You are on page 1of 218

Wind Power Integration:

from Individual Wind Turbine to Wind Park as a Power Plant

Yi Zhou

Wind Power Integration:


from Individual Wind Turbine to Wind Park as a Power Plant

PROEFSCHRIFT

ter verkrijging van de graad van doctor aan de Technische Universiteit Delft, op gezag van de Rector Magnificus prof. dr. ir. J. Fokkema, voorzitter van het College voor Promoties, in het openbaar te verdedigen op woensdag 14 oktober 2009 om 10.00 uur door

Yi Zhou
Master of Science, Delft University of Technology geboren te Hefei, Anhui, China

Dit proefschrift is goedgekeurd door de promotor: Prof. dr. J. A. Ferreira Copromotor: Dr. ir. P. Bauer

Samenstelling promotiecommissie: Rector Magnificus, voorzitter Prof. dr. J. A. Ferreira, Technische Universiteit Delft, promotor Dr. ir. P. Bauer, Technische Universiteit Delft, copromotor Dr. J. Pierik, Energieonderzoek Centrum Nederland Prof. ir. L. van der Sluis, Technische Universiteit Delft, Prof. dr. G. J. W. van Bussel, Technische Universiteit Delft Prof. dr. E. Lomonova, Technische Universiteit Eindhoven Prof. dr. L. Bertling, Chalmers University, Goteborg, Sweden

ISBN 978-90-9024673-4 Printed by Sieca Repro, Delft, The Netherlands.

To my family

Acknowledgements
This research has been carried out at the Delft University of Technology in the Netherlands, in the group of Electrical Power Processing (EPP). Financial support by We@Sea and PhD@Sea organization is gratefully acknowledged. This thesis could not be finished without the help and support of many people. I would like to take this opportunity to express my gratitude to them. First of all, I wish to express my deep and sincere gratitude to my Promoter Professor Braham Ferreira. I am truly grateful to him for trusting in my ability to complete this work and for his valuable ideas and suggestions during this work. I wish to extend my sincere thanks to my supervisor Dr. Paul Bauer, with whose guidance I could have worked out this thesis. The time we spent on discussions over this research and many other topics are most valuable and unforgettable. His patience and kindness are greatly appreciated. I am also extremely grateful to Mr. Jan Perik in the Energy Research Center of the Netherlands (ECN). His technical advice and guidance is indispensable through out this thesis. I also thank my doctoral examination committee for studying the draft of the thesis and giving their valuable comments. I would like to thank my friends and colleagues in the EPP research group. I owe special thanks to Dongsheng Zhao, Zhihui Yuan, Deok-Je Bang, Ghanshyam Shrestha, Yi Wang, Balazs Czech, and Juan Gutierrez Alcaraz for their friendship which makes my life in Delft enjoyable. I would like to thank Johan Morren, who helped me on the simulation at the beginning of this research. I especially would like to thank Rick van kessel and Ralph Hendriks, who translate the summary and propositions into Dutch. I would also like to thank my former colleagues and friends in the State Grid Corporation of China, especially Yuping Qiu, Zhongyuan Chen, Di Wu, and Haoran Hu. I owe special thanks to my former mentor Mrs. Yuping Qiu, whose supervision and inspiration help me to overcome the difficulties in the beginning of my professional career. Last but not the least, I would like to thank my parents, my wife, my brother, and my sister in law, to whom this thesis is dedicated, for their love and support throughout my life. I would specially thank my dad. Without his encouragement and stimulation, I could never make so far. I am also extremely grateful to my wife for her understanding and continuous support during all these years. Finally, to my dearest daughter, who is my driving force to accomplish this work.

vii

viii

Table of Contents
Chapter 1 Introduction
1.1 Introduction 1.2 Problem definition 1.3 Research Method 1.4 Thesis Layout 1.5 References

1
1 2 5 7 9

Chapter 2 Wind Turbine Concepts and Control


2.1 Introduction 2.2 Wind Turbine Concepts 2.2.1 Wind Energy Basics 2.2.2 Wind Turbine Components 2.2.3 Wind Turbine Trends 2.3 Wind Turbine Control 2.3.1 Aerodynamic Torque Control 2.3.2 Generator and Converter Control 2.3.3 Summary of FOC and DTC 2.4 Integration Problems 2.4.1 Static Problems 2.4.2 Power Quality 2.4.3 Transient Stability 2.4.4 Small-Signal Stability 2.4.5 Sub-Synchronous Oscillation 2.5 Conclusion 2.6 References

11
11 11 11 12 12 16 16 17 28 29 29 30 31 32 32 32 33

Chapter 3 Wind Generator System Modelling


3.1 Introduction 3.2 Dynamic Model 3.2.1 Wind Speed and Rotor 3.2.2 Shaft

37
37 37 38 38

ix

3.2.3 Generator 3.2.4 Power Electronic Converter 3.2.5 DC link 3.2.6 Phase Locked Loop 3.2.7 Grid 3.2.8 Interface between Grid and Unit 3.2.9 Model Implementation 3.2.10 Comparing FOCs and DPCs 3.2.11 Verification of Dynamic Model of Type C Wind Turbines 3.3 Small-Signal Model 3.3.1 Changes in Flux, Torque and Power 3.3.2 Aerodynamic Torque 3.3.3 Shaft 3.3.4 Wound Rotor Induction Generator 3.3.5 Converter 3.3.6 Generator VSC Controller 3.3.7 Grid VSC Controller 3.3.8 DC link and DC Voltage Controller 3.3.9 Phase Locked Loop 3.3.10 Interface between Unit and Grid 3.3.11 Power Grid Model 3.3.12 Model Implementation 3.3.13 Verification of Small-Signal Model of Type C Wind Turbines 3.3.14 Stability Analysis of Type C Wind Turbines 3.4 Conclusion 3.5 References

40 42 50 52 53 53 54 54 57 58 59 59 59 60 61 61 63 64 64 64 65 66 68 72 81 82

Chapter 4 Grid Integration of Individual Variable Speed Wind Turbine


4.1 Introduction 4.2 Unbalanced Grid-Voltage Conditions 4.2.1 Symmetrical vs. Unsymmetrical Voltage 4.2.2 Symmetrical Component and Instantaneous Reactive Power Theory 4.2.3 Real-Time Signal Processing 4.2.4 Dual-Sequence Rotor VSC Current Controllers 4.2.5 Dual-Sequence Grid VSC Current Controllers 4.2.6 Simulation Results x

85
85 86 87 89 90 92 93 94

4.3 Operation of Low-Inertia Power Grid 4.3.1 Comparison of Large-Inertia and Low-Inertia Power Grids 4.3.2 Analysis of the Low-Inertia Power Grid 4.3.3 Operation of Low-Inertia Power Grid with Central Control Unit 4.3.4 Operation of Low-Inertia Power Grid with Distributed Control Unit 4.4 Conclusions 4.5 References

98 99 102 103 108 111 113

Chapter 5 Wind Park Conceptual Design and Control


5.1 Introduction 5.2 Wind Park Integration Choice 5.2.1 HVAC Cable Transmission Capability 5.2.2 Offshore Platform 5.2.3 Case Study 1 Comparison of HVAC cable and HVDC VSC 5.2.4 Case Study 2 Comparison of HVDC VSC and HVDC LCC 5.3 Hybrid Wind Park Optimisation 5.3.1 Optimisation Method 5.3.2 Simple Steady-State Model 5.3.3 Optimisation Results 5.4 Wind Park Controller 5.4.1 Central Controller Concept 5.4.2 Case Study 5.4.3 Simulation Results 5.5 Conclusion 5.6 References

117
117 118 120 124 125 126 128 129 131 139 145 145 147 147 150 152

Chapter 6 Power System with Large Wind Power Installation


6.1 Introduction 6.2 Frequency Response 6.2.1 Frequency Response in Traditional Power Plants 6.2.2 Frequency Response in Wind Power Plants 6.3 Small-signal Stability 6.3.1 Multi-machine Power System 6.3.2 Model Implementation 6.3.3 Low Frequency Oscillation Modes

155
155 156 156 158 161 162 162 163

xi

6.3.4 Power System Stabiliser of Wind Power Plant 6.4 Limitation of Inertia Usage 6.5 Conclusion 6.6 References

166 169 171 173

Chapter 7 Conclusions and Recommendations


7.1 Conclusions 7.2 Recommendations

175
175 178

Appendix. A List of Symbols and Abbreviations


A. 1 Symbols A. 2 Subscripts A. 3 Notations A. 4 Abbreviations

179
179 180 180 181

Appendix. B Structures of Closed Loop Controlled Type C and D VSWT 183 Appendix. C Wind Turbine Parameters
C.1 Type A Wind Turbine C.2 Type C Wind Turbine C.3 Type D Wind Turbine

185
185 185 186

Appendix. D Two Area Four Machine System


D.1 Topology D.2 Bus Data D.3 Transmission Line Data D.4 Synchronous Generator and Excitation Data

189
189 189 190 191

List of Publications Summary Samenvatting

193 195 199 203

xii

CH.1 Introduction

Chapter 1
Introduction

1.1 Introduction
Wind energy is one of the most prominent renewable energy sources today. The increasing concerns with environmental issues are driving the search for more sustainable electrical sources. Wind energy along with solar energy, biomass and wave energy are possible solutions for environmental-friendly energy production. The initialization of wind power installation, which started in the beginning of 1980s, is very much related to the oil crises of the mid 1970s. During that period, a simple and robust wind turbine concept emerged and became very popular, which marked the beginning of the wind power industry. This simple and robust concept includes a three-bladed wind turbine rotor, a gearbox, and an induction generator directly connected to the grid. This early model was cheap and very robust, but the power quality was poor, and it has negative influence on the voltage stability of the grid. During the 1980s, most wind power installations were limited to a few hundred kilowatts. The small size of those installations did not threaten the power system stability. The 1990s marked an important breakthrough in the industry. New concepts emerged because of the demand for more efficient power production and because of the necessity to comply with power quality requirements. During the 1990s, wind turbines (and parks) grew in size and production, from just a few hundred kilowatts to megawatts. By the end of 2007, the wind generation in Europe had reached approximately 56GW. And the push to connect large-scale wind parks to the power grid is still growing. The 20% goal that wind power installation reaches 20% by 2010 no longer seems like an unattainable dream. However, the boost of wind power has become a challenge for the transmission system operator. This thesis focuses on grid integration aspects such as dynamic behavior of the wind power during disturbance, and dynamic behavior of power system with large wind power integration.

1.2 Problem definition

1.2 Problem definition


Successful operation of a power system is based on a progress that took decades to fully understand the characteristics of the dominating generator concept - synchronous generator (SG). For example, frequency deviation is often believed to be related with the active power balance. This phenomenon is due to the swing function of alternating current (AC) directly connected SG. For a power grid dominated by variable speed wind turbines (VSWT) whose rotation speeds are completely decoupled with the power grids frequency, the frequency deviation may not be related with active power balance anymore. Inspired by such considerations, the following characteristics of wind power plant are listed, which raise the research questions of this thesis: The rotation speed of variable speed wind turbines (VSWT) is completely decoupled from power grid frequency. Without an appropriate controller, total inertia of the system decreases as wind power installation increases. A wind power plant can also contribute to primary frequency and even secondary frequency support; however, the plants capability depends on its operation points and operation modes; Existing wind farm often uses long AC cable instead of AC overhead line, however the AC cable has much shorter transmission distance limitation. The limitation is caused by the reactive power produced by the large shunt capacitance. For the large wind farm and long transmission distance, high voltage alternating current (HVAC) cable is not feasible; The control of VSWTs is different from that of large synchronous generators. This may shape the eigen values of the entire system, and thus change the small-signal stability of the system; Both symmetrical and unsymmetrical faults are critical to wind power plants, while only symmetrical faults are considered critical to thermal power plants because of the influence on the short circuit current and transient stability of the synchronous generator [Kundur'94]; A wind power plant is composed of many small units, while a thermal power plant is often made up of several large units;

Due to the above analysis of characteristics of wind power plant, the following research questions arise: Does the decreasing inertia of the system introduce any stability problems? Is necessary to use wind turbine inertia for primary frequency support? In extreme conditions, if there is no inertia in the system, such as an isolated power grid dominated by voltage source converters (VSCs), is it possible to coordinate the use 2

CH.1 Introduction of a wind farm with other power sources to control the frequency and voltage of this isolated system in the case of a main power grid black out? Which method is the best for grid integration of a specific wind power plant with particular power rating and integration distance requirements? What criteria are crucial for comparing the integration methods between the HVAC, the high voltage direct current (HVDC) with line commutated converter (LCC), and the HVDC with voltage source converter (VSC)? How does a wind power plant change the small-signal stability of the whole system? Is it necessary and possible to install power system stabiliser (PSS) in a wind power plant? What is the difference between symmetrical and unsymmetrical faults, and how can wind turbines be controlled to successfully ride through both faults? Which method can be used to coordinate those many small wind turbines in the wind farm to behave as a single large unit? In addition to the above research questions, this thesis also deals with some aspects of how to design a wind farm: Is it possible to use an optimisation method to optimise the configuration of the wind farm to improve the reliability, power quality, voltage stability and etc while

limiting the investment of wind farm and power grid? The above questions represent the objectives of this thesis and result in the following issues. The detailed answers and studies will be dealt within the chapters of this thesis. These issues address: Frequency response with wind power plant In variable speed wind turbines (VSWT), the generator rotation speed is completely decoupled from the power grid frequency, which changes the frequency response of the power system. The frequency changes faster in a system with small inertia, which may lead to unacceptable frequency deviation. Even though the primary source of wind power cannot be controlled as power source of a thermal power plant, its inertia can be used for primary frequency support for several seconds. There are some fundamental limitations to using inertia, such as speed limitation, stability limitation and VSC capacity limitation, all of which need further study. Furthermore, in a low-inertia power grid, if no synchronous generator is directly connected to the system, the frequency will not vary during an active power imbalance and voltage will not vary during a reactive power imbalance. The traditional frequency and voltage droop controllers have to be studied. Integration choice The integration choices between an HVAC, HVDC LCC and HVDC VSC are compared regarding technical and economical aspects. The critical integration

1.2 Problem definition distance and critical power rating of a wind park at a certain voltage level will be calculated to illustrate the selection of an integration method. Small-signal stability The small-signal stability of a wind power plant is considered to be better than that of a thermal power plant. Constant speed wind turbines (CSWT) use an induction generator, which provides larger damping [Slootweg and Kling'03]. VSWTs do not see the power system frequency, thus it is not likely that VSWTs will participate in the oscillation of the power system. Moreover, it also provides additional damping with its power-speed curve. Generally speaking, with wind power penetration, the small-signal stability of the power system will be improved. However, a quantitative small-signal stability analysis is still required for the power system with many variable speed wind turbines. As with the power system stabiliser (PSS) of a thermal power plant, the wind power plant can also actively damp out the existing oscillations in the power system. The wind power plant can help to damp out the power system oscillations by either using the kinetic energy stored on its shaft or using its reactive power capability same as static var compensator (SVC). Instantaneous active and reactive power control under unbalanced system In an unbalanced three-phase system, voltages and currents have double fundamental frequency oscillations. These oscillations can have a negative effect on the power electronics and generators. For example, large DC voltage ripple exists in the DC link of back-to-back VSCs, hot spots on the stator and rotor windings of generators appear, and the mechanical stress of generator caused by the pulsating torque. It is necessary to develop proper controllers to mitigate these negative effects. Wind farm optimisation - The types of wind turbines along with the electrical configurations of wind farm are designed in such a way that the wind farm meets the technical requirements while economically performing well. Integral wind farm controller The master controller of a wind farm is compared to an individual wind turbine controller. The reactive power control can, for example, be implemented by voltage droop control of individual wind turbines, or by the master reactive power controller, which distributes total reactive power demand to individual wind turbines. Both controllers have their advantages and disadvantages. The individual wind turbine controller is simple and fast. However it cannot control the total reactive power of the wind farm at the point of common coupling (PCC) precisely. The master controller can control total reactive power at the PCC more accurately and it can dispatch the total reactive power demand to the wind turbines according to their reactive power capability. However it is more complex and slower.

CH.1 Introduction The purpose of this thesis is not to introduce new electrical system of wind turbine, but to improve the performance of the wind farm and power grid with existing wind turbine concepts, both under normal operation and disturbances.

1.3 Research Method


Four popular wind turbine concepts [Ackermann'05] are drawn in Figure 1-1, which are: Type A A CSWT with a squirrel cage induction generator (SCIG), soft starter and shunt capacitor bank; Type B A VSWT with a wound rotor induction generator (WRIG), variable rotor resistance, soft starter and shunt capacitor bank; Type C A VSWT with a WRIG and partial scale back-to-back VSC on the rotor (doubly fed induction generator - DFIG); Type D A VSWT with a permanent magnet generator (PMG) and a full-scale back-to-back VSC on the stator. In this thesis, only type A, C, and D as shown in Figure 1-1 are modelled and analyzed in detail, because of their dominant market share (> 95%), as will be discussed in Chapter 2. Simulation is considered as the main method to study the proposed research topics. Wind turbines and power electronic converters are simulated in such a way that they are as simple as possible to improve the simulation speed, while still having enough detail for the specific simulation purpose.
Table 1-1. Appropriate models for studies of different integration problems [Akhmatov'03]. Integration Problems Model Blocks Wind speed Rotor Static voltage and frequency Detailed Steady state Flicker Harmonics LVRT

Detailed Dynamic inflow, 3p effect* Unimportant

Unimportant Unimportant

Unimportant Constant power

Shaft

Lumped mass

Unimportant

Generator

Third order

Third order

Third order

Converter

Average

Average

Switching

Type A: two mass Type D: two mass Type D: lumped mass Type A: third order Type C: fifth order Type D: third order Average

* 3p effect is the phenomenon that the aerodynamic torque of the wind turbine decreases whenever the wind turbine blades passing through its tower.

1.3 Research Method Table 1-1 lists the required model details for different integration problem studies. For example, for flicker studies, both wind speed and rotor model have to use a detailed model including the turbulence, dynamic inflow and 3p effect, but for LVRT capability studies the wind speed and rotor model can be omitted and input for the shaft can be replaced with a constant mechanical torque.
Type A

M.Gear

G SCIG Soft starter Capacitor Bank

Grid

(a)
Type B Variable rotor resistance

M.Gear

G WRIG Soft starter Capacitor Bank

Grid

(b)
Type C

Partial VSC M.Gear G WRIG Grid

(c)
Type D

G PMG Full VSC

Grid

(d)
Figure 1-1. Four popular wind turbine concepts, (a) Type A wind turbine; (b) Type B wind turbine; (c) Type C wind turbine; (d) Type D wind turbine.

CH.1 Introduction

1.4 Thesis Layout


This section presents an outline of the thesis. The layout of this thesis is illustrated in Figure 1-2. The work in this thesis is in a bottom-up approach, starting with concepts for individual wind turbines, including control and modelling, followed by a conceptual wind park design and control, and finally on the highest level, wind power plant support aimed at improving power system performance.

Figure 1-2. Thesis layout.

Chapter 2 reviews the existing wind turbine concepts, and explains the control of individual wind turbines, including control strategies as connection to a strong, weak or islanded power grid; and control methods such as field-oriented current controllers and direct torque/power controllers. Chapter 3 develops dynamic and small-signal models of individual wind turbines. A dynamic model of a DFIG is calibrated through experimental results. A DFIG small signal model is also verified. The dynamic simulations of two control methods with different modelling approaches are compared. The compared simulations are field-oriented controller (FOC) with an average model, FOC with a switching model, and a direct torque control (DTC) with a switching model. The stability of a single DFIG with an infinite bus system is analysed by calculating the eigen value locus of the system. Chapter 4 investigates the control problems of existing wind turbine concepts, and proposes solutions to these problems. The investigated control problems are VSWTs under an unbalanced grid voltage, and VSWTs in an isolated, low-inertia power grid. Chapter 5 investigates the conceptual design and control of a large wind farm, including the choice of integration methods, types of wind turbines, and topology (string or star connected). Integration methods as HVAC cable, HVDC LCC and HVDC VSC are 7

1.4 Thesis Layout compared. The case of a hybrid wind farm optimisation is studied. The objectives of the optimization are to have low levelised production cost, high reliability, an acceptable flicker level and LVRT capability. The benefits of a central controller for the wind park are discussed and a generic central wind park controller is designed. Chapter 6 investigates the influence of a large amount of wind power on power system, including frequency response and small-signal stability problems. Frequency responses of power system both with and without wind power penetration are compared. Local and inter-area power system oscillations of traditional thermal power plants and wind power plants are identified by calculating the eigen values of the system. PSS of wind power plant is designed to damping out the existing oscillations inside the power system introduced by SGs. The fundamental limitations of using the inertia for short-term primary frequency support and power system oscillation damping are investigated. Chapter 7 summarizes the results of this research, draws conclusions from this thesis, and proposes recommendations for future research.

