may be left at the discretion of
the wind farm developer.
Disturbance Monitoring:

For the purpose of continued understanding of wind farm operation and


control, BCTC may consider requiring that each wind farm be fitted with a
disturbance monitor.
The monitor should at least monitor:
o For the entire farm (at the interconnection point)
Voltage and current
Active power transfer
Reactive power transfer
System frequency
o For a single turbine (of each kind if the farm has turbines from
multiple manufacturers)
Turbine mechanical speed
Generator active power output
Generator reactive power output (including shunt capacitor
banks at the turbine)
Generator terminal voltage

SCADA and Telemetry:


BCTC should consider requiring that the following information be made available by
SCADA, for each wind farm:
Average wind speed and direction at the site, as required to facilitate use in
forecasting models.
Megawatt and megavar output at the wind farm.
Number of turbines off-line due to high wind conditions.
74

Number of turbines off-line due to electrical or mechanical problems.


Status of circuit breakers on substation transformer and all shunt compensation
devices located at the wind farm substation.
The tap position of the wind farm substation transformer, if fitted with on-load
tap-changers.
Telemetry between wind farm (or wind farm operator) and the transmission
system operator to allow for curtailment of megawatts (i.e. limitation of plant
maximum output).
Telemetry between wind farm (or wind farm operator) and the transmission
system operator to facilitate specifying voltage regulation set-points for the
point of interconnection between the wind farm and the transmission grid.

Some of the above data may be used for forecasting purposes and so the data format
and quality should be in a format that is suitable for such use.

75

6 Conclusions & Recommendations


This document presents a summary of the experience with wind generation within the
North American continent and elsewhere in the world. In addition, a survey is given
on interconnection standards adopted through the world for integrating wind
generation into the utility grid. Based on this background information and ABBs
experience with wind turbine technologies and wind farm studies, recommendations
are given on the best practices and emerging standards for interconnecting large wind
farms to utility grid.
The intent of this report is to objectively discuss the issues and concerns related to the
integration of wind generation into the BCTC System. The outcome has been a series
of recommendations, all technically feasible and viable, for possible incorporation
into a new standard for interconnecting wind farms to the BCTC system. The goal of
any such standard will no doubt be to establish minimum technical requirements for
connecting wind farms to the transmission system. Such a standard should make no
distinction as to the preferred or not preferred technologies or types of wind
generation equipment, but rather rely on the ingenuity of the wind farm developers,
wind turbine generator manufacturers and other power equipment manufacturers to
come up with the most cost effective means of meeting and/or exceeding the technical
standards established. In the end, most of the issues surrounding integration of wind
generation into the bulk transmission system are commercial issues and limitations
and not technical limitations. The technical problems can be addressed regardless of
the type or vendor of the wind generation system, provided proper analysis and design
is performed upfront. As wind generation technology advances the use of WTGs is
becoming more and more economically viable and quite competitive compared to
other more traditional energy sources this in the end will help to ensure greater
utilization of our natural renewable resources helping to safe guard our global
environment.