CH.1 Introduction

1.5 References
[Ack05] [Akh03] T. Ackermann, Wind power in power systems, ISBN-0-470-85508-8, John Wiley & Sons, 2005. V. Akhmatov, Analysis of dynamic behavior of electric power systems with large amount of wind power, PhD thesis, Electrical Power Engineering Department, Technical University of Denmark, Copenhagen, Denmark, 2003. P. Kundur, Power System Stability and Control, ISBN-0-07-035958-x, McGraw-Hill Professional, 1994. J. G. Slootweg and W. L. Kling, "The impact of large scale wind power generation on power system oscillations," Electric Power Systems Research, vol. 67, issue 1, Page(s). 9-20, 2003.

[Kun94] [Slo03]

1.5 References

10

CH.2 Wind Turbine Concepts and Control

Chapter 2
Wind Turbine Concepts and Control

2.1 Introduction
This chapter reviews concepts and controls of individual wind turbines. Firstly, the basics of wind energy and topologies of wind turbines are reviewed. Secondly, wind turbine control methods are explained. Finally, existing integration problems are discussed.

2.2 Wind Turbine Concepts


2.2.1 Wind Energy Basics
The kinetic energy of wind depends on the air density, i.e. mass per unit of volume. Only part of the energy in wind can be extracted by wind turbines. This is known as the Betz limit. The power that can be extracted by a wind turbine is defined as [Ackermann'05, Burton'01, Hau'06]:

1 R2 Cp (; )(Vw )3 (2.1) 2 where is air density, R is the wind turbine radius, Cp (; ) is the power coefficient, is P =
the tip speed ratio, is the pitch angle, and Vw is the rotors effective wind speed. According to the Betz limit [Burton'01, Hau'06], the maximum power coefficient is

Cp

16 27

0:59.

Wind speed at a specific location is not constant. The probability or frequency of one particular wind speed can be statistically predicted, which is approximated by the Weibull distribution function [Ackermann'05, Burton'01, Hau'06]:

=1e
where scaling factor, and

Vw avg kwb Awb )

(2.2) is the

is the probability of wind speed,

is the average wind speed,

is the form parameter. These parameters are derived by analysing a

reliable database of wind speed measurements, which have an evaluation period of at least several years.

11

2.2 Wind Turbine Concepts

2.2.2 Wind Turbine Components


Wind turbines consist of two main parts: the aerodynamic part, which extracts kinetic energy from the wind and converts it to a mechanical torque, and the electric generating system, which converts the mechanical torque into electric power. The main components of wind turbines are shown in Figure 2-1. Not all wind turbine concepts contain all the components. However, Figure 2-1 illustrates the general working principle of a wind turbine and highlights the necessary components to be included in the simulation model.

Figure 2-1. Generic wind turbine structure.

2.2.3 Wind Turbine Trends


The number and size of wind power installations have increased tremendously in recent years. The unit size also has increased. As extractable wind power is proportional to the square of the rotor radius, the up-scaling of a unit is concluded as an efficient approach. The unit size growth can be observed in Figure 2-2 [Gardner, et al.'04]. There are many wind turbine concepts on the wind energy market today [Ackermann'05, Blaabjerg and Chen'06, Hansen and Hansen'07, Hau'06]. The concepts available can be categorised based on rotation speed, shafts, and the electrical equipment. The rotation speed can be categorised as: Constant speed Variable speed Multi-stage gearbox Single-stage gearbox

The shafts can be categorised as:

Directly driven And the electrical equipment can be categorised as:

12

CH.2 Wind Turbine Concepts and Control

Without electronic converter Partial-scale converter Full-scale converter

Figure 2-2. Growth in size of commercial wind rurbine designs [Gardner, et al.'04].

Surveys of the top 30 wind turbine manufactures list 10 configurations [Ackermann'05, Blaabjerg and Chen'06, Hansen and Hansen'07, Shrestha, et al.'08], as shown in Figure 2-3. The 10 wind turbine concepts drawn in Figure 2-3 are: Type A A CSWT with a SCIG, soft starter and a shunt capacitor bank; Type B A VSWT with a WRIG, variable rotor resistance, soft starter and a shunt capacitor bank; Type C A VSWT with a WRIG and partial scale back-to-back VSC on the rotor - DFIG; Type D A VSWT with a PMG and a full-scale back-to-back VSC on the stator; Type E A VSWT with a PMG, gearbox and a full-scale back-to-back VSC (multibrid); Type F A VSWT with a PMG, diode rectifier and a full-scale VSC inverter; Type G A VSWT with a SG, DC excitation and a full-scale back-to-back VSC; Type H A VSWT with a SCIG, and a full-scale back-to-back VSC; Type I A VSWT with an AC directly connected the SG, and a torque converter, such as a hydro-mechanical gearbox; Type J A VSWT with a multiple generator system of PMG and a full-scale backto-back VSC.

13

2.2 Wind Turbine Concepts

Type A

Type F

M.Gear

G G SCIG Soft starter Capacitor Bank Type G Grid PMG Diode Full VSC

Grid

Type B

Variable rotor resistance

M.Gear M.Gear G WRIG Soft starter Capacitor Bank Type E Grid

G SG Full VSC

Grid

Type C M.Gear Partial VSC M.Gear G WRIG Grid Type I G SCIG Full VSC Grid

Type D M.Gear

Torque Converter

Grid

G PMG Full VSC

Grid

G Type J

Type H G Bull Gear Grid G PMG Full VSC Grid

S.Gear

Figure 2-3. Wind turbine concepts

In this thesis, only type A, C, and D as shown in Error! Reference source not found. are modelled and analyzed in detail, because firstly from the grid point of view, Types E, F, G, H, I, J have similar behaviour as Type D; and secondly Types A, C, D constitute more than 95% of todays wind turbine market [Ackermann'05, Hansen and Hansen'07].
Table 2-1. Annual world wind power market share of different wind turbine concepts 1995-2005. 1995 69.5 16.2 0.0 14.3 1161 85.7 1996 62.7 23.4 0.1 13.7 1092 95.7 1997 53.5 27.0 3.1 16.3 1483 96.2 1998 39.6 17.8 26.5 16.1 2345 92.7 1999 40.8 17.1 28.1 14.0 3788 94.0 2000 39.0 17.2 28.2 15.6 4381 96.3 2001 30.6 15.2 37.3 16.9 7175 100.0 2002 27.8 5.2 46.7 20.3 7242 97.4 2003 19.2 3.1 59.8 17.9 8084 100.0 2004 24.7 2.2 54.8 18.3 8247 97.7 2005 17.1 1.0 60.8 16.1 11531 95.0

A B C D Total MW Total %

14

CH.2 Wind Turbine Concepts and Control

Wind Turbine Market Share (1995-2005)


80 70

Yearly Market Share (%)

60 50 40 30 20 10 0 1995 1996 1997 1998 1999 2000 2001 2002 2003 2004 2005

Type A (%) Type B (%) Type C (%) Type D (%)

Year

Figure 2-4. Annual world wind power market share of four popular wind turbine concepts 1995-2005. The annual wind power installation and market share of different wind turbine concepts are recorded in Table 2-1 and Figure 2-4. Table 2-2. Comparison of different wind turbine concepts. Concept A B C Variable speed No Limited Yes Control active power No Limited Yes Control reactive power No No Yes Reactive power compensation needed Yes Yes No Soft starter needed Yes Yes No Short circuit power contribution Yes Yes Limited Control bandwidth 1-10s 100ms 1ms Flicker production Yes Yes No Inertia contribution Limited Limited Yes Islanding operation No No Yes Investment ++ ++ + Maintenance ++ ++ ++ Very positive (very low cost); + Positive; - not competitive in cost (expensive)

D Yes Yes Yes No No Limited 0.5-1ms No Yes Yes +

Different wind turbine concepts have different advantages and disadvantages. Fixed-speed wind turbines have drawbacks such as uncontrolled reactive power consumption, large mechanical stress, and low power quality. As the total capacity of wind power increases, these drawbacks prevent further installation of wind power sources. Nowadays, variable speed wind turbines are the dominant wind turbine concepts. They are designed and controlled to achieve maximum aerodynamic efficiency over a wide range of wind speeds. To control the generator speed and reactive power production, induction or synchronous generators are used and partial-scale or full-scale power electronic converters are supplied. The advantages of variable speed wind turbines are increased aerodynamic efficiency,

15

2.3 Wind Turbine Control

reduced mechanical stress, and improved power quality and power factor. The introduction of variable speed wind turbines has facilitated an increasing number of wind power installations. The advantages and disadvantages of Type A, B, C and D wind turbines are compared in Table 2-2.

2.3 Wind Turbine Control


The control of wind turbine includes two parts: one part controls the aerodynamic torque, and the other part controls the output of both active and reactive power to fulfil the grid codes.

2.3.1 Aerodynamic Torque Control


In order to control the mechanical power of wind turbines, we have to control the torque force . Four methods can be used to control the torque force , namely: passive stall control, active stall control, pitch angle control, and yaw control [Bianchi, et al.'07, Hau'06, Hoffmann'02].

Figure 2-5. Aerodynamic torque controls (a) passive stall control, (b) active stall control, (c) pitch control.

The yaw control changes the angle between the facade of wind turbine and the direction of the actual wind. The passive stall control, active stall control and pitch angle control are

16

CH.2 Wind Turbine Concepts and Control

used in large wind turbines to change the mechanical power input. Figure 2-5 illustrates the mechanisms of these three control methods [Hoffmann'02]. Although the active stall control may physically resemble the pitch angle control, actually they are quite different. In To initiate stall effect, the two stall controls use different means to increase the attack angle , thus reducing the lift force decreases the attack angle and torque force . The pitch angle control to limit all the aerodynamics forces. The pitch control rotates

the blades in the opposite direction of the active stall control. The active stall control works like a barrier to the wind, thus the thrust force which tries to turn down the turbine increases. The increased thrust force limits the life time of a wind turbine. All three methods have their advantages and disadvantages. They are compared in Table 2-3.

Table 2-3. Comparison of passive stall control, active stall control and pitch angle control. Controls Passive stall control Advantages Does not require moving parts, and simplifies the control system. Disadvantages Requires complex aerodynamic design, may cause vibrations due to the stall effect; Only designed to limit the power when wind speed is above the rated speed. The influence of thrust force on the blade is significant, thus the turbine has a short lifetime. The change in speed of angle has to be large.

Active stall control

Pitch angle control

The change speed of blade angle can be small; The input mechanical power can be changed when necessary. All forces on the blade are limited, thus the turbine has a long lifetime; The mechanical power input can be changed when necessary.

2.3.2 Generator and Converter Control


Type A wind turbines (i.e., CSWT) have very limited control possibilities. The active power can only be controlled when the wind speed is faster than the rated wind speed, using the aerodynamic control methods discussed in 2.3.1. The reactive power cannot be controlled by the wind turbine itself, thus an additional reactive power compensator is required, such as a shunt capacitor bank, a static var compensator (SVC), or a static compensator (STATCOM). For VSWT, both the active and reactive power can be independently controlled using Type C and Type D wind turbines. In general, the generator-side VSC controls the generator

17

2.3 Wind Turbine Control

speed (or electric torque)\, and the flux level (or stator reactive power), and the grid-side VSC controls the DC voltage level and terminal AC voltage (or VSCs reactive power). In this section, the electrical controllers of VSWTs will be discussed. Firstly, the control strategy of VSWTs will be explained. It generates the reference active and reactive power of VSWTs and is therefore defined here as the control strategy. Secondly, the actual control methods of VSWTs will be described, which include the generator control and power electronic converter control. Two control methods, namely the field-oriented control (FOC) and the direct torque/power control (DTC/DPC), will be explained in detail later in section 2.3.2.3 and section 2.3.2.4.

2.3.2.1 Control Strategies


The VSWT can be integrated into a strong power grid, or within an islanded grid as main power source. These two situations define the two control strategies.

Strong Power Grid


When wind turbines are connected to a strong power grid, it is often a requirement that the maximum amount of active power is produced, or that the active power is produced as designated by the grid code. The maximum power production mode (MPP) is the most often adopted control strategy of wind turbines. Since the effective rotor wind speed cannot be measured very accurately, the mechanical speed of either the turbine or the generator is measured to calculate the active power reference using the so-called power-speed curve [Akhmatov'03, Bianchi, et al.'07, Burton'01, Hau'06]. The reactive power reference can be set to 0, or it can be calculated from terminal AC voltage with a voltage droop controller. In addition, a frequency droop control loop can be implemented to increase short-term primary frequency support using the kinetic energy.

Islanded Power Grid


In remote areas, the local power grid often has only one transmission line connected to the main power grid. It is very possible that the local power grid disconnects from the main power grid occasionally. During the disconnection period, the local power grid still has to operate sufficiently to maintain local voltage and frequency. This is called as the islanded operation. There are two main control strategies to operate the islanded power grid. (1) Use one main power source to control the voltage and frequency of the grid. The advantage is that the control implementation is quite straight-forward. The disadvantages are: the grid is not flexible, DC units cannot be easily added or removed; and the main

18

CH.2 Wind Turbine Concepts and Control

power source has to switch between grid-connected mode and islanded mode. A fast islanding-detection method is required to transition between the two modes seamlessly. (2) Use multiple power sources to control voltage and frequency. The advantage is that reliability is high because more than one power source carries out the control. The disadvantage is that the control is not easy to implement. As the local grid often has many small DG units, the second strategy is more realistic. However there are conflicting views on control implementation. Some literature suggests using a master controller to control the grid voltage and frequency and distribute the active and reactive power requirements to all the units. However, this requires communication between the various DG units, but the micro-grid does not allow the DG units to be integrated or removed freely. A better solution is to remove the master controller, thus eliminating the communications between DG units, and to use only local signals to control the DG units. The designation of this controller will be discussed in Chapter 4.

2.3.2.2 Control Methods


The VSWT control methods include using a field-oriented controller (FOC) and direct torque controller (DTC). There are also other advanced controllers such as the gainscheduled adaptive controller and the fuzzy-logic controller [Bose'02]. However, these advanced controllers were developed to deal with special control problems, such as nonlinearity and machine parameter uncertainty. Their control mechanisms are the same as the FOC or DTC, and therefore will not be treated in this thesis.

2.3.2.3 Field Oriented Controller


FOC Concept
The field-oriented controller controls the AC machine the same way as a DC machine [Bose'02, Kazmierkowski, et al.'02, Vas'98]. Similar to the expression of the electromagnetic torque of a separately-excited DC machine, the instantaneous electromagnetic torque of an induction motor can be expressed as the product of a fluxproducing current and a torque-producing current, if a synchronously rotating flux-oriented reference frame is used, i.e. if an FOC is employed. In this case, the stator and rotor-current components are transformed into a new rotating reference frame which rotates together with a selected flux-linkage space vector. Figure 2-6 illustrates the similarity between a DC machine and an FOC-controlled AC machine. In Figure 2-6 (b), the stator currents are

19

2.3 Wind Turbine Control

aligned with rotor flux r, where id controls the magnitude of the flux similar to If of the DC machine, and iq controls the torque similar to Ia of the DC machine.

Figure 2-6. (a) DC machine (b) FOC-controlled AC machine.

In an FOC, the rotating angle e of the reference frame is crucial. Depending on how e is obtained, FOCs are divided into two categories: (1) the direct FOC, which measures or estimates the flux position e; and (2) the indirect FOC, which uses a machine model to R calculate e , for example, e = (wr wslip )dt for the generator mode. Depending on how the generator speed signal !r is obtained, indirect FOCs can also be divided into two categories: (1) FOCs with a speed sensor, which use an encoder to measure the speed; and (2) sensor-less FOCs, which use voltages and currents to estimate the speed. In this thesis, it is assumed that all the voltages, currents or speeds are accurately measured. Methods to estimate flux and speed can be found in [Bose'02, Vas'98]. The circuit vector diagram of an induction generator is shown in Figure 2-7 (a). At steady state, the derivative terms Figure 2-7 (b).
~ ds dt

= 0 and

~ dr dt

= 0, the space vector diagram of the rotor flux,

the air gap flux, the stator flux and the grid flux of the induction generator can be seen in

20

CH.2 Wind Turbine Concepts and Control

(a)

(b)
Figure 2-7. (a) Equivalent circuit (b) Space vector diagram at steady state.

In Figure 2-7 (b), the stator voltage is chosen as the reference frame. At steady state, the rotor flux r, stator flux s, and air gap flux m all rotate at the same synchronous speed

!e. However, because of the existence of stator resistance, stator leakage inductance and
rotor leakage inductance, there will be small angle differences between them. The choice of the reference frame is arbitrary in FOCs. The reference frame can be chosen to be aligned with the stator flux, rotor flux or grid flux, or the air gap flux, depending on how easily the flux position angle e can be calculated. However, sometimes incorrect choices will cause a coupling effect which has to be compensated by decoupling compensation current [Bose'02, Vas'98]. For example, a squirrel cage induction machine (SCIG) with a VSC on stator should choose the rotor flux reference. Choosing the stator flux reference will cause a cross-coupling term, as shown in (2.3) [Bose'02]: dds dids ds + Tr = Ls ids + Ls Tr Ls Tr !sl iqs (2.3) dt dt

21

2.3 Wind Turbine Control

where Tr =

Lr Rr

and = 1

L2 m Lr Ls

. The term Ls Tr !sl iqs is the cross-coupling term,

which means that the stator flux ds is not only controlled by stator current ids, but it is also influenced by iqs. Type C wind turbines have a VSC on the rotor of the WRIG, and the stator is directly connected to the power grid. Therefore it is normal to select the stator flux as the reference frame. Type D wind turbines have a VSC on the stator of the SG, so the rotor flux reference frame is often chosen. For grid-side VSCs, the grid-voltage or grid-flux orientation is chosen. The grid-flux orientation can also be chosen for Type C wind turbines because the stator resistance is usually small and can be omitted. FOCs can be divided into three parts; the reference-current generation, the current controller, and VSC modulation. A general structure of a generator-side FOC is shown in Figure 2-8, and the control equations are represented by equations (2.4) to (2.7).
vq

Te

id *

id *

id

vd
vd * vd *

Te

*
m

iq*

iq*

iq

vq*

vq*

Figure 2-8. (a) FOC reference current generation (b) FOC current controller (c) VSC modulation.
i = kp (Te Te ) + ki d

(Te Te )

(2.4) (2.5) (2.6) (2.7)

Z i = kp ( ) + ki ( ) q Z vd = kp (i id ) + ki (i id ) !q d d Z vq = kp (i iq ) + ki (i iq ) + !d q q

22

CH.2 Wind Turbine Concepts and Control

Reference Current Generation


The reference currents of generator FOCs are derived from an electric torque/mechanical speed controller, and from a generator flux/reactive power controller. For grid-side VSCs, the reference currents are derived from an active power/DC link voltage controller, and from a reactive power/AC voltage controller.

Current Controller
After obtaining the reference currents i i, a current controller is required to control the d q VSC to output the desired currents. In order to compensate the cross-coupling between id and iq , feed-forward decoupling terms are added after the PI controllers. The decoupling terms, which are sensitive to parameter variations, are !e Liq and !e Lid . In grid-side VSC FOCs, additional feed-forward terms of grid voltage vd and vq can be added.

FOC of Type C and D Wind Turbines


The full structure of Type C and Type D wind turbine FOC controllers are shown in Appendix A. They include both the generators FOC and the grid-side VSCs FOC.

2.3.2.4 Direct Torque/Power Controller


DTC Concept
Besides the FOC discussed in section 2.3.2.3, instantaneous torque controls with fast torque response can also be obtained by employing direct torque control (DTC) [Bose'02, Kazmierkowski, et al.'02, Lai and Chen'01, Malinowski, et al.'04, Noguchi, et al.'98, Vas'98]. In a DTC-controlled machine, flux linkage and electromagnetic torque are controlled directly and independently by the selection of optimum inverter switching modes. Unlike FOCs, which use the direct and quadrature currents id iq to control torque and flux, DTC controls the flux and torque errors within the hysteresis bands. The required optimal switching-voltage vectors can be obtained by simple physical considerations involving the position of the flux-linkage space vector, the required torque and flux linkage, and the available switching vectors. DTC does not use PI controllers and feed-forward decoupling terms, thus it is a fast control that is insensitive to parameter variations. The electric torque equation can be expressed as the cross production of rotor and stator fluxes. Lm Lm ~ ~ Te = p r = p js j jr j sin (2.8) 2 ) s Lr (Ls Lr Lm Lr (Ls Lr L2 ) m 23

2.3 Wind Turbine Control

The linearising equation (2.8) gives the torque control function as (2.9). The torque can be directly controlled by angle between s and r.

Te = kt s0 r0

(2.9)

Here a generator with a stator VSC is used to illustrate the DTC mechanism. According to Figure 2-7 (a), the following stator voltage vector equation in stationary reference frame can be obtained by: ~ ds ~ ~ = Vs + Rs Is (2.10) dt The stator voltage reference frame choice eliminates the speed term j!e s, as shown in Figure 2-7 (a). By linearising (2.10) and neglecting the second order terms and stator resistance, we can obtain the flux change equation: ~ ~ s = Vs t (2.11)

Thus, the flux vector change is determined by the selected voltage vector and applied time, ~ as shown in Figure 2-9 (a). By selecting voltage vector V6, the stator flux increases. Angle

decreases, which implies the decrease in electric torque. A two-level VSC has six active ~ ~ ~ ~ voltage vectors V1 ! V6 and two zero vectors V0, V7 , as shown in Figure 2-9 (b).
A clarification should be made between the DTC concept in Figure 2-9 (a) and FOC concept in Figure 2-7 (b). In the DTC concept, the chosen stationary reference frame eliminates the speed term j!e s , while in the FOC concept the synchronously rotating reference frame is chosen.
~ V3 (010) ~ V2 (110)

~ V4 (011)

~ s0

~ V (100) ~ V0 (000) 1 ~ V7 (111)

~ r
~ V5 (001)
!e

~ ~ s = V6 t

~ V6 (101)

Figure 2-9. Director torque controller: (a) flux change due to selected voltage vector (b) two-level VSC voltage vectors.

24

CH.2 Wind Turbine Concepts and Control

Figure 2-10. Voltage vectors build from VSC switching.

The voltage vectors shown in Figure 2-9 (b) are realised by switching of the two-level VSC. For example, the voltage vector are opening. is built when the switching states of the VSC is 100, is closing and the upper switches of phase and which means the upper switch of phase

~ ~ In general, it can be concluded that s and r can be controlled by ~s and ~r, respectively. v v
For example, the rotor flux of the Type C wind turbine can be controlled by rotor voltage, and the stator flux of the Type D wind turbine can be controlled by stator voltage, depending on the location of the VSC. The general structure of DTC is shown in Figure 2-11.
Te*

dT

Te

s
s*

Figure 2-11. General structure of DTC.

25

2.3 Wind Turbine Control

Voltage Vector Selector


In Figure 2-11, DTC limits the torque and flux in the specified bandwidth by selecting the optimum voltage vector. In the voltage vector selector, the entire plane is divided into six sectors. In each sector, depending on the torque error and flux error, the optimum voltage vector can be selected. The two zero-voltage vectors and do not change flux or torque, but are used to minimise the switching. The voltage vector selector is shown in Figure 2-12 and the optimum voltage vector reference table is shown in Table 2-4. In Table 2-4, ds and dTe can be defined as: 1 if s > s + bd ds = 1 if s < s bd 8 < 1 if Te > Te + Tbd 0 if Te Tbd Te Te + Tbd dTe = : 1 if Te < Te Tbd

(2.12)

Table 2-4. Optimum voltage vector selection table of the stator VSC of a Type D wind turbine.

ds
1

dTe
1 0 -1 1

Sector 1

Sector 2

Sector 3

Sector 4

Sector 5

Sector 6

-1

0 -1

~ V3 ~ V0 ~ V5 ~ V2 ~ V7 ~ V6

~ V4 ~ V7 ~ V6 ~ V3 ~ V0 ~ V1

~ V5 ~ V0 ~ V1 ~ V4 ~ V7 ~ V2

~ V6 ~ V7 ~ V2 ~ V5 ~ V0 ~ V3

~ V1 ~ V0 ~ V3 ~ V6 ~ V7 ~ V4

~ V2 ~ V7 ~ V4 ~ V1 ~ V0 ~ V5

It should be noticed that the generator operation is considered in this thesis. Thus the voltage vector selected to increase or decrease the torque is opposite to motor operation [Bose'02, Kazmierkowski, et al.'02, Vas'98, Wu'06].

DTC of Grid-Side VSC


The previous section discussed the DTC of the stator VSCs of Type D wind turbines. The grid-side VSC does not have actual flux. However, it is possible to build a virtual grid flux as (2.13): ~ dg ~ ~ = Vg + Rg Ig (2.13) dt

26

CH.2 Wind Turbine Concepts and Control

~ V2 (F I; T I)

~ V3 (F I; T D)

~ V4 (F I; T D) ~ V5 (F D; T D)

~ V1 (F D; T I)

~ V4 (F D; T D)

~ V1 (F I; T I)

s
~ V3 (F I; T I)
~ V2 (F D; T I)

~ V6 (F D; T I)

~ V3 (F D; T D)

~ V2 (F I; T D)

s s
~ V5 (F I; T D) ~ V6 (F D; T D)

!r
~ V5 (F D; T I)

~ V6 (F I; T I)

s
~ V3 (F D; T I)

s
~ V2 (F D; T D)

~ V1 (F D; T D) ~ V4 (F I; T I) ~ V6 (F I; T D)

~ V4 (F D; T I)

~ V1 (F I; T D)

~ V5 (F I; T I)

Figure 2-12. DTC voltage vector selection of the generator VSC of a Type D wind turbine. increase; : flux decrease; : torque increase; : torque decrease.

: flux

With the virtual grid flux, the control of grid VSCs is similar to that of the generator VSCs. The voltage vector selector is the same as in Figure 2-12 and Table 2-4. The two controlled variables Te and s are replaced by active power Pg and reactive power Qg , respectively. An important feature is that for DTC of grid VSCs, to increase power output, the converter flux has to be controlled to rotate counter-clockwise. Thus its voltage vector selection in Table 2-5 is the opposite of the selection in Table 2-4.

27

2.3 Wind Turbine Control

Table 2-5. Optimum voltage vector selection table of the DTC of grid VSCs of Type D wind turbines with virtual grid flux.

dQg
1

dPg
1 0 -1 1

Sector 1

Sector 2

Sector 3

Sector 4

Sector 5

Sector 6

-1

0 -1

~ V5 ~ V0 ~ V3 ~ V2 ~ V7 ~ V6

~ V6 ~ V7 ~ V4 ~ V1 ~ V0 ~ V3

~ V1 ~ V0 ~ V5 ~ V2 ~ V7 ~ V4

~ V2 ~ V7 ~ V6 ~ V3 ~ V0 ~ V5

~ V3 ~ V0 ~ V1 ~ V4 ~ V7 ~ V6

~ V4 ~ V7 ~ V2 ~ V5 ~ V0 ~ V1

DTC of Type C Wind Turbines


For Type C wind turbines (DFIG), the DTC of the rotor VSC is based on controlling the rotor flux r. The voltage vector selection is the same as that of the DTC of grid VSCs, as shown in Table 2-5. And the DTC of grid VSCs is also the same as in Table 2-5. The complete structures of DTC of Type C and Type D wind turbines are very similar to the FOC of Type C and D wind turbines shown in Appendix A, where only the FOC controllers are replaced by DTC controllers. They are omitted here.

2.3.3 Summary of FOC and DTC


FOC and DTC have their advantages and disadvantages, which are summarised in Table 2-6.
Table 2-6 Qualitative comparison of FOCs and DTC. FOC Require coordinate transformation Yes Require rotor position Yes and precisely DSP time Coupling between P and Q Parameter sensitivity Torque ripple + Harmonic + Dynamic response Current protection + + signifies better performance; - signifies worse performance DTC No Yes but approximately + + + + -

A major disadvantage of DTC is the varying switching frequencies, which are difficult to be filtered out. Some advanced controllers [Lai and Chen'01, Malinowski, et al.'04, Vas'98] are proposed to solve this problem. However, these methods require either hysteresis

28

CH.2 Wind Turbine Concepts and Control

bandwidth online variation or coordinate transformation, which require more CPU time for the DSP; therefore the advantage of DPC simplicity is lost.

2.4 Integration Problems


Integration of large-scale wind power into the power grid raises several integration problems which are not encountered in the operation of the previous power system, which is dominated by traditional thermal power plants. These new integration problems arise from the characteristics of wind power and the various technologies adopted in the wind energy industry. Most of the studies have concentrated on low-voltage ride through capability [Akhmatov'03, Holdsworth, et al.'01, Holdsworth, et al.'03, Kanellos and Hatziargyriou'02, Kehrli, et al.'03, Koch, Erlich and Shewarega'03, Koch, Erlich, Shewarega, et al.'03, Pierik, et al.'04, Srensen, et al.'03, Srensen, et al.'01, Wiik, et al.'00], while the others have focused on power quality [Larsson'00, '02, Thiringer, et al.'01]. Some studies have focused on the small-signal stability of wind energy [Mendonca and Lopes'05, Slootweg, et al.'03, Tsourakis, et al.'08, Wang, et al.'08]. In [Pourbeik, et al.'03, Rosas'03], the sub-synchronous oscillations excited by wind power variations are recommended for future research. In this section, the general integration problems and solutions will be reviewed.

2.4.1 Static Problems


The steady-state problems relate to overloading and static voltage/frequency deviations. It is caused by the variations in electric power due to the long-term fluctuations of wind speed. The possible solutions could be: Improve the wind speed forecast, or use an energy storage system [Ackermann'05, Holttinen, et al.'06]. Design coordinated central wind farm control and individual wind turbine control to make the electric power output of the wind farm smooth and schedulable. [Hansen, et al.'06, Ko, et al.'07, Zhou, et al.'08]. Limit the total active power reserve requirement for wind power by changing the power system balance plan from day-ahead to hour-ahead [Holttinen, et al.'06], since the short-term wind predictions (hour-ahead) are much more accurate than the long-term (day-ahead) predictions.

29

2.4 Integration Problems

2.4.2 Power Quality


Two important power-quality problems related to integration of wind power are flicker and harmonics. Flicker is the 1-25 Hz voltage variations. These low-speed voltage variations can be caused by short-term fluctuations of wind speed, turbulence, tower shadow, wind shear effect, and start-up and shut-down of the wind turbines. Flicker can be a very serious problem for Type A and B wind turbines, because the fluctuations of wind speed are directly reflected into the electric power output. For wind turbine Types C and D, because the rotation speed is decoupled from power grid, the variable speed shafts become buffers which can absorb the fluctuations of the wind speed. Since the wind-speed fluctuations are quite a localised phenomenon, a large geographical area will contribute to a spatial smoothing effect on the electric power output of the large wind farm. The possible solutions could be: Reinforce the transmission system, i.e. to decrease the transmission impedance. Implement an anti-flicker algorithm in the reactive power control systems of wind turbine Types C and D. Design appropriate start-up and shut-down strategies for wind turbines to avoid large fluctuations of active power and reactive power. Use dynamic reactive power compensation to control the voltage at the point of connection with the power grid.

Harmonics are components with frequencies which are multiples of the supply frequency, i.e., 100 Hz, 150 Hz, 200 Hz, etc. Inter-harmonics are similarly defined as components having frequencies located between the harmonics of the supply frequency. Harmonics can be caused by switching off either the power converters, the saturation of transformers or the unbalanced loads. The possible solutions could be: Passive filters, which can improve the power factor and reduce high-frequency harmonics. Their performance depends on tuning the reactors according the unknown source impedance. Active filters, which can inject either the anti-phase PWM currents or voltages into the grid. Use PWM control techniques to eliminate specific dominating orders of harmonics. Reactive power compensation system STATCOM can also be used to mitigate the specific orders of harmonics.

30

CH.2 Wind Turbine Concepts and Control

2.4.3 Transient Stability


One challenge of integrating a large-scale wind farm with the power grid is the required fault ride through capability (i.e., low-voltage ride through capability). This capability is normally related to the transient-voltage stability of the wind generator during voltage dips. Type A and B wind turbines are basically induction generators which require reactive power for magnetizing. The limited electric torque during voltage dips will lead to generator acceleration and an increase in the slip. From the torque-speed curve and reactive power-speed curve, it is understood that when the slip is larger than the critical value, the electric torque decreases, and the generator absorbs large amounts of reactive power. The increased reactive power consumption makes the voltage continue to drop while the generator shaft continues to accelerate. Without proper remediation, the positive feedback will ultimately lead to voltage collapse and over-speed of the generator. Type B wind turbines have a variable rotor resistance which can increase the value of the critical slip. Thus Type B wind turbines have better LVTR capability than Type A wind turbines. For Type C DFIG wind turbines, which are equipped with partially-rated power VSC converters on the rotor, if the voltage dip is not severe, the rotor converters can control the DFIG to produce reactive power for the grid. However if the voltage dip is severe, then the rotor converter may face a transient over-current and will therefore have to be blocked while the rotor is short circuited with resistances (active crowbar). Then the DFIG becomes a normal induction machine and requires substantial reactive power. The grid VSC of the DFIG is used to produce reactive power during this period. However because its power rating is limited, voltage collapse and over-speed of the generator may occur. Type D wind turbines with full-scale back-to-back VSC converters on the stator have the best low-voltage ride through (LVRT) capability. The problems are over-current of the grid VSC and possible DC over-voltage. The grid VSC can be protected by over-current protection and the DC voltage can be controlled by shunt DC resistance (i.e., a DC chopper). Some general solutions to improving the transient voltage stability of wind park could be: Use dynamic reactive power compensations such as an SVC, STATCOM or the grid-side VSCs of Type C and Type D wind turbines to help the recovery of voltage. Use an HVDC link to decouple the wind farm from the power grid.

31

2.5 Conclusion

2.4.4 Small-Signal Stability


Small-signal stability is the ability of the power system to maintain synchronism when subjected to small disturbances. The transient stability simulations in the time domain cannot cover all the possible oscillation modes of the power system, thus the small-signal analysis is provided by calculating the Eigen values of the state matrix of the power system. A possible method to improve small-signal stability is: Implement power system stabilizer (PSS) on synchronous generator, or variable speed wind turbines (Type C and D wind turbines) to increase the damping. Use SVC or STATCOM to increase the damping. Reinforce the transmission grid.

2.4.5 Sub-Synchronous Oscillation


Sub-synchronous oscillations are the 0.2 Hz to 2 Hz low-frequency mechanical oscillations of shafts which are excited by power oscillations in the grid. Without proper damping, the oscillation may continue for a long time or even increase the oscillation magnitude. The sub-synchronous oscillation increases the fatigue of the shafts of generators and has to be avoided. Only very few studies [Akhmatov'03, Pourbeik, et al.'03, Rosas'03] mention this problem. The sub-synchronous oscillations of large synchronous generators excited by nearby wind power variations are recommended for future research.

2.5 Conclusion
In this chapter, the basics of wind energy electrical system are reviewed. The existing wind turbine concepts on the market are surveyed. There are three dominant wind turbine concepts: Type A constant speed wind turbine with squirrel cage induction generator, Type C variable speed wind turbine with a doubly-fed induction generator, and Type D variable speed wind turbine with a full-scale converter and synchronous generator. The behaviour of these concepts is crucial to the power system operation and therefore must be analysed. Two types of control methodsthe field-oriented controller and the direct torque controllerare explained and qualitatively compared. The quantitative comparison between the FOC and DTC will be shown in Chapter 3. The general integration problems and solutions are also reviewed in this chapter.

32

CH.2 Wind Turbine Concepts and Control

2.6 References
[Ack05] [Akh03] T. Ackermann, Wind power in power systems, ISBN-0-470-85508-8, John Wiley & Sons, 2005. V. Akhmatov, Analysis of dynamic behavior of electric power systems with large amount of wind power, PhD thesis, Electrical Power Engineering Department, Technical University of Denmark, Copenhagen, Denmark, 2003. F. D. Bianchi, H. De Battista, and R. J. Mantz, Wind Turbine Control Systems: Principles, Modelling and Gain Scheduling Design, ISBN-13: 9781846284922, Springer Verlag, 2007. F. Blaabjerg and Z. Chen, Power Electronics for Modern Wind Turbines, Morgan & Claypool Publishers, 2006. B. K. Bose, Modern Power Electronics and AC Drives, ISBN-7-111-11296-2, Prentice Hall, 2002. T. Burton, Wind Energy Handbook, ISBN 0-471-48997-2, Wiley, 2001. P. Gardner, A. Garrad, P. Jamieson, H. Snodin, and A. Tindal, "Wind Energy the Facts - Technology," the European Wind Energy Association 2004. A. D. Hansen and L. H. Hansen, "Wind Turbine Concept Market Penetration over 10 Years (1995-2004)," Wind Energy, vol. 10, issue 1, Page(s). 8197, 2007. A. D. Hansen, P. Sorensen, F. Iov, and F. Blaabjerg, "Centralised power control of wind farm with doubly fed induction generators," Renewable Energy, vol. 31, issue 7, Page(s). 935-951, 2006. E. Hau, Wind Turbines: Fundamentals, Technologies, Application, Economics, ISBN 3-540-24240-6, Springer, 2006. R. Hoffmann, a comparison of control concepts for wind turbines in terms of energy capture, Elektrotechnik und Informationstechnik Department, Technischen University at Darmstadt, Darmstadt, Germany, 2002. L. Holdsworth, N. Jenkins, and G. Strbac, "Electrical stability of large, offshore wind farms," Seventh International Conference on AC-DC Power Transmission, 2001., 2001, Page(s). 156-161. L. Holdsworth, X. G. Wu, J. B. Ekanayake, and N. Jenkins, "Comparison of fixed speed and doubly-fed induction wind turbines during power system disturbances," in IEE Proceedings of Generation, Transmission and Distribution, 2003, Page(s). 343-352. H. Holttinen, P. Meibom, C. Ensslin, L. Hofmann, J. McCann, J. Pierik, J. O. Tande, E. Hagstrom, A. Estanqueiro, and H. Amaris, "Design and Operation of Power Systems with Large Amounts of Wind Power Production, IEA collaboration," in European Wind Energy Conference 2006, EWEC2006, Athens, Greece, 2006. F. D. Kanellos and N. D. Hatziargyriou, "The effect of variable-speed wind turbines on the operation of weak distribution networks," IEEE Transactions on Energy Conversion, vol. 17, issue 4, Page(s). 543-548, 2002. M. P. Kazmierkowski, R. Krishnan, and F. Blaabjerg, Control in Power Electronics: Selected Problems, Academic Press, 2002.