76

References
[1] American Wind Energy Association, The Most Frequently Asked Questions About Wind
Energy, 2002. (Document produced in cooperation with US Department of Energy and the
National Renewable Energy Laboratory). www.awea.org
[2] Part 1 - Early History Through 1875, http://telosnet.com/wind/early.html
[3] S. Heier, Grid Integration of Wind Energy Conversion Systems, John Wiely & Sons, 1998.
[4] www.bwea.org
[5] J. Pedersen, P. B. Eriksen and P. Mortensen, Present and Future Integration of Large-Scale Wind
Power Into Eltras Power System, Eltra, www.eltra.com
[6] Harley Lee, Wind Energy: A Success Story, http://www.endlessenergy.com/windenergy.shtml
[7] News briefs, US renewables increase after five years of decline, Refocus, Nov/Dec 2003.
[8] The European Wind Energy Association, Wind is Power, www.ewa.org
[9] T. Bulow, Eltra Prepares Storage of Wind Power, www.eltra.dk
[10] Eltra on Lookout for New Electricity Storage Facility, January 19, 2004 www.eltra.dk
[11] European Commission, Wind Energy The Facts, Volumes 1 5, www.ewea.org
[12] Australian Wind Energy Forges Ahead, January 12, 2005, www.auswea.com.au
[13] D. J. Appleyard, Lord of the Winds, Refocus, Sept/Oct 2004.
[14] Global Wind Energy Market Report, for 2003, www.awea.com
[15] S. B. Green and B R. Lemon, How Patent Law is Affecting the Wind Power Market, North
American Windpower, Volume 1, Number 11, December 2004.
[16] L. L. Freris, Wind Energy Conversion Systems, 1990.
[17] H. Slootweg and E. de Varies, Inside wind turbines Fixed vs. variable speed, Renewable
Energy World, January/February 2003.
[18] D. S. Zinger and E. Muljadi, Annualized Wind Energy Improvement Using Variable Speeds,
IEEE Transactions on Industrial Applications, November/December 1997.
[19] P. Gipe, Pitch Versus Stall: The Numbers are In, Wind-Works.org, June 20, 2003 www.windworks.org
[20] NEG Micon (brochure), NEG Micon Power Quality, 2003
[21] P. Pourbeik, R. J. Koessler, D. Dickmander and W. Wong, Integration of Large Wind Farms into
Utility Grids (Part 2 - Performance Issues), Proceedings of IEEE PES General Meeting, July 2003.
[22] GE Wind Energy (brochure), Low-voltage ride-through technology, 2003.
[23] J. Niiranen, Application of Power Electronics, presentation on Distributed Generation at
University of Vaasa, May 23, 2003.
[24] Wind Chimes In, EDN December 7, 2004, pg 38.
[25] IEC 61400-1, Wind turbine generator systems Part 1: Safety requirements, 1999.
[26] IEC 61400-12, Wind turbine generator systems Part 12: Wind turbine power performance
testing, 1999.
[27] IEC 61400-21, Wind turbine generator systems Part 21: Measurement and assessment of power
quality characteristics of grid connected wind turbines, 1999.
[28] Alberta Electric System Operator, Wind Power Facility Technical Requirements, 15 November
2004. www.aeso.ca
[29] American Wind Energy Association, Standardizing Generator Interconnection Agreements and
Procedures Doket No. RM02-1-001, submitted to FERC www.ferc.gov
[30] Transcript of the Technical conference on Interconnection for Wind Energy and Other Alternative
Technologies, Docket Nos. PL04-15-000 et al., September 24, 2004 www.ferc.gov
[31] NERC/WECC Planning Standards, August 2002.
[32] Hydro Quebec, Technical Requirements for the Connection of Generation Facilities to the HydroQuebec Transmission System: Supplementary Requirements For Wind Generation, May 2003.
[33] MISO, Generator Interconnection Procedures and Agreement, December, 2002.
www.midwestiso.org
[34] Xcel Energy, Interconnection Guidelines for Transmission Interconnected Producer-Owned
Generation, Version 1.0L, 1/27/03.
[35] G. Wolf, PNMs Taiban Mesa switching station connects wind farm to power grid in record
time, Transmission & Distribution World, November 2003.
[36] The Effects of Integrating Wind Power on Transmission System Planning, Reliability and
Operations: Phase 1 Preliminary Overall Reliability Assessment, prepared by GE Power Systems
Energy Consulting, for NYSERDA, February, 2004 www.nyserda.org