[Bia07]

[Bla06] [Bos02] [Bur01] [Gar04] [Han07]

[Han06]

[Hau06] [Hof02]

[Hol01]

[Hol03]

[Hol06]

[Kan02]

[Kaz02]

33

2.6 References

[Keh03]

[Ko07]

[Koc03]

[Koc03]

[Lai01]

[Lar00] [Lar02]

[Mal04]

[Men05]

[Nog98]

[Pie04]

[Pou03]

[Ros03]

[Shr08]

A. Kehrli, M. Ross, A. Superconductor, and W. I. Middleton, "Understanding grid integration issues at wind farms and solutions using voltage source converter FACTS technology," in IEEE Power Engineering Society General Meeting, 2003. H. S. Ko, S. Bruey, J. Jatskevich, G. Dumont, and A. Moshref, "A PI Control of DFIG-Based Wind Farm for Voltage Regulation at Remote Location," in IEEE Power Engineering Society General Meeting, 2007., 2007, Page(s). 1-6. F. W. Koch, I. Erlich, and F. Shewarega, "Dynamic Simulation of Large Wind Farms Integrated in a Multi Machine Network," in IEEE Power Engineering Society General Meeting, 2003, Page(s). 13-17. F. W. Koch, I. Erlich, F. Shewarega, and U. Bachmann, "Dynamic interaction of large offshore wind farms with the electric power system," in 2003 IEEE Power Tech Conference Proceedings, Bologna, 2003. Y. S. Lai and J. H. Chen, "A new approach to direct torque control of induction motor drives for constant inverter switching frequency and torque ripple reduction," IEEE Transactions on Energy Conversion, vol. 16, issue 3, Page(s). 220-227, 2001. A. Larsson, The Power Quality of Wind Turbines, Department, Chalmers University of Technology, Department of Electric Power Engineering, 2000. A. Larsson, "Flicker emission of wind turbines caused by switching operations," IEEE Transactions on Energy Conversion, vol. 17, issue 1, Page(s). 119-123, 2002. M. Malinowski, M. Jasinski, and M. P. Kazmierkowski, "Simple direct power control of three-phase PWM rectifier using space-vector modulation (DPC-SVM)," IEEE Transactions on Industrial Electronics, vol. 51, issue 2, Page(s). 447-454, 2004. A. Mendonca and J. A. P. Lopes, "Impact of large scale wind power integration on small signal stability," in Future Power Systems, 2005 International Conference on, 2005. T. Noguchi, H. Tomiki, S. Kondo, and I. Takahashi, "Direct power control of PWM converter without power-source voltagesensors," IEEE Transactions on Industry Applications, vol. 34, issue 3, Page(s). 473-479, 1998. J. T. G. Pierik, J. Morren, E. J. Wiggelinkhuizen, S. W. H. de Haan, T. G. van Engelen, and J. Bozelie, "Electrical and Control Aspects of Offshore Wind Farms II (Erao II) - Volume 1: Dynamic Models of Wind Farms," Energieonderzoek Centrum Nederland (ECN), Petten 2004. P. Pourbeik, R. J. Koessler, D. L. Dickmander, W. Wong, A. B. B. Inc, and N. Y. Latham, "Integration of large wind farms into utility grids (part 2performance issues)," in IEEE Power Engineering Society General Meeting, 2003. P. Rosas, Dynamic influences of wind power on the power system, PhD thesis, Electrical Power Engineering Department, Technology University of Denmark, 2003. G. Shrestha, H. Polinder, D. J. Bang, and J. A. Ferreira, "Review of Energy Conversion System for Large Wind Turbines," in European Wind Energy Conference 2008, EWEC2008, Brussels, Belgium, 2008.

34

CH.2 Wind Turbine Concepts and Control

[Slo03]

[Sor03]

[Sor01]

[Thi01]

[Tso08]

[Vas98] [Wan08]

[Wii00]

[Wu06] [Zho08]

J. G. Slootweg, H. Polinder, and W. L. Kling, "Representing wind turbine electrical generating systems in fundamental frequency simulations," IEEE Transactions on Energy Conversion, vol. 18, issue 4, Page(s). 516-524, 2003. P. Srensen, A. D. Hansen, P. Christensen, M. Mieritz, J. Bech, B. BakJensen, and H. Nielsen, "Simulation and verification of transient events in large wind power installations," Ris National Laboratory 2003. P. Srensen, A. D. Hansen, L. Janosi, J. Bech, and B. Bak-Jensen, "Simulation of interaction between wind farm and power system," Ris National Laboratory 2001. T. Thiringer, T. Petru, and C. Liljegren, "Power quality impact of a sea located hybrid wind park," IEEE Transactions on Energy Conversion, vol. 16, issue 2, Page(s). 123-127, 2001. G. Tsourakis, B. M. Nomikos, and C. D. Vournas, "Effect of wind parks with doubly fed asynchronous generators on small-signal stability," Electric Power Systems Research, 2008. P. Vas, Sensorless Vector and Direct Torque Control, ISBN-0-19-856465-1, Oxford Science Publication, 1998. C. Wang, L. Shi, L. Wang, and Y. Ni, "Modelling Analysis in Power System Small Signal Stability with Grid-connected Wind Farms of DFIG Type," Wind Engineering, vol. 32, issue 3, Page(s). 243-264, 2008. J. Wiik, J. O. Gjerde, and T. Gjengedal, "Impacts from Large Scale Integration of Wind Farms into Weak Power Systems," in International conference on Power System Technology, 2000. B. Wu, High-Power Converters and AC Drives, ISBN-13 978-0-471-73171-9, John Wiley & Sons, 2006. Y. Zhou, P. Bauer, and J. A. Ferreira, "Reactive power limitation of doubly fed induction generator and its application on coordinated hybrid wind farm control," in European Wind Energy Conference 2008, EWEC2008, Brussels, Belgium, 2008.

35

2.6 References

36

CH.3 Wind Generator System Modelling

Chapter 3
Wind Generator System Modelling 3.1 Introduction
This chapter introduces the modelling of individual wind turbines, including dynamic modelling and small-signal modelling. Dynamic model is used to study the integration problems of wind turbines, such as flickers, harmonics, transient stability, and low-voltage ride through capability. Small-signal model is used to study integration problems, such as stability of the closed-loop controller, and local and inter-area power system oscillations of the power system during large wind-power penetration. In this chapter, the dynamic simulation results of one wind turbine with different control methods and modelling approaches are compared. These methods are a field-oriented controller with an average model, a field-oriented controller with a switching model, and a direct torque controller with a switching model. The dynamic model of a Type C wind turbines with a fieldoriented controller during a grid-voltage dip is verified by experiments. The Type C smallsignal model is validated by comparing its time-step responses with the dynamic model. The stability of a single Type C wind turbine integrated into an infinite bus system is analysed, the results of which determine the stability margin of the closed-loop controlled wind turbine.

3.2 Dynamic Model


In Matlab/Simulink, dynamic models of different wind turbine concepts are derived based on the general topology of a wind turbine, as shown in Figure 3-1. In this thesis, all electrical components are modelled in a dq synchronously rotating reference frame, as discussed in the literature [Morren, et al.'05, Ong'98, Pierik, et al.'04]. The voltages vd, vq , and currents id , iq are DC values at steady state, thus in Matlab/Simulink the simulation is much faster than the classical abc models such as the power system blockset. In Figure 3-1, the direction of the arrow indicates whether the corresponding variable is an input or an output. If the arrow is pointed at a block, then the corresponding variable is an input of the block, otherwise it is an output. For example, the shaft block has two inputs: mechanical torque and electrical torque, and two outputs: turbine speed and generator speed.

37

3.2 Dynamic Model

Figure 3-1. General structure of a wind turbine dynamic model.

3.2.1Wind Speed and Rotor


The wind speed and rotor models are only required in flicker, fatigue and wind turbine control studies. For general electro-mechanical simulation purposes, the detailed wind speed and turbine rotor models can be omitted. For example, in low-voltage ride through capability studies, the detailed wind speed and rotor models can be represented simply by a constant input mechanical torque. Detailed wind speed, rotor and wind turbine aerodynamic controller models have been developed [Pierik, et al.'04, Van Engelen and Wiggelinkhuizen'03] and therefore can be connected to electrical dynamic models when necessary.

3.2.2 Shaft
For wind turbines, the generator is only softly coupled with wind turbines, i.e., the shaft has a slight stiffness which allows large speed difference between generator and turbine.
gear Gearbox Ratio

Mechanical Torque Electrical Torque

Figure 3-2. Shaft model.

Because of the large transient speed change in the generator, the elastic energy stored in the shaft will be released during grid disturbance. The released energy will have a considerable influence on the electric torque output, current and active power [Ackermann'05, Akhmatov'03, '05]. Thus two-mass shaft models are necessary for dynamic modelling if the generator stator is directly connected to the power grid:

38

CH.3 Wind Generator System Modelling

dwt 1 !gen = !wt dt gear

(3.1)

d!wt 1 1 = Taero Tshaf t dt Jwt Jwt


1 d!gen 1 = Tshaf t Te dt Jgen gear Jgen Tshaf t = Cdr !wt
where

(3.2) (3.3)

Cdr !gen + Kdr wt gear


and

(3.4) are the inertia of the wind is the stiffness;

is the twisted angle of the turbine shaft,

turbine and the generator, respectively; is the ratio of gearbox; generator.

is the damping coefficient;

is the input mechanical torque from the rotor blades; is the electrical torque from

is the torque produced from shaft rotation; and

In this thesis, all electrical components are modeled in a dq 0 synchronously rotating reference frame using Park and inverse Park transformations [Morren, et al.'05, Ong'98, Park'29, Pierik, et al.'04]. Under balanced three-phase conditions, voltages and currents in

dq 0 coordinates are DC values at steady state. For an unbalanced three-phase system,


voltages and currents in dq 0 coordinates vary, i.e., the dq -sequence values have two times the fundamental frequency components while the zero-sequence value has a fundamental frequency component. Voltages and currents of the average models are directly expressed in a dq synchronously rotating reference frame. For the switching models of power electronic converters, Park and inverse Park transformations are required to convert voltages and currents between the abc and dq 0 reference frames, because the switching functions of both the CSC and VSC are expressed in an abc reference frame (see Chapter 3). In this thesis, only the three-phase three-wire system is studied with the zero sequence omitted. The zero sequence is omitted due to the absence of a physical zero circuit for the wind generator system [Anderson and Fouad'03, Kundur'94]: The generator transformer is often Y/-connected; The neutral point of the stator winding of the generator is not grounded.

39

3.2 Dynamic Model

3.2.3 Generator 3.2.3.1 Induction Generator

Figure 3-3. Circuit diagram of a wound rotor induction generator (a) d-axis (b) q-axis.

The circuit diagram of a WRIG for Type C wind turbines is presented as Figure 3-3; its dynamic equations are listed in a dq synchronously rotating reference frame as equations (3.5) and (3.6). The rotor voltages vdr and v (vdr ; vqr = 0).
q r

are produced by a VSC. The squirrel cage

induction generator model is omitted here, because its rotor simply short-circuited circuits

vds = Rs ids !s qs + dds dt dqs vqs = Rs iqs + !s ds + dt vdr = Rr idr (!s p!m )qr + vqr = Rr iqr + (!s p!m )dr +
The flux equations are:

ddr dt dqr dt

(3.5)

ds qs dr qr

= (Ls + Lm )ids Lm idr = (Ls + Lm )iqs Lm iqr = (Lr + Lm )idr Lm ids = (Lr + Lm )iqr Lm iqs

(3.6)

40

CH.3 Wind Generator System Modelling

and the torque equation can be expressed by multiple equations. The torque equation used in simulation on the actual controlled variables. For example, in type C wind turbine, the torque equation is expressed by (3.7), because the rotor currents variables in type C wind turbine. and are controlled

Te = p(dr iqr qr idr )

(3.7)

3.2.3.2 Synchronous Generator


Two types of synchronous generators are used in wind turbine concepts. One type uses field winding and DC excitation: the circuit diagram and equations can be found in [Anderson and Fouad'03, Kundur'94] but have been omitted here. The other type uses a permanent magnet instead of field winding, which is called a PMG. The PMG model is similar to the classical synchronous generator, only the damping windings are missing and the field winding is replaced by a permanent magnet. Its circuit diagram is shown as Figure 3-4, and its dynamic equations are listed in a dq synchronously rotating reference frame as (3.8) and (3.9).

Figure 3-4. Circuit diagram for a permanent magnet generator (a) d-axis (b) q-axis.

41

3.2 Dynamic Model

The stator voltage equations are:

vds = Rs ids !s qs + vqs = Rs iqs + !s ds +


and the flux equations are:

dds dt dqs dt

(3.8)

ds = (Ls + Lm )ids f qs = (Ls + Lm )iqs


The torque equation is:

(3.9)

where f is the excitation flux of the permanent magnets linked to the stator windings.

Te = piqs [ids (Lds Lqs ) + f ]

(3.10)

3.2.4 Power Electronic Converter


The power electronic converter can either be modelled as a controlled voltage source which is an average model, or as a switching model which simulates the actual switching of the components. The average model is fast and accurate enough for electro-mechanical simulation [Akhmatov'03, Morren, et al.'05]. However, in some situations, it is necessary to use the switching model, for example, for the DTC controller of the VSC, the harmonics study, etc. In this thesis, both the average model and switching model of a VSC and phasecontrolled CSC are developed in Simulink. The model of the switching CSC based on IGBT is similar to the VSC and is omitted in the thesis. The topologies of a VSC and a phase-controlled CSC are shown in Figure 3-5.
Idc Idc

Vdc

Vdc

Figure 3-5. (a) Voltage-source converter (VSC) using IGBT (b) phase-controlled current source converter (CSC) using thyristor.

42

CH.3 Wind Generator System Modelling

3.2.4.1 Average Model


In average models, power electronic converters are modelled either as a controlled currentsource or a controlled voltage-source. The switching of semiconductors is omitted in the average model. The phase controlled CSC is modelled as a controlled DC voltage source and AC current source. The VSC is modelled as a DC current source and controlled AC voltage source. The AC current and voltage are modelling directly in a dq reference frame. Park and inverse Park transformations are not required.

Current Source Converter


Balanced Grid-Voltage Condition CSCs using IGBTs are similar to VSCs using IGBTs, and can be modelled similarly as the VSCs using IGBTs, but only with different modulation strategies. Only a CSC using a thyristor bridge is modelled in this thesis. The diode bridges average model is the same as that of the thyristor bridge, only where . The average CSC model is a controlled DC voltage source and an AC current source, as represented by equations (3.11) to (3.15). Under symmetrical grid voltage, the DC voltage of CSC is calculated as: p q 2 3 2 Vd + Vq2 cos 3!s Lc Idc Vdc = (3.11) where Vd and Vq are the instantaneous AC voltages in a dq frame, is the rectifier angle, !s is the grid frequency, Lc is the commutation AC inductance, and Idc is the DC current. Because of the AC inductance Lc , there is a commutation angle: p 2!s Lc Idc = q 2 + V 2 sin Vd q

(3.12)

The root mean square (RMS) AC current is the fundamental frequency component calculated from the Fourier analysis: p 6Idc Irms = k (3.13) where k is the coefficient: p [cos 2 cos(2 + 2)]2 + [2 + sin 2 sin(2 + 2)]2 k= 1 4[cos cos( + )]

(3.14)

43

3.2 Dynamic Model

Thus the instantaneous AC current can be calculated as: p 3 2Idc sin vi id = p 3 2Idc iq = cos vi and current. Unbalanced Grid-Voltage Condition

(3.15)

where vi = + 0:5 is the power factor angle, i.e. the angle between AC phase voltage

Under unbalanced grid-voltage conditions, vd and vq are not constant DC values even at steady state and the RMS values of the phase voltages cannot be calculated from vd and vq . A Fourier analysis has to be used to calculate the RMS values and phase angles of the three-phase AC voltages. The average DC voltage is much more complex compared to the balanced grid-voltage conditions of equation (3.11) [Arrillaga and Arnold'90]. p 2 Vdc = Vab [cos(v ab + v ac + ) cos(v bc + v ab + )] p 2 + Vac [cos v ab + v ab ) cos v ac + v ab )] (3.16) p 2 + Vbc [cos(v ac + v bc ) cos(v ab + v bc + )] Idc !s (Lc1 + Lc2 + Lc3 )

Vab , Vbc , Vac are the RMS values of the line-to-line voltages, v

ab ,

bc ,

ac

are the

corresponding phase angles of the line-to-line voltages, all of which are calculated from a Fourier analysis of the instantaneous AC voltages; Idc is the DC current; Lc1, Lc2, Lc3 are the commutation phase inductances on the AC side; and is the fire angle. The conduction periods of each phase are:

Ta = v Tb = v Tc = v

v ac bc v ba + ac v bc +
ba

(3.17)

The RMS values of AC phase currents are calculated with conduction periods (3.17).

44

CH.3 Wind Generator System Modelling

p 2 2Idc Ta sin( ) Ia = 2 p 2 2Idc Tb Ib = sin( ) 2 p 2 2Idc Tc Ic = sin( ) 2


Commutation angles are calculated as:

(3.18)

a = arccos(cos

Idc !s (Lc1 + Lc3 ) p ) 2Vac Idc !s (Lc1 + Lc2 ) p ) b = arccos(cos 2Vba Idc !s (Lc2 + Lc3 ) p c = arccos(cos ) 2Vbc

(3.19)

Phase angles of phase currents are calculated as:

i a = v i b i c

a 2 b = v b 2 c = v c 2
a

(3.20)

Instantaneous AC currents in an abc reference frame can be calculated from (3.17) to (3.20). p ia (t) = 2Ia sin(!s t + i a ) p (3.21) ib (t) = 2Ib sin(!s t + i b ) p ic (t) = 2Ic sin(!s t + i c ) Finally, Park and inverse Park transformations are used to convert voltages and currents between abc and dq reference frames.

Voltage Source Converter


The average VSC model is the same for both balanced and unbalanced grid-voltage conditions. It is modelled as a DC current source and a controlled AC voltage source in equations from (3.22) to (3.24), which display the reciprocals of the CSC model. The AC voltage is simply the reference voltages which are the outputs of the current controller:
vd = vd

(3.22) (3.23)

vq = vq

45

3.2 Dynamic Model

and the DC current is:

Idc =

vd id + vq iq Vdc

(3.24)

where Vdc is the DC voltage, and vd ; vq and id ; iq are the instantaneous AC voltage and current.

3.2.4.2 Switching Model


For simulation purposes, such as the DTC controller and the harmonic study, the switching model of a power electronic converter is required. The switching model is implemented by the switching function, which is derived in an abc reference frame. Then Park and inverse Park transformations are used to convert voltages and currents between abc and dq reference frames.

Current Source Converter


Without Commutation Effects The commutation effect of AC inductance causes a drop in the DC side voltage at normal operating points and is omitted at first to simplify the model and improve the simulation speed. Without the commutation effects, the logic of a thyristor bridge is that the thyristor stays blocked until it receives a firing pulse, after which it compares the three-phase voltages to determine whether the conducting thyristor is in the most positive phase (1, 3, 5) or in the most negative phase (4, 6, 2). The phase current flows when the thyristor of that phase is conducting current. In Simulink, the above discussed switching logic is implemented by comparing the instantaneous three-phase AC voltages Va ; Vb ; Vc to each R other and then comparing their instantaneous electric angle !s dt versus fire angle . The

states of the three-phase thyristor bridge (as provided in Table 3-1) are influenced by two aspects: Comparison between the three-phase instantaneous AC voltages; R Comparison between the instantaneous electric angle !s dt and fire angle .
Table 3-1. CSC states. Conducting T1 = 1; T3 = 1; T5 = 1 Blocked T1 = 0; T3 = 0; T5 = 0

Components 1, 3, 5 4, 6, 2

T4 = 1; T6 = 1; T2 = 1

T4 = 0; T6 = 0; T2 = 0

46

CH.3 Wind Generator System Modelling

Table 3-2. Conducting sequence of the three-phase thyristor bridge in one cycle. Conducting Components 1, 2 3, 2 3, 4 5, 4 5, 6 1, 6 Positive DC voltage va Negative DC voltage vc Total DC voltage va vc

vb vb vc vc va

vc va va vb vb

vb vc

vb va vc vb

vc va va vb

The DC voltage of the three-phase thyristor bridge in one cycle can be calculated as Table 3-2. The DC voltage can be calculated from AC voltage as:

Vdc = (T1 + T4 )va + (T3 + T6 )vb + (T5 + T2 )vc


The three-phase AC current can be calculated from DC current as:

(3.25)

ia = (T1 + T4 )Idc ib = (T3 + T6 )Idc ic = (T5 + T2 )Idc


between abc and dq reference frames. With Commutation Effect

(3.26)

Park and inverse Park transformation are used to convert the AC voltages and currents

Omitting the commutation effect can cause large errors in the DC voltage when the AC commutation inductance that is connected to the thyristor bridge is large. Because of this, an instantaneous thyristor bridge model with commutation effect is developed [Zhou, et al.'06]. Compared to the instantaneous model without commutation effect, the main differences of the model with commutation effect are: The current is a trapezoidal waveform, the increasing and decreasing slopes are determined by the AC line-to-line voltage and AC commutation inductance; Besides the influence of the AC phase voltage and fire angle, the switch-off logic of the thyristor is also influenced by the instantaneous current flow through the thyristor; the thyristor is only blocked when its current is less than 0. The switchoff logic of thyristor is implemented by a zero-hit crossing block in Matlab/Simulink. The switching function, which takes commutation effect into consideration, is complex and is omitted here for simplicity. The detailed instantaneous model of a phase-controlled CSC with and with commutation effects can be seen in [Zhou, et al.'06]. 47

3.2 Dynamic Model

Voltage Source Converter


The switching model of a VSC is implemented with its modulation strategy. The states of VSCs are shown in Table 3-3, which are different from the states of CSCs in Table 3-1.
Table 3-3. VSC states. Conducting Components 1, 3, 5 4, 6, 2 States

Sa = 1; Sb = 1; Sc = 1 Sa = 0; Sb = 0; Sc = 0

The states of VSCs are calculated in the Simulink model from the modulation strategy. The modulation strategy uses a reference voltage or reference current as the inputs, and also generates the gate signals, which are the states in simulation. There are many modulation methods, such as: Hysteresis band current control PWM Sinusoidal carrier waveform-based PWM (SPWM) Space-vector PWM (SVPWM)

Hysteresis band current control PWM is simple, because if used, then the PI controller is not required. Removing the PI controller makes the current controller faster and insensitive to parameter variation. However, compared to SPWM and SVPWM, the large disadvantage is: the switching frequencies are not constant, thus the harmonics are difficult to be filtered out. SVPWM can be seen as a special form of SPWM [Kazmierkowski, et al.'02, Zhou and Wang'02], where a special zero-sequence waveform is superimposed onto the carrier waveform. With the development of microprocessors, space-vector modulation has become one of the most important PWM methods for three-phase converters. The space-vector concept is used to compute the duty cycle of the switches, which is simply the digital implementation of PWM modulators. The notable features of space vector modulation are: easy digital implementation capability and a wide linear modulation range for the output line-to-line voltages. In this thesis, only SVPWM is considered and simulated. SVPWM is illustrated as Figure 3-6.

48

CH.3 Wind Generator System Modelling

~ V3 (010)
~ V2 T2

~ V2 (110)
!s
V Ts

Ts

~ V4 (011)

V1 T1 ~ V0 (000) ~7 (111) V

~ V1 (100)

~ V5 (001)

~ V6 (101)

T0 2

T1 2

T2 2

T7

T2 2

T1 2

T0 2

Figure 3-6. SVPWM concepts (a) space vector diagram (b) duty ratio in time domain.

The whole space plane, which corresponds to a one cycle in time domain, is divided into 6 sectors. In each sector, the required voltage vector can be built using two adjacent voltage vectors and two zero vectors. The detailed algorithm for each sector is listed in Table 3-4.
Table 3-4. Space vector modulation algorithm.

Sector 1

Sector 2
3)

Sector 3
2 3 )

(0 <

( 3

p 3 T1 = m Ts sin( ) 2 3 p 3 T2 = m Ts sin 2 Ts T1 T2 T0 = T7 = 2
4 3 )

p 3 T2 = m Ts sin( ) 2 3 p 3 2 T3 = m Ts sin( ) 2 3 Ts T2 T3 T0 = T7 = 2

<

( 2 < ) 3
T3 =

p 3 m Ts sin( ) 2 p 3 2 T4 = m Ts sin( ) 2 3 Ts T3 T4 T0 = T7 = 2

Sector 4

Sector 5

Sector 6
5 3 )

( <

( 4 3

p 3 T4 = m Ts sin( ) 2 p 3 4 T5 = m Ts sin( ) 2 3 Ts T4 T5 T0 = T7 = 2

p 3 5 T5 = m Ts sin( ) 2 3 p 3 4 T6 = m Ts sin( ) 2 3 Ts T5 T6 T0 = T7 = 2

<

( 5 < 2) 3

p 3 5 T6 = ) m Ts sin( 2 3 p 3 T1 = m Ts sin(2 ) 2 Ts T6 T1 T0 = T7 = 2

49

3.2 Dynamic Model

~ ~ ~ For example, in sector 1 the required voltage can be built from voltage vectors V1, V2, V0 ~ and V7 , as in Figure 3-6 (a). The time durations T1 , T2 , T0 and T7 , which produce the
specific voltage vector V \, can be calculated from sector 1 of Table 3-4. p 3 T1 = m Ts sin( ) 2 3 p 3 T2 = m Ts sin 2 Ts T1 T2 T0 = T7 = 2 where m =
V Vdc .

(3.27)

With the time durations T1, T2 , T0 and T7 , which are calculated from (3.27), the states

Sa ; Sb ; Sc can be determined from Figure 3-6 (b). Finally, the AC phase voltages of VSCs
for balanced conditions can be calculated from the switching function:

va =

Vdc (2Sa Sb Sc ) 3 Vdc (2Sb Sa Sc ) vb = 3 Vdc (2Sc Sa Sb ) vc = 3

(3.28)

The DC current can be calculated as:

Idc = Sa ia + Sb ib + Sc ic
between abc and dq reference frames.

(3.29)

Park and inverse Park transformation are used to convert the AC voltages and currents

3.2.5 DC link 3.2.5.1 Current Source Converter


In phase-controlled CSCs, the DC link shown as in Figure 3-7is a series connected resistor and inductor circuit. The DC link of a CSC is modelled as (3.30).

50

CH.3 Wind Generator System Modelling

Figure 3-7. DC link of a CSC.

dIdc Vdc rec Vdc inv Rdc Idc = dt Ldc

(3.30)

The DC links of CSCs have initialisation problems in Simulink. During initialisation, the AC voltage of the rectifier and inverter changes rapidly. The AC voltage fluctuation will easily saturate the DC current controller. Either an anti-windup PI controller or switches are used to help the CSC model ride through the initialisation period. A second problem is that even if Vdc rec < Vdc inv , the DC current Idc cannot change its direction unless the rectifier and inverter change their operation modes. This phenomenon must be modelled correctly.

3.2.5.2 Voltage Source Converter


In voltage source converters, the DC link is a DC capacitance. It is modelled as (3.31); a PI controller is used to control the DC voltage.

Figure 3-8. DC link of a VSC.

51

3.2 Dynamic Model

Idc dVdc = dt

gen

Idc grid Cdc

(3.31)

As with the DC links of CSCs, the initialisation of VSC model in Simulink can cause problems. The amount of energy stored on the DC capacitor is quite small. Large oscillations in Pgen or Pgrid during initialisation will cause significant DC voltage fluctuation, which easily causes the DC voltage controller to saturate. The solution is to use either an anti-windup PI controller or switch to help the VSC model ride through the initialisation period in Simulink.

3.2.6 Phase Locked Loop


A phase-locked loop (PLL) is used to measure the frequency and phase angle of the gridvoltage vector. PLLs can also be used to measure the rotation speed and position of the flux vector of a generator. The PLL is basically a PI controller [Harnefors and Nee'00, Kaura, et al.'97]. It is illustrated in Figure 3-9, which presents a generic structure of a PLL containing only the essential parts. Blocks such as a low-pass filter and frequency averaging are omitted.
qs
qP LL

Vqs
VdP LL

dP LL

s
ds

!P LL
!s

P LL

!0

P LL

0
Va
Vb Vc

R
Vd0
Vq0

Vd0 cos() + Vq0 sin()

R
VdP LL
!
!P LL

Vdref = 0

Figure 3-9. (a) Space vector diagram of a PLL (b) Generic structure of a PLL.

52

CH.3 Wind Generator System Modelling

The PLL forces the voltage vector to align with the quadrature

axis, which means

vdP LL = 0. If the rotating speed measured !P LL is higher than the actual rotating speed !s,
then the direct axis voltage calculated in a PLL reference frame is VdP LL > 0. As can be seen in Figure 3-9 (b), the error between actual VdP LL and Vdref will force the PI controller to decrease the rotating speed by ! . The PI controller will continue to work until

VdP LL = 0, when the voltage vector is aligned with the quadrature


actual speed !P LL and position angle P LL are measured.

axis. At that point, the

In Figure 3-9 (b), !0 is the estimated rotating speed, which works as a feed-forward term for the PI controller of PLL. With this feed-forward term, it is possible to use the small control parameters kp ; ki to increase PLL stability. From Figure 3-9 (b), the PLL model is derived by omitting the filters.

!P LL =

dP LL = !0 + kp (0 vd ) + ki dt

(0 vd )dt

(3.32)

3.2.7 Grid
Depending on the simulation purpose, the grid is modelled either as a constant voltage source or as a large synchronous generator with a speed and AC voltage controller.

3.2.8 Interface between Grid and Unit


It should be emphasised that there are multiple rotating reference frames in one simulation. A common reference frame is marked as DQ, which always rotates at the synchronously rotating reference speed !s. All the passive grid components, i.e., cable, transformer, etc., are modelled in the common reference frame. However, the synchronous generators have their own dq reference frames which rotate at their own rotating speeds !r. Also the VSWT and VSC have their own dq reference frames when they are controlled by stator-flux oriented or grid-flux oriented methods. Comparisons of different reference frames are shown in Figure 3-10. An interface between unit and grid is required to switch between the different reference frames DQ and dq , as in (3.33) and (3.34). xd cos sin x = D xq sin cos xQ

(3.33)

xD cos = xQ sin

sin x d cos xq

(3.34)

53

3.2 Dynamic Model

Where is the angle difference between the common reference frame DQ and the unit reference frame dq ; xD, xQ are variables in the DQ frame, and xd , xq are variables in the

dq frame.

Q
q

!r Units reference frame


!s Common reference frame

Stationary reference frame


Figure 3-10. Common reference frame DQ, unit reference frame dq , and stationary reference frame .

3.2.9 Model Implementation


By connecting the blocks from 3.2.1 to 3.2.8, the dynamic models of wind turbines can be implemented in Matlab/Simulink.

3.2.10 Comparing FOCs and DPCs


DPCs, FOCs with SVPWM switching models, and FOCs with the average model are compared using dynamic simulations of a grid-connected VSC. In simulation, the discrete time Ts is set to 5e-5s for both DPCs and FOCs with a switching model, which is the sampling time of a DSP in a real digital controller. Some of the important settings of the simulation are listed in Table 3-5.
Table 3-5. Simulation settings.

DPC Discrete time Ts Switching Frequency Simulation Time Step reference set reference set 5e-5 second 20 kHz

FOC-SVPWM 5e-5 second 20 kHz

FOC-Avg none none 0.01 second

Ts =10

Ts =10

Step up from 1MW to 2MW at 1.5 second Step up from 0 MVar to 0.5MVar at 1.0 seconds

54

CH.3 Wind Generator System Modelling

Simulated converter currents are shown in Figure 3-11. With the same sampling time, the DPC controller has many more ripples than FOC-SVPWM. The behaviour of FOCSVPWM with a switching model is similar to an FOC with an average model. The large transients in FOCs are caused by the interaction between the PI controller and the low-pass filter. The large transients show the difficulty of tuning the PI controller in practice. Moreover, there is also a coupling effect between the active and reactive current components in the FOC. The zoomed-in view of the direct axis current is shown in Figure 3-12. The DPC has a low, varying switching frequency (around 5 kHz), while the FOC-SVPWM has a high, constant switching frequency (20kHz), even though they have the same sampling time. The hysteresis bandwidth controller of the DPC causes the varying switching frequency. This problem can be seen more clearly in Figure 3-13 with the spectrum analysis of the converter voltage.
DPC 0.5
id (kA) iq (kA)

DPC 1.5 1 0.5 0 0.8 1 1.2 1.4 1.6

-0.5 0.8

1.2

1.4

1.6

time (s)

time (s)

FOC-SVPWM 0.5
id (kA) iq (kA)

FOC-SVPWM 1.5 1 0.5 0 0.8 1 1.2 1.4 1.6

-0.5 0.8

1.2

1.4

1.6

time (s)

time (s)

FOC-AVG 0.5
id (kA) iq (kA)

FOC-AVG 1.5 1 0.5 0 0.8 1 1.2 1.4 1.6

-0.5 0.8

1.2

1.4

1.6

time (s)

time (s)

Figure 3-11. Direct and quadrature axis currents of DPC, FOC-SVPWM, and FOC-Avg.

55

3.2 Dynamic Model

Zoomed in View of id 0.5 0.4 id (kA) 0.3 0.2 0.1 0 1.14 DPC FOC-SVPWM FOC-Avg

1.142 1.143 time (s) Figure 3-12. Zoomed-in view of direct axis current.

1.141

1.144

1.145

Nomalized Spectrum of Phase Voltage with DPC Nomalized Spectrum of Phase Voltage with FOC-SVPWM 1 1 0.8 0.6 |Y(f)| 0.4 0.2 0 |Y(f)| 0.4 0.2 0 0 2 4 6 Frequency (Hz) 8 10 x 10
4

0.8 0.6

4 6 Frequency (Hz)

10 x 10
4

Figure 3-13. Spectrum of converter voltage of DPC and FOC-SVPWM.

From Figure 3-13 it is clear that the DPC has low, varying switching frequencies distributed between 1 to 20 kHz which are difficult to filter out, while the FOC-SVPWM has almost a constant switching frequency around 20, 40 kHz, etc. Some advanced controllers of DPC have been proposed to solve this problem [Aurtenechea, et al.'06, Malinowski, et al.'04]. However, the advantage of DPC simplicity is then lost. For electromechanical simulation purposes, such as LVRT capability studies, the simulation results from an average model are similar to those of a switching model, the simulation speed of which is more than 100 times faster than the switching model. Thus, in all the later chapters, FOC with an average model will be used in all the analysis, except when special notification is made.

56

CH.3 Wind Generator System Modelling

3.2.11 Verification of Dynamic Model of Type C Wind Turbines


The dynamic simulation of Type C (DFIG) wind turbines under both symmetrical and unsymmetrical grid fault is compared with experimental results provided by Chalmers University [Petersson, Thiringer, et al.'05, Zhou, et al.'09], as shown in Figure 3-14. The generator parameters are listed in Table 3-6.
Table 3-6. Parameters of DFIG.

Type 2.75MW 850kW 850kW

Rs 1.2e-3 6.9e-3 2.62e-2

Lls 1.18e-4 9.7e-3 2.04e-2

Rr 5.2e-3 6.49e-4 3.86e-4

Llr 3.08e-4 6.49e-4 6.30e-4

Lms 4.6e-3 1.51e-2 3.52e-2

Lmr 1.82e-2 1.51e-2 3.52e-2

Source ECN Chalmers Gamesa

DFIG model, measured instant voltage dip input 0.18 0.16 Pmeas, Psimu (pu) 0.14 0.12 0.1 0.08 0.06 0.04 1.9 1.95 2 2.05 2.1 2.15 2.2 2.25 2.3 2.35

0.06 0.04 Qmeas, Qsimu (pu) 0.02 0 -0.02 -0.04 -0.06 -0.08 1.9 1.95 2 2.05 2.1 2.15 2.2 2.25 2.3 2.35

Figure 3-14. Measured and simulated p and q of DFIG during symmetrical and unsymmetrical voltage dips, red curves = measurement, blue curves = simulation using Chalmers data, green curves = simulation using Gamesa data.

In Figure 3-14, active power oscillations from measurements and from simulation agree well during both symmetrical and unsymmetrical voltage dips. However, the active power oscillations after voltage recovery and the reactive power oscillations still differ. The

57

3.3 Small-Signal Model

deviances between simulation and measurement of reactive power can be caused by the controller strategies choice and parameters, none of which were not further investigated for this thesis.

3.3 Small-Signal Model


By linearising the dynamic model around an operating point, a small-signal model can be derived. The small-signal model is important for control design and stability analysis of a complex system, which are often time-consuming and sometimes even impossible with dynamic simulation. However, the small-signal model can find the origin of the instability of the system; for example, the application of a power system stabiliser on synchronous generator was initialised from analysis of a small-signal model of the power system [Kundur'94]. There are two methods that derive the small-signal model of a dynamic system. One is a numerical method which perturbs the dynamic simulation at a specific operating point. Another one is an analytic method which expands the dynamic model using Taylor expansion, neglecting the 2 high-order terms and eliminating the steady-state terms [Bose'02, Krause, et al.'86, Kundur'94, Rogers'00]. The numerical method is simple to perform. However, dynamic simulation of a large system is difficult and time-consuming. And sometimes it is impossible to reach stable dynamic simulation results, for example, when tuning the control parameters. The analytic method is superior to the numerical method for the analysis of a complex system. Although it is initially time-consuming to build the linearised model of a generator, converter and controller, it is easy to expand the system to include more units later on. Moreover, the stability margin of the system can clearly be seen by drawing the eigen value locus of the system, when tuning the control parameters. The general small-signal model of a system is:

_ x = Ax + Bu y = Cx + Du

(3.35)

where x; u; y are the changes in state variables, inputs and outputs, respectively. In general, state variables are the outputs of integrators, such as current flow through inductors, voltages across the capacitors, etc. The inputs are often reference values for the controllers, such as the desired DC voltage, desired grid frequency, etc. By linearising the dynamic models derived in 3.2, the small-signal models are analytically developed in the following sections. As with the dynamic models, the small-signal models are also considered in a rotating reference frame. For VSWTs and VSCs, only FOCs using

58

CH.3 Wind Generator System Modelling

PI controllers are modelled. Each PI controller has one integrator, thus to each PI controller, one state variable x is added, which is the change in control error. linearised equation of a PI controller is: The general

_ x = u u y = kp (u u) + ki x

(3.36) (3.37)

where u is the reference value, u is the actual value, and kp and ki are the proportional and integral control parameters, respectively. In the next sections, the small-signal models of generators, converters and controllers are derived.

3.3.1 Changes in Flux, Torque and Power


Changes in flux, electric torque, active power and reactive power are expressed as:

ds = (Ls + Lm )ids Lm idr qs = (Ls + Lm )iqs Lm iqr dr = (Lr + Lm )idr Lm ids qr = (Lr + Lm )iqr Lm iqs P = id0 vd + iq0 vq + vd0 id + vq0 iq Q = id0 vq iq0 vd + vq0 id vd0 iq Te = pLm (iqr0 ids idr0 iqs iqs0 idr + ids0 iqr )

(3.38) (3.39) (3.40) (3.41) (3.42) (3.43) (3.44)

3.3.2Aerodynamic Torque
The linearisation of the aerodynamic torque equation gives:

Ta = (R3 Cq0 Vw0 0:5kopt R4 !wt0 )Vw + 0:5kopt R4 Vw0 !wt = aVw Vw + a!wt !wt

(3.45)

3.3.3 Shaft
The small-signal model of a two-mass shaft is:

_ wt = !wt

1 gear

!m

(3.46)

59

3.3 Small-Signal Model

a! Kdr Kdr Cdr aV _ !wt = wt + !m wt + w Vw Jwt Jwt gear Jwt Jwt


_ !m = Kdr Kdr Cdr 1 !wt !m + wt Te 2 Jgen gear Jgen gear Jgen gear Jgen

(3.47)

(3.48)

where !wt and !m are the changes in rotation speed of a low-speed shaft and highspeed shaft, respectively. wt is the change in the shaft twisting angle, which is referred to the low-speed shaft. model in section 3.2. is the change of wind speed. is the change of wind speed. The other parameters such as are the same as in equations of dynamic

3.3.4 Wound Rotor Induction Generator


The linearised stator winding equations of an induction generator are:

_ ids =

1 (Rs Lr ids + [!e0 Ls Lr + (!e0 p!m0 )L2 ]iqs !e0 Lr Lm iqr m L2 Ls Lr m Rr Lm idr p(Lr iqr0 + Lm iqs0 )Lm !m + Lr Vds Lm Vdr ) (3.49) L2 m 1 ([!e0 Ls Lr (!e0 p!m0 )L2 ]ids + Rs Lr iqs + !e0 Lr Lm idr m Ls Lr Rr Lm iqr + p(Lr idr0 + Lm ids0 )Lm !m + Lr Vqs Lm Vqr ) (3.50)

_ iqs =

The rotor winding equations are:

_ idr =

1 (Rs Lm ids p!m0 Ls Lm iqs + [!e0 L2 + (!e0 p!m0 )Ls Lr ]iqr m Ls Lr L2 m Rr Ls idr p(Lr iqr0 + Lm iqs0 )Ls !m + Lm Vds Ls Vdr ) (3.51) 1 (p!m0 Ls Lm ids + Rs Lm iqs + [!e0 L2 (!e0 p!m0 )Ls Lr ]idr m Ls Lr L2 m Rr Ls iqr p(Lr iqr0 + Lm iqs0 )Ls !m + Lm Vqs Ls Vqr ) (3.52)

_ iqr =

where Vdr and Vqr are the changes in the output voltages of rotor VSC, which are

calculated from the current controllers in (3.62) and (3.63). ids and iqs are the changes of the stator currents. idr and iqr are the changes of the rotor currents. ids0 and iqs0 are the stator currents at the linearized operation point. idr0 and iqr0 are the rotor currents at the linearized operation point.