77

[37] K. Porter, C. Smith and S. Wiese, Wind Energy Interconnection, September 2003, National
Wind Coordinating Committee.
[38] White paper, The need for voltage ride-through performance standards for wind turbines,
February, 2003.
[39] WECC, Generator Electrical Grid Fault Ride Through Capability, October 2004.
[40] M. Mc Govern, Steep Learning Curve for Grid Operator, Windpower Monthly, December
2003.
[41] Eltra, Specifications for Connecting Wind Farms to the Transmission Network, (unofficial
translation of Danish Eltra doc. No. 74174), Second edition, April 26, 2000.
[42] J. K. Jensen, Towards a Wind Energy Power Plant, Eltra, January 2002. www.eltra.dk
[43] E.ON Netz GmbH, Grid Code: High and extra high voltage, August 2003.
[44] F. Santjer, New Supplementary Regulations for Grid Connection by E.ON Netz GmbH, DEWI
Magazin, Nr. 22, February 2003.
[45] Commission for Energy Regulation, Wind Generator Connection Policy, 9th July, 2004.
www.eirgrid.com
[46] ESB National Grid, WFPS1: Wind Farm Power Station Grid Code Provisions, July, 2004.
www.eirgrid.com
[47] ESB National Grid, Models Received and Tested by ESBNG, 20 December 2004.
www.eirgrid.com
[48] IEEE Task Force on Load Representation for Dynamic Performance, Standard Load Models for
Power Flow and Dynamic Performance Simulation, IEEE Trans. PWRS, August 1995.
[49] RED Electrica de Espana, Condiciones Tecnicas Aplicables a la Generacion de Regimen Especial
no Gestionable, April 2003.
[50] REE Annual Report 2003, www.ree.es
[51] Policy and Planning Guideline for Development of Wind Energy Facilities in Victoria
www.auswea.org
[52] Resources & Networks Branch, Ministry of Economic Development, Facilitating Distributed
Generation: A discussion paper, September 2003. www.windenergy.org.nz
[53] V. Akhmatov, Voltage Stability of Large Power Networks with a Large Amount of Wind
Power, Proceedings of 4th International Workshop on Large-Scale Integration of Wind Power
and Transmission for Offshore Wind Farms, October 29-21, 2003, Billund, Denmark.
[54] V. Akhmatov, Voltage Stability of Large Power Networks with a Large Amount of Wind Power,
Proceedings of 4th International Workshop on Large-Scale Integration of Wind Power and
Transmission for Offshore Wind Farms, October 29-21, 2003, Billund, Denmark. (Presentation
slides available on Eltra website)
[55] J. Niiranen, Voltage Dip Ride Through of a Doubly-Fed Generator Equipped with an Active
Crowbar, Nordic Wind Power Conference, 1-2 March 2004.
[56] BC Hydro, 69 kV to 500 kV Interconnection Requirements for Power Generators, January 2002.
[57] H. E. House and P. D. Tuttle, Current-Carrying Capacity of ACSR, AIEE Transactions, pp.
1169-1177, February 1959.
[58] M. Bahrman, E. Larsen, R. Piwko, H. Patel, Experience with HVDC Turbine Generator
Torsional Interaction at Square Butte, IEEE Transactions on Power Apparatus and Systems, Vol.
PAS-99, pp. 966-975, May/June 1980.
[59] N. Rostankolai, R. J. Piwko, E. V. Larsen, D. A. Fisher, M. A. Mobarak and A. E. Poitras,
Subsynchronous Torsional Interactions With Static Var Compensators - Concepts and Practical
Implications, IEEE Transactions on Power Systems, Vol. 5, No. 4, November 1990.
[60] C. T. Wu, K. J. Peterson, R. J. Piwko, M. D. Kankam and D. H. Baker, The Intermountain Power
Project Commissioning - Subsynchronous Torsional Interaction Tests, IEEE Transactions on
Power Delivery, October 1988.
[61] Electric Power Research Institute, HVDC Systems Control for Damping of Subsynchronous
Oscillations, EPRI EL-2708, Final Report, October 1982.
[62] IEEE Rotating Machinery Committee, IEEE Screening Guide for Planned Steady-State
Switching Operations to Minimize Harmful Effects on Steam Turbine-Generator, IEEE Trans.
PAS, pp. 1519-1521, 1980.
[63] P. Kundur, Power System Stability and Control, McGraw-Hill, 1994.
[64] IEC Standard 1000-3-7, Electromagnetic Compatibility: Part 3 Limits Section 7. Assessment
of emission limits for fluctuating loads in MV and HV power systems, 1996.
[65] B. K. Parsons, Y. Wan and B. Kirby, Wind Farm Power Fluctuations, Ancillary Services, and
System Operating Impact Analysis Activities in the United States, Presented at the European Wind
Energy Conference, Copenhagen, Denmark July 2-6, 2001.

78

[66] J. Lewis Blackburn, Symmetrical Components for Power Systems Engineering, Marcel Dekker,
1993.
[67] C. F. Wagner, Self-Excitation of Induction Motors with Series Capacitors, AIEE Transactions,
pp.1241-1247, Vol. 60, 1941.
[68] M. Milligan and B. Parsons, A Comparison and Case Study of Capacity Credit Algorithms for
Intermittent Generators, Presented at Solar97, Washington DC, April 27-30, 1997.
[69] M. J. Gibbard, Co-ordination of Multimachine Stabiliser Gain Settings for a Specified Level of
System Damping Performance, IEE Proc. Part-C, pg 45-48, March 1982.
[70] F. P. De Mello and C. Concordia, Concepts of Synchronous Machine Stability as Affected by
Excitation Control, IEEE Trans. PAS, pg 316-329, Apr. 1969.
[71] P. Pourbeik and M. J. Gibbard, Damping and Synchronizing Torques Induced on Generators by
FACTS Stabilizers in Multimachine Power Systems, IEEE Trans. PWRS, November 1996, pages
1920-1925.
[72] P. Pourbeik and M. J. Gibbard, Simultaneous Coordination of Power System Stabilizers and
FACTS Device Stabilizers in a Multimachine Power System for Enhancing Dynamic
Performance, IEEE Trans. PWRS, May 1998, pages 473-479.
[73] M. Klein, G. J. Rogers and P. Kundur, A Fundamental Study of Inter-Area Oscillations in Power
Systems, IEEE Trans. PWRS, pg 914-921, August 1991.
[74] United State of America Federal Energy Regulatory Commission, Interconnection for Wind
Energy and Other Alternative Technologies, January 24, 2005, Docket No. RM05-4-000.

79

You might also like