60

CH.3 Wind Generator System Modelling

3.3.5 Converter
It is not necessary to model the switching of a converter in the small-signal model. The converter is modelled as an average model, which is a controlled voltage source and produces the outputs of the field-oriented current controllers.

3.3.6 Generator VSC Controller 3.3.6.1 Reference Torque and Power Generation
The reference torque of VSWTs is generated from the speed controller and grid frequency droop controller:
Te =

2kopt !wt0 !wt + Dp (!P LL ! ) gear

(3.53)

where kopt =

0:5R5 Cq0 , 2 0

which is determined by the Cq curve at 0 ; !wt is the

change in rotation speed of the low-speed shaft; Dp is frequency droop ratio; !P LL and

! are the changes in the measured grid frequency and desired grid frequency,
respectively.
The reactive power reference Q is similar to Te : s

jVs j =

vds0 vds + vqs0 vqs q 2 2 vds0 + vqs0

(3.54)

Q = Q0 + Dq (jVs j Vs ) s s

(3.55)

where Dq is the voltage droop ratio; Qs0 is the change in reactive power order; jVs j and

Vs are the changes in the measured grid voltage and desired grid voltage, respectively.

3.3.6.2 Torque/Flux Controller


The torque/flux controller of FOCs controls the torque and flux errors and generates reference currents. The linearised equations of Type C (DFIG) wind turbines are as follows. The changes in torque and reactive power errors are:
_ Te = Te Te

(3.56) (3.57)

_ Qs = Q Qs s

61

3.3 Small-Signal Model

where Te and Q are the changes in the reference values, which are calculated from s 2kopt !wt0 Te = !wt + Dp (!P LL ! ) gear (3.53) and (3.55). Te and Qs are the

changes in the actual values, which are calculated from(3.44) and (3.43). The outputs of the torque/flux controller are the reference rotor currents, which are:

i = kp Qs (Q Qs ) + ki Qs Qs dr s
i = kp T e (Te Te ) + ki T e Te qr

(3.58) (3.59)

The torque and flux controllers of Type D wind turbines can be derived similarly and are therefore omitted here. The difference is that the direct current idr is used to control the flux instead of the stator reactive power.

3.3.6.3 Current Controller with Decoupling


For FOC of VSCs, current controllers are required to produce the desired voltage. The linearised equations of DFIG rotor current controllers are as follows. The changes in the direct and quadrature current errors are:

_ idr = i idr dr

(3.60) (3.61)

_ iqr = i iqr qr
(3.58) and (3.59). The outputs of the current controllers are the desired rotor voltages, which are:

where i and i are changes in the reference currents, which are calculated from qr dr

Vdr = kp (i idr ) + ki idr (!P LL0 p!m0 )qr qr 0 !P LL + pqr0 !m dr (3.62) Vqr = kp (i iqr ) + ki iqr + (!P LL0 p!m0 )dr + dr0 !P LL pdr0 !m qr (3.63)

where !m is the change in the mechanical rotation speed of the high-speed shaft. p is the number of pole pairs. dr and qr are the changes of the rotor fluxes. is the change of measured grid frequency from PLL. The final 3 terms in (3.62) and (3.63) are the decoupling terms in the current controller.

62

CH.3 Wind Generator System Modelling

3.3.7 Grid VSC Controller 3.3.7.1 Active and Reactive Power Controllers
The active and reactive power controllers generate the reference currents. The active power
reference of grid VSCs is Pgrid , which is calculated from the DC link voltage controller

(3.74). The reactive power reference is the same as the stator reactive power reference in (3.55). PI controllers can be used to generate reference currents same as the torque/flux controller of machine VSCs. However, since the grid voltages vds vqs are often constant, the reference currents can simply be calculated using mathematical equations. Furthermore, if the gridvoltage vector is aligned with the quadrature axis, then vds = 0. The linearised reference currents of grid VSCs are:

i = dg

Q Qgrid0 grid Vqs 2 Vqs0 Vqs0


Pgrid Pgrid0 Vqs 2 Vqs0 Vqs0

(3.64)

i = qg

(3.65)

3.3.7.2 Current Controller with Decoupling


Current controllers of grid VSCs are similar to rotor VSCs. The changes in direct and quadrature current errors are:

_ idg = i idg dg _ iqg = i iqg qg


from (3.64) and (3.65). Outputs of the current controllers are the desired grid VSC voltages, which are:

(3.66) (3.67)

where i and i are changes in reference of grid VSC currents, which are calculated qg dg

Vd vsc = Vdg + kp (i idg ) + ki idg Lg iqg0 !P LL !P LL0 Lg iqg (3.68) dg

Vq vsc = Vqg + kp (i iqg ) + ki iqg + Lg idg0 !P LL + !P LL0 Lg idg (3.69) qg


The final 2 terms in (3.62) and (3.63) are the decoupling terms in the current controller, which are the feed-forward speed voltages. idg and iqg are changes in grid VSC currents. Vdg and Vqg are the measured grid voltages, which work as feed-forward terms to improve the stability of the current controller.

63

3.3 Small-Signal Model

3.3.7.3 Converter Impedance


The impedance of grid VSCs is a series RL circuit. The linearised equations are:

1 1 Rg _ idg = Vd vsc Vdg idg + !e0 iqg Lg Lg Lg 1 1 Rg _ iqg = Vq vsc Vqg iqg !e0 idg Lg Lg Lg

(3.70)

(3.71)

As with the generator VSCs, the changes in grid VSC voltages are caused by outputs from
the current controllers Vd vsc and Vq vsc , which are calculated in (3.68) and (3.69).

3.3.8 DC link and DC Voltage Controller


The linearised equation for DC voltage (3.31) is:

Pgen Pgrid _ Vdc = Cdc Vdc0


respectively. They are calculated as (3.42). The linearised DC voltage controller equations are:
_ Vdc = Vdc Vdc Pgrid = kp (Vdc Vdc ) + ki Vdc + Pgen

(3.72)

where Pgen and Pgrid are the active power from the generator side and grid side,

(3.73) (3.74)

where Pgen is the feed-forward term used in the DC voltage controller to improve stability; its linearised equation is in (3.42).

3.3.9 Phase Locked Loop


After linearising (3.32), the small-signal model of the PLL becomes:

_ P LL = 0 Vd _ !P LL = P LL = kp (0 Vd ) + ki P LL

(3.75) (3.76)

3.3.10 Interface between Unit and Grid


The interface between unit and grid in the small-signal model is derived from (3.33) and (3.34), and linearized as (3.77) and (3.78). xd cos 0 sin 0 xD xD0 sin 0 + xQ0 cos 0 = + (3.77) xq sin 0 cos 0 xQ xD0 cos 0 xQ0 sin 0 64

CH.3 Wind Generator System Modelling

xD cos 0 = xQ sin 0

sin 0 xd xd0 sin 0 xq0 cos 0 + (3.78) cos 0 xq xd0 cos 0 xq0 sin 0

3.3.11 Power Grid Model


The non-interconnected system model is:

_ x = Ax + Bv V + Bu u y = Cx + Dv V + Du u

(3.79) (3.80)

where V is the change in the node voltage, which are influenced by both x and u. For example, a reactive power reference change of one unit will not only influence its own terminal voltage, but also other node voltages in an interconnected system. A power grid model is required to eliminate states x and inputs u. The power grid is composed of transmission lines and transformers, the units of which are all connected together on the grid. Each line or transformer is modelled as a series RL circuit, and a small-shunt capacitance is modelled at each node (terminator of the unit) to convert the node voltages Vds and Vqs to the state variables. The linearised equations of the grid are written in (3.81) to (3.86).
1 1 _ iDDF IG iDcb + !e0 VQt VDt = Ct Ct 1 1 _ iQDF IG iQcb !e0 VDt VQt = Ct Ct

(3.81) (3.82) (3.83) (3.84) (3.85) (3.86)

1 1 Rg i_Dg = VDpcc VDg iDg + !e0 iQg Lg Lg Lg 1 1 Rg i_Qg = VQpcc VQg iQg !e0 iDg Lg Lg Lg
_ VDpcc = 1 1 iDcb iDg + !e0 VQpcc Cpcc Cpcc 1 1 iQcb iQg !e0 VDpcc Cpcc Cpcc

_ VQpcc =

where iDcb iQcb are the grid currents flowing through the low-voltage cable, iDDF IG and
iQDF IG

are the total currents flowing out of the DFIG, iDg iQg are the grid currents flowing

through the grid integration impedance, VDt and VQt are the terminal voltages of the DFIG,
VDpcc

and VQpcc are the voltages at the point of common coupling, VDg and VQg are the

65

3.3 Small-Signal Model

voltages of the infinite bus, and the subscripts D; Q denote those variables that are in the common reference frame, as has been discussed in section 3.2.8. However, this modelling method greatly increases the number of state variables and is therefore not efficient for simulation of a large system. The technique for large power system modelling will be discussed in Chapter 6.

3.3.12 Model Implementation


From equations (3.38) to (3.86), the complete system is interconnected and modelled, however only state variables and inputs are considered. A Matlab program is used to call the unit models to form a complete system. The flow chart of the Matlab program is shown in Figure 3-15. Once the state space model of the power system is derived, it is easy to use Matlab to analyse the system for many purposes. The eigen values of the state matrix A show the stability of the system, and the participation factor matrix indicates the degree of eigen values influenced by the state variables. With the calculated eigen value locus, the system stability trend becomes clear by the change in either the system operation points or control parameters. The transfer function between any inputs and outputs can also be easily derived and its Bode diagram can be used to tune the PI controllers. The model of a synchronous generator and its controller is described in the literature [Anderson and Fouad'03, Kundur'94, Rogers'00]. In this thesis, the small-signal model of Type C wind turbines (DFIG) is implemented. The model, which is controlled by FOCs, provides both grid-frequency and grid-voltage support. The DFIG is chosen for analysis because of its dominant market share [Ackermann'05, Hansen and Hansen'07, Shrestha, et al.'08], and its technical complexity. Combining equations from (3.49) to (3.78), the complete small-signal model of Type C (DFIG) wind turbines is derived. Those variables which are neither state variables nor inputs have to be eliminated from the equations. For example, the changes in the rotor fluxes dr and qr in (3.62) and (3.63) have to be substituted by (3.40) and (3.41). Also the change in the measured grid frequency !P LL in (3.62) and (3.63) has to be substituted by (3.76). The complete model of a DFIG is programmed as a function block in Matlab, which can be called by the main program to form a complete power system model as shown in Figure 3-15. The schematic of a smallsignal model of a DFIG is drawn in Figure 3-16, which has a total of 29 state variables. For Type D wind turbines, the generator model is the same as the SG, and the back-to-back VSC controller is the same as in Type C wind turbines. It can be implemented in a same way as Type C wind turbine developed in this thesis, thus will not be discussed in detail.

66

CH.3 Wind Generator System Modelling

Figure 3-15. Flow chart of small-signal stability program.

67

3.3 Small-Signal Model

Ps0 ; Qs0

Ps ; Q s

i ; i qr dr

Ps droop ; Qs droop

Taero
f; jVt j

!gen
Te

vdr ; vqr

vDt ; vQt

vdt ; vqt

ids ; iqs
idr ; iqr

iDs ; iQs

iD

df ig ; iQ df ig

vDt ; vQt

iD vsc ; iQ vsc

idr ; iqr
vdr ; vqr

Pr
Vdc
Pvsc

vdt ; vqt

i vsc ; i vsc q d

vd vsc; vq

vsc

Q vsc

id vsc ; iq

vsc

Pvsc

vd vsc; vq id vsc ; iq

vsc vsc

Figure 3-16. Small-signal model of a DFIG with a field-oriented current controller and droop controller.

3.3.13 Verification of Small-Signal Model of Type C Wind Turbines


The small-signal model is validated by comparing its step responses with those of the dynamic model, as the dynamic model of DFIG has already been verified in experiments [Petersson, Thiringer, et al.'05]. The reference frame is chosen in such a way that the gridvoltage vector is aligned with the q axis.

vds = 0 vqs = jVs j

(3.87) (3.88)

Thus, ids and idr are the reactive currents, and iqs and iqr are the active currents. A single DFIG with an infinite bus system (as shown in Figure 3-17) is modelled by both a dynamic modelling and small-signal modelling approach. The parameters of a DFIG are listed in Table 3-7. At 1.0s the infinite bus voltage, the - q axis voltage vQg drops to 0.9 pu for 1.0s, the simulated results of which are plotted from Figure 3-18 to Figure 3-21.

68

CH.3 Wind Generator System Modelling

Figure 3-17. Simulated single DFIG with an infinite bus system.

Table 3-7. DFIG and grid parameters [Pierik, et al.'04]. DFIG Pnom
Jwt

2.75MW 1.26e7kg m 240kg m 70.65 92m 1.18e-4H 9.1e-3H 1.2e-3


2 2

Rr

5.2e-3 2.5e-2,1.55e-1 3.2e-4,3.2e-3 2e-2,3,2 2e-1,3.2e-1 4% 1.25e-2 1e-4H

Cdc

1e-3F 2e4,2e5 1e-2,1e-1 1e-2,1e-1 5e1,1e3 1.6e-3 6.6e-5H 20

kp ; ki Te control kp ; ki Qs control kp ; ki id r control kp ; ki iq r control Dp ; Dq droop ratio Rconv Lconv

kp ; ki Vdc control kp ; ki id conv control kp ; ki iq Rtraf o Ltraf o


conv

Jgen

Gear ratio
Dwt Lls Lm Rs

control

kp ; ki P LL

SC Ratio

Because of the terminal voltage droop controller, the stator reactive power demand of the DFIG Qs ref increases during a 0.1pu grid-voltage dip, as shown in Figure 3-18. In Figure 3-19, the rotor current idr is controlled to increase its stator reactive current ids and its actual output stator reactive power Qs. From Figure 3-18 and Figure 3-19 it can be seen that for the reactive power components, the small-signal model agrees very well with the dynamic model. The small differences can be attributed to the following factors: firstly, some small capacitors of the dynamic model are omitted in the small-signal model to limit the number of integrators; secondly, in the small-signal model, the pre-disturbed working point is calculated, around which the model is linearised. However, during a grid-voltage dip, the working point of this nonlinear system changes.

69

3.3 Small-Signal Model

Changes in Stator Voltage 400 DFIG vqs change 200 0 -200 -400 0.5 x 10
6

Linear Model Dynamic Model

1.5

2.5

Changes in Stator Reactive Power Linear Model Dynamic Model

2 DFIG Qs change 1 0 -1

-2 0.5

1.5 time (s)

2.5

Figure 3-18. Changes in DFIG terminal voltage vqs and reactive power Qs.
Changes in d-axis Stator Current DFIG ids change 2000 0 -2000 0.5 1 1.5 Linear Model Dynamic Model 2 2.5

Changes in d-axis Rotor Current DFIG idr change 1000 0 -1000 0.5 1 1.5 time(s) Linear Model Dynamic Model 2 2.5

Figure 3-19. Changes in DFIG stator reactive current ids and rotor reactive current idr.

70

CH.3 Wind Generator System Modelling

Changes in q-axis Stator Current 500 DFIG iqs change Linear Model Dynamic Model 0

-500 0.5

1.5 2 time (s) Changes in q-axis Rotor Current

2.5

500 DFIG iqr change

0 Linear Model -500 0.5 Dynamic Model 1 1.5 time (s) 2 2.5

Figure 3-20. Changes in DFIG stator active current iqs and rotor active current iqr.
x 10
4

Changes in Electric Toruqe Linear Model Dynamic Model

1 DFIG Te change 0.5 0 -0.5

-1 0.5 x 10
-3

1.5 Changes in Generator Speed

2.5

DFIG wm change (pu)

Linear Model Dynamic Model

-5 0.5

1.5 time (s)

2.5

Figure 3-21. Changes in DFIG electric torque Te and generator speed -

71

3.3 Small-Signal Model

Figure 3-20 and Figure 3-21 prove that the stator active current iqs, active current iqr and electric torque Te changes in the small-signal model also agree very well with the dynamic model. However, there is very little difference in the steady state speed between the smallsignal and dynamic model, as shown in Figure 3-21. The speed difference may be caused by the fact that the friction loss of the turbine and generator included in the dynamic model is omitted in the small-signal model. However, it can be seen that the transient speed signal response of the small-signal model agrees quite well with dynamic model.

3.3.14 Stability Analysis of Type C Wind Turbines


As the VSWTs dominate the wind power market [Ackermann'05, Hansen and Hansen'07, Shrestha, et al.'08], VSWTs with FOCs have been well studied in recent years. Most of the research has focused on the dynamic modelling and LVRT capability [Akhmatov'05, Erlich, et al.'07, Hansen and Michalke'07, Hughes, et al.'05, Morren and de Haan'05, Morren, et al.'05, Petersson, Lundberg, et al.'05, Petersson, Thiringer, et al.'05, Xiang, et al.'06]. Dynamic analysis is necessary to ensure that the system remains operative under different system conditions and following the most common types of faults. However, the smallsignal model/linearised system model indicates whether the systems steady state is stable or not. In addition, this system model can be used to design controls which may be necessary to maintain the system's stability following small changes. The small-signal modelling and stability analysis of VSWTs is seldom taken into consideration and the control parameters of FOCs are often chosen arbitrarily. The PLL and droop controller can also influence the stability. Moreover, with the increase in wind power penetration, the following research questions can be raised: what is the small-signal stability of a power system with a large wind power installation? Is it necessary to install a PSS for the wind power plant similar to that for the thermal power plant? In [Akpinar, et al.'93], a slip energy recovery induction generator which uses a current source converter is analysed. In [Abdin and Xu'00], the small-signal model of a constant speed induction generator with a static Var compensator (SVC) is presented. In [Mei and Pal'07], an open-loop small-signal model of a DFIG is developed and the stability of an open-loop DFIG is analysed. In this section, the stability margin of a single DFIG with an infinite bus system is determined by considering the eigen values and participation factors of the closed-loop controlled system. From a technical point of view, the small-signal model of Type C VSWT is similar to, but even more complex than, Type D VSWTs.

72

CH.3 Wind Generator System Modelling

In Chapter 6, the small-signal model of a DFIG will be integrated into other power sources, such as large synchronous generators with a high-gain exciter, thus allowing an analysis of the small-signal stability of an interconnected power system which consists of both thermal power plants and wind power plants.

3.3.14.1 Overview of Eigen Values


A single DFIG with an infinite bus system as shown in Figure 3-17 is analysed, and its parameters including parameters of the DFIG and the grid are listed in Table 3-7. It has a total of 29 state variables, 19 of which are states of the DFIG with its transformer, and the other 10 states of which are of the shunt load and grid components. The eigen values at the operating point of rated wind speed 11.5m/s are listed in Table 3-8.
Table 3-8. Eigen values and mode descriptions at full power operating point. (* ) Damping ratio Frequency (Hz) 1 1 1 1 1 1 1 0.03185 3.38172 7.73885 10.93949 24.3949 46.28981 49.96815 385.3344 487.4841 925.4936 1025.876 Mode Description DFIG d axis rotor current controller DFIG q axis rotor current controller DFIG DC voltage controller grid VSC d axis current controller DFIG stator reactive power controller grid VSC q axis current controller DFIG torque controller turbine two mass shaft mode phase locked loop DFIG electro-mechanical mode DFIG rotor currents mode DFIG DC voltage mode DFIG stator currents mode shunt load mode grid electrical mode grid electrical mode grid electrical mode grid electrical mode

1 2 3 4 5 6 7 8,9 10,11 12,13 14,15 16,17 18,19 20,21 22,23 24,25 26,27 28,29

-0.2207 -0.1358 -0.0126 -0.0087 -0.005 -0.0045 -0.0001

-0.0017 0.0002i 0.9932 -0.0301 0.0213i 0.8162 -0.0212 0.0486i 0.3998 -0.0124 0.0687i 0.1776 -0.1261 0.1532i 0.6355 -0.0435 0.2907i 0.1480 -0.3099 0.3138i 0.7027 -0.1187 2.4199i 0.0490 -0.0667 3.0614i 0.0218 -0.0368 5.8121i 0.0063 -0.0228 6.4425i 0.0035

Eigen values 22 to 29 are high-frequency oscillation modes related to grid components of the series RL circuit and shunt capacitance; these values are usually eliminated in large

73

3.3 Small-Signal Model

power system analysis. However, they are included here because the grid inductance and shunt capacitance are modelled by integration equations.

3.3.14.2 Stability Margin of Grid Short Circuit Ratio


Decrease Short Circuit Ratio 80 Gen Rotor Elec Mode

70

imaginary

60 Gen Elect-Mech Mode

50

40

30 -25

-20

-15

-10

-5 real

10

15

Figure 3-22. Eigen value locus when short circuit ratio at PCC decreases from 20 to 1.

In Figure 3-22, point + denotes the operating point when the short-circuit (SC) ratio at PCC is 20; point o denotes the operating point when the SC ratio is 1; point denotes the boundary operating point when the SC ratio is 2.5. It is clear that the stability of the DFIG decreases when the grid strength (i.e., SC ratio) decreases. When the SC ratio is smaller than 2.5, the system becomes unstable.

3.3.14.3 Stability Margin of Phase Locked Loop


Figure 3-23 shows the eigen value locus when and and when and of the phase-locked loop increase from 10% to 500% of the values in Table 3-7. Point + denotes the operating point when of the phase-locked loop are 10%; point o denotes the operating point when of the phase-locked loop are 500%; point denotes the boundary operating point and of the phase-locked loop are 400%. It can be seen that when and are

400% of the values of Table 3-7, the stator current mode becomes unstable.

74

CH.3 Wind Generator System Modelling

increase PLL Gain k p _ PLL k i _ PLL 300 250 Stator Currents Mode 200 imagine 150 DC Link Mode 100 50 0 -120 PLL Mode -100 -80 -60 real
Figure 3-23. Eigen value locus when increasing the kp and ki of the PLL.

-40

-20

3.3.14.4 Stability Margin of Droop Controller


When the droop ratio and of the frequency and voltage droop controllers increase, one high-frequency electrical mode becomes unstable. This high-frequency electrical mode is the mode of the cable and transformer inductors and capacitors, which build up a grid LCL filter. Point + denotes the operating point when denotes the operating point when is quite stable. and and are 0.01; point o are 0.5; point denotes the boundary

condition when Dp and Dq equal 0.24. The normal droop ratio is 0.04 [Kundur'94], which

75

3.3 Small-Signal Model

increase droop ratio Dp Dq 4000


Grid LCL Filter

3000 imaginary
Grid LCL Filter

2000

1000

0 -600

-500

-400

-300

-200 real

-100

100

200

Figure 3-24. Eigen value locus when increasing the droop ratios Dp and Dq from 0.01 to 0.5.

3.3.14.5 Stability Margin of Generator Field-Oriented Controller


Figure 3-25 shows the eigen value locus when the and of the torque and reactive and power controllers of the DFIG increase. Point + denotes the operating point when point o denotes the operating point when when the and and

of the torque and reactive power controllers are 100% of the values listed in Table 3-7; of the torque and reactive power controllers are 5000% of the values of Table 3-7; point denotes the boundary condition are 1200% of the values of Table 3-7.

76

CH.3 Wind Generator System Modelling

increase Kp Ki of Torque controller 4000

3000 imaginary

Grid LCL Filter Mode

2000

1000

0 -600

-400

-200 real

200

400

increase Kp Ki of Qs controller 4000 Grid LCL Filter Mode

3000 imaginary

2000

1000

0 -600

-400

-200 real

200

400

Figure 3-25. Eigen value locus when the kp and ki of the torque and reactive power controller of DFIG increase.

77

3.3 Small-Signal Model

3.3.14.6 Stability Margin of DC Voltage Controller


increase Kp Ki of DC voltage controller 4000 Grid LCL Filter 3000 imaginary Grid LCL Filter

2000 DFIG DC Link 1000

0 -600

-400

-200 real

200

400

Figure 3-26. Eigen value locus when.the kp and ki of the DC voltage controller increase.

increase Kp Ki of DC voltage controller 4000 Grid LCL Filter 3000 imaginary Grid LCL Filter

2000 DFIG DC Link 1000

0 real Figure 3-26 shows the eigen value locus when the

0 -600

-400

-200

200

400

and

of the DC voltage controller

increase. Point denotes the boundary condition when kp and ki of the DC voltage controller are 105 times larger than the values of Table 3-7.

78

CH.3 Wind Generator System Modelling

3.3.14.7 Influence of Operation Points


The root locus during the change in the operation points is plotted in Figure 3-27 and Figure 3-28. The wind speed increases from 6m/s to 15m/s, which corresponds to rotational speed increasing from 0.7pu to 1.2pu. The point indicated by + denotes the operation point at the lowest wind speed (6m/s) or lowest rotational speed (0.7pu); point o denotes the operation point at the highest wind speed (15m/s) or highest rotational speed (1.2pu); point denotes the operation point at the synchronous rotational speed (1.0pu). It can be seen that all the operation points are stable. The damping of the electro-mechanical mode reaches the maximum at the synchronous speed, as shown in Figure 3-28, which is similar to the analysis in [Banakar, et al.'06].
350 300 250 imagine 200 DC Link Mode 150 100 50 0 -250 Non oscillatory Mode -200 -150 real
Figure 3-27. DFIG eigen value locus when rotational speed increases from 0.7pu to 1.2pu.

Gen Stator Elec Mode

Gen Rotor Elec Mode Gen Elect-Mech Mode PLL Mode -100 -50 0

79

3.3 Small-Signal Model

Electro-mechanical Mode When Rotation Speed Increases from 0.7 - 1.2 pu 50 49 48 imaginary 47 46 45 44 -24

-23.5

-23

-22.5

-22 real

-21.5

-21

-20.5

-20

Figure 3-28. Eigen value locus of DFIG electro-mechanical mode when rotational speed increases from 0.7pu to 1.2pu.

3.3.14.8 Summary of DFIG Stabilty


From the above analysis, the stabilities of a grid connected wind turbine with DFIG are as the followings: The DFIG is stable under the full speed operation range that is the rotation speed between 0.7pu to 1.2pu. The rotor current mode of DFIG becomes unstable when the short circuit ratio at PCC decreases to 2.5pu. Under the normal when of phase locked loop (50, 1000), all modes of grid connected DFIG are stable. The stator current mode of DFIG becomes unstable of phase locked loop increases to 200, 4000 respectively. Under the normal droop ratio 4%, the DFIG is stable with both grid frequency and terminal voltage droop controller. However, one grid LCL filter mode (mode 26,27 in Table 3-8) becomes unstable when droop ratio increases to 24%. The grid LCL filter mode (mode 26,27 in Table 3-8) is also strongly influenced by electrical torque, reactive power controller and DC voltage controller DFIG. The grid LCL filter is composed by inductor of transformer, shunt capacitor and inductor of cable as shown in Figure 3-17.

80

CH.3 Wind Generator System Modelling

Thus, all the modes of DFIG itself (mode 1 to 19 in Table 3-8) are stable under the most investigated conditions except for the decreasing of short circuit ratio at PCC. However, the stability of a grid connected DFIG is determined by the grid LCL filter mode (mode 26,27 in Table 3-8), which becomes unstable when increasing of PLL, droop ratio, electrical torque controller, reactive power controller and DC voltage controller. This implies that this grid LCL filter mode should be considered in the controller design of wind turbine using DFIG.

3.4 Conclusion
In this chapter, dynamic and small-signal models of VSWTs are derived. The dynamic model can be used to study the dynamic response of wind turbines, for example, the lowvoltage ride through capability. The small-signal model is used for control design and stability analysis. It can also be used to analyse low-frequency power system oscillations with large-scale wind power integration. It will be discussed in further detail in Chapter 6. The dynamic model of a Type C wind turbine is verified through experiment. Its responses during symmetrical and unsymmetrical grid-voltage dips agree quite well with experiment results. The dynamic simulations of three different dynamic models (i.e., different control and modelling approaches) are compared. They are FOC with an average model, FOC with a switching model, and DTC with a switching model. It is concluded that FOC with an average model are the fastest; moreover, they are accurate enough for electro-mechanical simulation. In the next chapters, FOC with an average model will be used in all the analyses, unless otherwise indicated. The small-signal model validation is completed by comparing the step responses of the small-signal model and dynamic model of a single type C wind turbine with an infinite bus system. The results prove the accuracy of the small-signal model. The stability of a single type C wind turbine integrated with an infinite bus system is analysed, which determines the stability margin of the closed-loop controlled type C wind turbine.

81

3.5 References

3.5 References
[Abd00] E. S. Abdin and W. Xu, "Control design and dynamic performance analysis of a wind turbine - induction generator unit," IEEE Transactions on Energy Conversion, vol. 15, issue 1, Page(s). 91-96, 2000. T. Ackermann, Wind power in power systems, ISBN-0-470-85508-8, John Wiley & Sons, 2005. V. Akhmatov, Analysis of dynamic behavior of electric power systems with large amount of wind power, PhD thesis, Electrical Power Engineering Department, Technical University of Denmark, Copenhagen, Denmark, 2003. V. Akhmatov, "Modelling and ride-through capability of variable speed wind turbines with permanent magnet generators," Wind Energy, vol. 9, issue 4, 2005. E. Akpinar, R. E. Trahan, and A. D. Nguyen, "Modeling and analysis of closed-loop slip energy recovery induction motor drive using a linearization technique," IEEE Transactions on Energy Conversion, vol. 8, issue 4, Page(s). 688-697, 1993. P. M. Anderson and A. A. Fouad, Power system control and stability, ISBN0-471-23862-7 Wiley-Interscience, New York, 2003. J. Arrillaga and C. P. Arnold, Computer analysis of power systems, John Wiley & Sons, Inc. New York, NY, USA, 1990. S. Aurtenechea, M. A. Rodrguez, E. Oyarbide, and J. R. Torrealday, "Predictive direct power control-a new control strategy for dc/ac converters," IEEE Transactions on Industrial Electronics, 2006. H. Banakar, C. Luo, and B. T. Ooi, "Steady-state stability analysis of doublyfed induction generators under decoupled P-Q control," IEE ProceedingsElectric Power Applications, vol. 153, issue 2, Page(s). 300, 2006. B. K. Bose, Modern Power Electronics and AC Drives, ISBN-7-111-11296-2, Prentice Hall, 2002. I. Erlich, H. Wrede, and C. Feltes, "Dynamic Behavior of DFIG-Based Wind Turbines during Grid Faults," in IEEE Power Conversion Conference, 2007, Nagoya, Japan, 2007, Page(s). 1195-1200. A. D. Hansen and L. H. Hansen, "Wind Turbine Concept Market Penetration over 10 Years (1995-2004)," Wind Energy, vol. 10, issue 1, Page(s). 8197, 2007. A. D. Hansen and G. Michalke, "Fault ride-through capability of DFIG wind turbines," Renewable Energy, vol. 32, issue 9, Page(s). 1594-1610, 2007. L. Harnefors and H. P. Nee, "A general algorithm for speed and position estimation of AC motors," IEEE Transactions on Industrial Electronics, vol. 47, issue 1, Page(s). 77-83, 2000. F. M. Hughes, O. Anaya-Lara, N. Jenkins, and G. Strbac, "Control of DFIGBased Wind Generation for Power Network Support," IEEE Transactions on Power Systems, vol. 20, issue 4, Page(s). 1958-1966, 2005.

[Ack05] [Akh03]

[Akh05]

[Akp93]

[And03] [Arr90] [Aur06]

[Ban06]

[Bos02] [Erl07]

[Han07]

[Han07] [Har00]

[Hug05]

82

CH.3 Wind Generator System Modelling

[Kau97]

[Kaz02] [Kra86] [Kun94] [Mal04]

[Mei07]

[Mor05]

[Mor05]

[Ong98] [Par29] [Pet05]

[Pet05]

[Pie04]

[Rog00] [Shr08]

[Eng03]

V. Kaura, V. Blasko, R. Autom, A. B. Co, and W. I. Mequon, "Operation of a phase locked loop system under distorted utilityconditions," IEEE Transactions on Industry Applications, vol. 33, issue 1, Page(s). 58-63, 1997. M. P. Kazmierkowski, R. Krishnan, and F. Blaabjerg, Control in Power Electronics: Selected Problems, Academic Press, 2002. P. C. Krause, O. Wasynczuk, and S. D. Sudhoff, Analysis of electric machinery, ISBN 0-7803-1101-9, McGraw-Hill Singapore, 1986. P. Kundur, Power System Stability and Control, ISBN-0-07-035958-x, McGraw-Hill Professional, 1994. M. Malinowski, M. Jasinski, and M. P. Kazmierkowski, "Simple direct power control of three-phase PWM rectifier using space-vector modulation (DPC-SVM)," IEEE Transactions on Industrial Electronics, vol. 51, issue 2, Page(s). 447-454, 2004. F. Mei and B. Pal, "Modal Analysis of Grid-Connected Doubly Fed Induction Generators," IEEE Transactions on Energy Conversion, vol. 22, issue 3, Page(s). 728-736, 2007. J. Morren and S. W. H. de Haan, "Ride through of wind turbines with doubly-fed induction generator during a voltage dip," IEEE Transactions on Energy Conversion, vol. 20, issue 2, Page(s). 435-441, 2005. J. Morren, J. T. G. Pierik, S. W. H. de Haan, and J. Bozelie, "Grid Interaction of Offshore Wind Farms. Part 1. Models for Dynamic Simulation," Wind Energy, vol. 8, issue 3, Page(s). 265-278, 2005. C. M. Ong, Dynamic Simulation of Electric Machinery: Using MATLAB/SIMULINK, ISBN 0-13-723785-5, Prentice Hall PTR, 1998. R. H. Park, "Two Reaction Theory of Synchronous Machines," AIEE Transactions, vol. 48, Page(s). 716-730, 1929. A. Petersson, S. Lundberg, and T. Thiringer, "A DFIG Wind-Turbine RideThrough System Influence on the Energy Production," Wind Energy, vol. 8, issue 3, Page(s). 251-263, 2005. A. Petersson, T. Thiringer, L. Harnefors, and T. Petru, "Modeling and experimental verification of grid interaction of a DFIG wind turbine," IEEE Transactions on Energy Conversion, vol. 20, issue 4, Page(s). 878-886, 2005. J. T. G. Pierik, J. Morren, E. J. Wiggelinkhuizen, S. W. H. de Haan, T. G. van Engelen, and J. Bozelie, "Electrical and Control Aspects of Offshore Wind Farms II (Erao II) - Volume 1: Dynamic Models of Wind Farms," Energieonderzoek Centrum Nederland (ECN), Petten 2004. G. Rogers, Power System Oscillations, ISBN-0-7923-7712-5, Kluwer Academic Pub, 2000. G. Shrestha, H. Polinder, D. J. Bang, and J. A. Ferreira, "Review of Energy Conversion System for Large Wind Turbines," in European Wind Energy Conference 2008, EWEC2008, Brussels, Belgium, 2008. T. G. Van Engelen and E. J. Wiggelinkhuizen, "ECN Design Tool for Control Development Revised Points of Departure," Energieonderzoek Centrum Nederland (ECN), Petten 2003.

83

3.5 References

[Xia06]

[Zho02]

[Zho06]

[Zho09]

D. Xiang, L. Ran, P. J. Tavner, and S. Yang, "Control of a doubly fed induction generator in a wind turbine during grid fault ride-through," Energy Conversion, IEEE Transaction on, vol. 21, issue 3, Page(s). 652-662, 2006. K. Zhou and D. Wang, "Relationship between space-vector modulation and three-phasecarrier-based PWM: a comprehensive analysis," IEEE Transactions on Industrial Electronics, vol. 49, issue 1, Page(s). 186-196, 2002. Y. Zhou, P. Bauer, J. A. Ferreira, and J. Pierik, "Aggregated models of offshore wind farm components for grid study," in IEEE Power Electronics Specialists Conference, 2006, Korea, 2006. Y. Zhou, J. Pierik, P. Bauer, and J. A. Ferreira, "Operation of GridConnected DFIG Under Unbalanced Grid Voltage Condition," IEEE Transactions on Energy Conversion, vol. 24, issue 1, Page(s). 240-246, 2009.

84

CH.4 Grid Integration of Individual Variable Speed Wind Turbine

Chapter 4
Grid Integration of Individual Variable Speed Wind Turbine 4.1 Introduction
The controller and the model of an individual VSWT are discussed in Chapter 2 and Chapter 3. General wind turbine integration problems and solutions are also reviewed. However, some grid integration problems of individual VSWTs are seldom treated and require further studies. The low-voltage ride through (LVRT) capability of wind turbines has been well studied, but often only the balanced grid-voltage dip is considered. During a balanced grid-voltage dip, the LVRT problem is the over-current of the VSC. During an unbalanced grid-voltage dip, the most severe problem is the electric torque oscillation and DC voltage ripple. The cause of and solution for LVRT problems during an unbalanced grid-voltage dip need further study. In traditional power systems, in which a synchronous generator (SG) dominates, the rotation speed of the SG is directly linked to the grid electric frequency. The grid frequency is considered to be mainly related to active power balance due to the swing function of the SG, and the grid voltage is mainly related to reactive power balance because of the large X/R ratio of the transmission line of the power system. However, the rotation speed of VSWTs is completely decoupled from the grid electric frequency by using power electronic converters. Furthermore, the trend of integrating a motor load with a power electronic converter attenuates this problem. As installation of wind power and other renewable energy sources increase drastically, the total inertia of a power system decreases. In a low-inertia power system, the relations of the grid frequency and the voltage to active and reactive power have to be re-investigated, and the operation of VSWTs in such a low-inertia power system has to be studied. In this chapter, the above-mentioned aspects are discussed and solutions are proposed, including VSWTs in an unbalanced grid-voltage condition, and VSWT operation in a lowinertia power grid.

85

4.2 Unbalanced Grid-Voltage Conditions

4.2 Unbalanced Grid-Voltage Conditions


Nowadays, wind turbines are required to have LVRT capability. For CSWTs (Type A wind turbines), the LVRT problem is the reactive power consumption by the wind turbine which may lead to voltage instability, and which is solved by dynamic reactive power compensation using an SVC or a STATCOM [Akhmatov'03, Chompoo-inwai, et al.'05, Molinas, et al.'08]. For VSWTs (Type C/D wind turbines), the LVRT problem is the transient over-current of power electronic converter. Crowbar protection [Akhmatov'03, Morren and de Haan'05, Petersson'05, Petersson, et al.'05, Xiang, et al.'06] is designed to short-circuit the Type C wind turbines rotor during a symmetrical grid-voltage dip, both to protect the rotor side VSC and to damp out the oscillations faster. Over-current protection and a DC chopper are designed for Type D wind turbines to protect the VSC and DC capacitance [Akhmatov'03, '05]. Most of these studies focus on the behaviour and protection of wind turbines during a symmetrical voltage dip. In reality, however, unsymmetrical faults occur much more frequently than symmetrical faults. Furthermore, it has been identified that the unbalanced voltages can occur in a weak power grid even during normal operation. During unbalanced grid voltage, the most severe operation problem is not the transient over-current, but both the large electric torque pulsations, which cause wear and tear of the gearbox and decrease its lift-time, and the large voltage ripple in the DC link of the back-to-back VSCs, which may decrease the lifetime of the DC capacitance. For Type C, both electric torque oscillation and DC voltage ripple exist because the stator is directly connected to the power grid. For Type D, only DC voltage ripple exists, because the full-scale back-to-back VSCs completely decouple the generator from power grid. As discussed above, Type C wind turbines (DFIG) have been selected to study the problem of VSWT under unbalanced grid-voltage dip, because they have both torque oscillation and DC voltage ripple under unbalanced grid-voltage conditions. In the literature [Bendl, et al.'05, Brekken and Mohan'07, Pena, et al.'07, Seman, et al.'06], researches have been done on the control system of DFIGs under unbalanced grid-voltage conditions. [Brekken and Mohan'07] use a feed-forward loop on the classical field-oriented current (FOC) controller to limit torque ripple, which is simple and robust. However, the loop performance depends on the filter. In [Pena, et al.'07], the grid VSC is controlled as a STATCOM and supplies reactive current to compensate the unbalanced grid voltage. Performance depends on both the current ratings of the grid VSC as well as impedance between the generator terminal and fault location. A control scheme is proposed to control DFIGs during unbalanced grid-

86

CH.4 Grid Integration of Individual Variable Speed Wind Turbine

voltage disturbances [Bendl, et al.'05]. The control objective is to limit the stator current unbalance, but not to limit the electric torque oscillation or the DC voltage ripple. In all, a theoretical analysis of DFIGs during unsymmetrical faults is still missing, and a systematic approach to limit both torque pulsation and DC voltage ripple is required. In this thesis, the behaviour of DFIGs under unbalanced grid-voltage conditions is thoroughly analysed, and a dual-sequence FOC controller is proposed based on symmetrical components theory [Lyon'54, Paap'00] and instantaneous reactive power theory [Akagi, et al.'84, Nabae and Tanaka'96]. The rotor VSC is controlled to limit the electric torque pulsation, and the grid VSC is controlled to limit the DC voltage ripple. In the following section, DFIG under unbalanced grid-voltage conditions will be compared by means of simulation using the dynamic model developed in Chapter 3.

4.2.1 Symmetrical vs. Unsymmetrical Voltage


Simulation results of DFIGs during symmetrical and unsymmetrical voltage dips are shown in Figure 4-1. Figure 4-1 shows that during symmetrical voltage dips, the most severe problem for DFIGs is the transient over-current in the rotor, which may damage the rotor VSC. This overcurrent problem is generally protected by what is referred to as active crowbar protection [Akhmatov'03, Morren and de Haan'05, Petersson'05, Petersson, et al.'05, Xiang, et al.'06], which uses a thyristor-controlled resistor bank to short-circuit the rotor windings. On the other hand during unsymmetrical voltage dips, the maximum rotor currents can be smaller, but they have 100 Hz oscillations. The torque oscillations and DC voltage ripples can also be observed in Figure 4-1. It is worth noting that the magnitude of rotor-current oscillations depends on the moment of the voltage dip. The starting point of the voltage dip determines the initial conditions of the stator currents, and thus determines the natural response of the stator currents. The natural response of stator currents has a large influence on the total transient rotor currents. The cause of electric torque oscillation and DC voltage ripple under unbalanced gridvoltage conditions will be analysed in the following sections using the instantaneous reactive power theory [Akagi, et al.'84, Nabae and Tanaka'96].

87

4.2 Unbalanced Grid-Voltage Conditions

Stator Voltage Under Symmetrical and Unsymmetrical Voltage Dips -600 Vqs -800 -1000

5.1

5.2

5.3

5.4

Rotor Current Under Symmetrical and Unsymmetrical Voltage Dips 2000 Idr 0 -2000 5 2 5.1 5.2 5.3 5.4

4 x 10 Torque Under Symmetrical and Unsymmetrical Voltage Dips

Te

-2

5.4

5.5

5.6

5.7

5.8

DC Voltage Under Symmetrical and Unsymmetrical Voltage Dips 1400 Vdc 1200 1000 5.4 5.5 5.6 time (s) 5.7 5.8

Figure 4-1. Simulation results of DFIG during symmetrical and unsymmetrical voltage dips. Solid line DFIG under three phase 30% voltage dip, dotted line DFIG under single phase 30% voltage dip.

88

CH.4 Grid Integration of Individual Variable Speed Wind Turbine

4.2.2 Symmetrical Component and Instantaneous Reactive Power Theory


An unsymmetrical system can be broken down into positive, negative and zero sequences [Lyon'54, Paap'00]. The positive and negative sequences are balanced three-phase systems. To fully decouple positive and negative sequences, the following assumptions are made: DFIG stator and rotor windings are assumed to be symmetrical. Grid VSC three-phase AC inductances and resistances are symmetrical.

Instantaneous reactive power theory [Akagi, et al.'84, Nabae and Tanaka'96] provides the definition of instantaneous active and reactive power, i.e. they are the production of instantaneous voltages and currents. Here, only a three-phase three-wire system is studied, where zero sequence is omitted. The choice is justified by the following reasons: The transformer is often Y/-connected. The neutral point of stator winding of DFIG is not grounded. and in a stationary

In a three-phase three-wire system, the instantaneous power reference frame are given in (4.1): p v i + v i = q v i + v i The stationary as in (4.2):

(4.1)

Equation (4.1) is true for both balanced and unbalanced three-phase three-wire systems. reference frame can be decoupled into positive and negative sequences

+ f f + f = + f f + f

(4.2)

The voltages and currents are further transformed into positive dq and negative dq reference frames, because they become DC values at steady state, which means a simple PI controller can be used. With Eq. (4.1), (4.2) and Parks transformation, the instantaneous active and reactive power in positive and negative dq sequences can be derived as (4.3): + + p p+p ~ vd id + vq + iq + + vd id + vq iq = = q q+q ~ vq + id + vd + iq + + vq id vd iq + v i vq + id + vq id + vd iq + (4.3) + d+ q sin 2 vd id + vq + iq + vd id + + vq iq + + v i + vq + iq + vd id + + vq iq + + d + d cos 2 vd iq vq + id + vq id + vd iq + The terms of sin 2 and cos 2 in (4.3) are the oscillating instantaneous active and reactive

~ ~ ~ power p , q . The oscillation active power p causes DC voltage ripple and electric torque
oscillation, as shown in Figure 4-1. The oscillating active power has to be maintained at

89

4.2 Unbalanced Grid-Voltage Conditions

zero to eliminate DC voltage ripple and torque pulsation. For current controllers of rotor and grid VSCs, the reference currents can be derived by rearranging Eq. (4.3). The controllers will be referred to as dual-sequence current controllers in this thesis.

4.2.3 Real-Time Signal Processing


The proposed dual-sequence current controller requires fast and accurate separation of positive and negative sequences. The two methods most often used are the low-pass filter method and the signal-delay cancellation method [Saccomando and Svensson'01]. These two methods will be reviewed and compared. The low-pass filter method is shown in Figure 4-2. The negative sequence component appears as a second order harmonic in the synchronous rotating frame (positive dq ), and the positive sequence component appears as a second order harmonic in the negative synchronous rotating frame (negative dq ). Low-pass filters are used to bypass the DC values and suppress the high-frequency oscillations. Thus the positive and negative sequences are separated in real time.
Positive Park Transform Positive dq Reference Low Pass Filter Positive Sequence

abc Negtive Park Transform Negtive dq Reference Low Pass Filter Negtive Sequence

Figure 4-2. Separated positive and negative sequences by low-pass filters.

The second method is the signal-delay cancellation method, which is shown in Figure 4-3. The abc system is first transformed into a stationary reference frame using Clarks transformation, then it is delayed for T/4. The positive and negative sequences can be calculated by adding or subtracting the present real-time signal with the delayed signal. This process will now be explained in detail.

90

CH.4 Grid Integration of Individual Variable Speed Wind Turbine

Tdq ( )
abc Clark Transform

Equation (4.6), Separate positive and negative sequence in a

Positive Sequence

Tdq ( ) Sequence
Figure 4-3. Separate positive and negative sequences by the signal-delay cancellation method.

Negtive

In the signal-delay cancellation method, the abc system can be transformed into a stationary reference frame using Clarks transformation and can be expressed in positive and negative sequence as (4.4): + + v (t) v (t) + v (t) v cos(!t + + ) + v cos(!t + ) = = + v (t) v sin(!t + + ) + v sin(!t + ) v (t)+ + v (t) Delaying (4.4) for T=4 is given by: # " + v (t T ) v (t T ) + v (t T ) 4 4 4 = + v (t T ) v (t T ) + v (t T ) 4 4 4 + v sin(!t + + ) v sin(!t + ) = v + cos(!t + + ) + v cos(!t + ) By combining (4.4) and (4.5), the positive and negative sequences in the reference frame can be calculated as (4.6): 2 2 + 3 v (t) 1 0 0 1 6 v (t)+ 7 1 6 0 1 1 0 6 7 6 4 v (t) 5 = 2 4 1 0 0 1 0 1 1 0 v (t) (4.8):

(4.4)

(4.5)

stationary

32

The voltages can be further transformed into a dq rotational reference frame with (4.7) and

3 v (t) 76 7 v (t) 76 7 5 4 v (t T =4) 5 v (t T =4)

(4.6)

+ + vd (t) cos() sin() v (t) = sin() cos() vq (t)+ v (t)+ vd (t) cos() sin() v (t) = sin() cos() vq (t) v (t)
without applying the two separation methods, are shown in Figure 4-4.

(4.7)

(4.8)

The unbalanced voltages in the positive synchronous rotating dq reference frame, with and

91

4.2 Unbalanced Grid-Voltage Conditions

x 10

-0.5

x 10

positive vd v

-1 0 without seperation signal delay cancellation low pass filter 5 time s 5.05 vq v -1.5

-5 4.95

-2 4.95

without seperation signal delay cancellation low pass filter 5 time s 5.05

(a) (b) Figure 4-4. Unbalanced voltages in a positive dq reference frame with two separation methods (a) vd (b) vq.

Without the two separation methods, the unbalanced three-phase voltage and current have second-order harmonics (100Hz) in both positive and negative synchronous rotating dq reference frames. Both separation methods have time delays before their outputs can reach the steady states. Because the signal-delay cancellation is much faster than the low-pass filter method, the signal-delay cancellation method is adopted in the dual-sequence controller of Type C wind turbines for unsymmetrical voltage dips.

4.2.4 Dual-Sequence Rotor VSC Current Controllers 4.2.4.1 Induction Generator Equation
With the assumption that the DFIG itself is symmetrical, the generator equations in positive and negative dq rotational reference frames are listed as (4.9): 3 2 3 2 ds Rs ids (!s qs ) + ddt vds 7 dqs 6 vqs 7 6 Rs iqs + (!s ds ) + dt 7 6 7=6 4 vdr 5 6 R i (! ! ) + ddr 7 4 5 r dr s r qr dt vqr dqr Rr idr + (!s !r )dr + dt

(4.9)

where !s is the stator electrical angular velocity, and !r is both the production of mechanical speed !m and the number of the pole pair . The positive and negative sequences are completely decoupled, as shown in (4.9), thus the DFIG can be controlled in positive and negative dq sequences independently.

92

CH.4 Grid Integration of Individual Variable Speed Wind Turbine

4.2.4.2 Controller Design


The dual-sequence current controller of rotor VSCs is described in Figure 4-5 (a). The grid voltage and current are transformed into positive and negative sequences using the signaldelay cancellation method derived from equations (4.6) to (4.8). The active power reference , is derived from the power-speed curve of the wind turbine, and the is derived from the AC voltage droop controller. With , , , , and ready, the references for the state currents in both positive and can be calculated using equation (4.10). The state are controlled by the rotor VSC voltages with rotor reactive power reference negative sequences reference currents

current PI controllers. There are cross-coupling terms between the direct and quadrature axes, namely (!s !r )dr and (!s !r )qr in (4.9). Those cross-coupling terms are added to the outputs of the PI controllers as feed-forward terms. Finally, the negative sequence rotor voltage the output simulation is in a positive 2 3 2 p Pref 6 q 7 6 Qref 6 7 6 4 psin 2 5 = 4 0 ~ pcos 2 ~ 0 has to be converted into a positive sequence and added to of the positive sequence rotor current controller because the reference frame. 3 2 + + vd id + vq + iq + + vd id + vq iq 7 6 vq + id + vd + iq + + vq id vd iq 7=6 + 5 4 vd iq vq + id + vq id + vd iq + vd + id + vq + iq + vd id + + vq iq +

3 7 7 5

(4.10)

4.2.5 Dual-Sequence Grid VSC Current Controllers 4.2.5.1 Grid VSC Impedance Equation
vd conv vq conv = " vd grid + Rid + (!s Liq ) + L did dt vq
grid

+ Riq (!s Lid ) + L

diq dt

(4.11)

where !s is the electrical angular velocity of the grid. The positive and negative sequences are completely decoupled, as shown in (4.11), thus the grid VSC can be controlled in positive and negative dq sequences independently.

4.2.5.2 Controller Design


The grid VSC controller is designed from (4.11), which uses grid converter voltages to control grid currents, as shown in Figure 4-5 (b). The rotor reference currents are calculated from DC link voltage, grid AC voltage, and rotor oscillation power, as shown in

93

4.2 Unbalanced Grid-Voltage Conditions

(4.12). Unlike normal VSCs during unbalanced voltage [Saccomando and Svensson'01], the DC voltage ripple in the DC link of the DFIG is caused by both the rotor VSC and grid VSC. Thus, the instantaneous active power of the grid VSC has to be controlled in coordination with the rotor VSC in order to eliminate the DC voltage ripple. Feed-forward loops are used to compensate rotor oscillation power psin 2 and pcos 2, as ~ ~ shown in Figure 4-5 (b) . 2 3 2 p Pref 6 q 7 6 Qref 7 6 .6 4 psin 2 5 = 4 psin 2 Rotor ~ pcos 2 ~ pcos 2 Rotor

In grid VSC controllers, the cross-coupling terms between d and q are !s Lid and the negative sequence rotor voltage and added to the output the simulation is in a positive has to be converted into a positive sequence

3 vd + id + + vq + iq + + vd id + vq iq 7 6 vq + id + vd + iq + + vq id vd iq 7 7=6 + 7 5 4 vd iq vq + id + vq id + vd iq + 5 (4.12) vd + id + vq + iq + vd id + + vq iq + 3 2

!s Liq . They are feed-forwarded and added to the outputs of the PI controllers. Finally,
of the positive sequence rotor current controller, because reference frame.

4.2.6Simulation Results
The effects of the proposed dual-sequence FOC controllers can be seen in Figure 4-6. The new rotor VSC controller limits the torque pulsation to less than 20% of the original value, while the grid VSC limits the DC voltage ripple. The electric torque, DC voltage, stator active and reactive power are compared in Figure 4-7, which shows the zoomed in view of Figure 4-6. When the dual-sequence FOC is applied to the rotor VSC, the stator power oscillation is limited, but the rotor power oscillation increases slightly. As the stator power often dominates the total power of the generator, the total power oscillation is still significantly limited. When the grid VSC is controlled separately in positive and negative sequences and in coordination with the rotor VSC, the DC voltage ripple is drastically limited to less than 10% of the original value. The coordination of grid VSC and rotor VSC means that the instantaneous active power of grid VSCs is intentionally controlled to have second order ~ ~ harmonics in order to compensate the rotor power oscillation Psin 2 and Pcos 2.

94

CH.4 Grid Integration of Individual Variable Speed Wind Turbine

vs _ dq _ pos

vs _ dq _ neg

vg _ dq _ pos

vg _ dq _ neg

Rotor VSC controller

Grid VSC controller

Ps ref
Qs ref
Stator reference currents calculation

Pcref
Qcref
Grid VSC reference currents calculation

Psin 2
Pcos 2

is _ ref _ dq _ neg

is _ ref _ dq _ pos + is _ dq _ pos is _ dq _ neg


PI

igc _ ref _ dq _ neg

igc _ ref _ dq _ pos + igc _ dq _ pos igc _ dq _ neg


PI

PI

PI

dq decoupler

dq decoupler

vr c _ dq _ neg vr c _ dq _ pos
Neg Pos

vgc _ dq _ neg vgc _ dq _ pos


Neg Pos

vrotor _ conv _ dq

vgrid _ conv _ dq

(a)

(b)

Figure 4-5. Dual-sequence current controller of DFIG, (a) controller of rotor VSC, (b) controller of grid VSC

95

4.2 Unbalanced Grid-Voltage Conditions

4 2 Te 0 -2

x 10

Torque with Single Sequence Controller

5 x 10
4

5.5

6.5

Torque with Dual Sequence Controller

4 2 Te 0 -2

5.5

6.5

DC Voltage with Single Sequence Controller 1300 1200 Vdc 1100 1000 5 5.5 6 6.5 7

DC Voltage with Dual Sequence Controller 1300 1200 Vdc 1100 1000 5 5.5 6 time (s) 6.5 7

Figure 4-6. Comparison of the DFIG responses during unsymmetrical voltage dips from Single sequence FOC controller and Dual-sequence FOC controller.

96

CH.4 Grid Integration of Individual Variable Speed Wind Turbine

4 2

x 10

1300 1200 Vdc

Te

1100 1000 900

0 -2 6 4 2 Ps 0 -2 6 6.02 6.04 time (s)

6.02

6.04

800 6 2 1 Pr 0 -1 6

6.02

6.04

6.02 6.04 time (s)

Figure 4-7. Comparison of the DFIG responses during an unsymmetrical voltage dip from different controllers. Dotted line: Classical Single sequence FOC controller; Solid line: Proposed Dualsequence FOC controller. Graphs from left to right and from top to bottom are electrical torque Te, DC voltage Vdc, stator power Ps, and grid VSC power Pconv, respectively.

Under unbalanced grid-voltage conditions, the classical rotor and grid VSC controller has to be tuned very fast to limit torque pulsation and DC link voltage ripple, because otherwise its stability margin will decrease. Compared to the literature on DFIGs during unbalanced grid-voltage disturbance [Bendl, et al.'05, Brekken and Mohan'07, Pena, et al.'07, Seman, et al.'06], this study has made some improvements: Methods to separate the positive and negative sequence signals; Reference current calculation; The use of the generator VSC to limit torque oscillation, and grid VSC to limit DC voltage ripple.

97

4.3 Operation of Low-Inertia Power Grid

4.3 Operation of Low-Inertia Power Grid


Classical power system operation assumes that the grid frequency is mainly related to active power while grid voltage is mainly related to reactive power [Anderson and Fouad'03, Kundur'94]. In the classical power system, where the synchronous generator that is directly connected to the grid dominates, the grid frequency is proportional to the mechanical rotation speed of the synchronous generator. The rotation speed of the synchronous generator is determined by the input mechanical power and output electrical power. Thus the assumption that the grid frequency is mainly related active power is true. However, nowadays installation of wind power and other renewable energy sources is increasing drastically, thus the mechanical speed of generator is completely decoupled from grid electrical frequency by using power electronic converters. Furthermore, the trend of motor-load integration using a power electronic converter attenuates this problem. In the extreme case, a power system may contain a very small or even no synchronous generator in direct connection with AC grid, in which case the grid frequency is not directly related to active power and grid voltage is not directly related to reactive power [Karlsson, et al.'05, Katiraei and Iravani'05, Leva'08, Piagi and Lasseter'06]. An isolated power grid possibly involves very low-inertia, if the power source and load are integrated by power electronic converters. There are two main control strategies to operate this low-inertia power grid: With a central unit and/or master control loop With distributed units and/or a local control loop [Guerrero, et al.'05, '07, Karlsson,

et al.'05, Lasseter and Piagi'04] The first method uses one main power source, namely a master control loop, to control the voltage and frequency of the grid. The advantage is that its control implementation is easy. However, there are several disadvantages. Firstly, the grid operation is not flexible enough to add or remove distributed generation DG units freely. Secondly, the main power source has to have a large enough power capacity, otherwise the system will have to rely on communication between all the units participating in frequency and voltage control. Finally, the control unit has to switch between grid connected mode and islanded mode. A fastislanding detection method is required for the central unit to transfer seamlessly between the two modes. The second method uses multiple power sources to control voltage and frequency with only local signals. As the local grid often has many small DG units, the second method is more realistic. The advantages are that the reliability is high, the units do not require a very large power capacity, and communication is not essential. However, its control

98

CH.4 Grid Integration of Individual Variable Speed Wind Turbine

implementation is not very self-evident. The operation of the main AC power grid is based on the second control strategy without a master controller. All SGs only use local signals, such as their rotation speeds and terminal voltages, to participate in grid frequency and voltage control. Just as in the main power grid, the islanded micro-grid also works sufficiently when the SGs are connected. However, the isolated grid becomes unstable when the number of directly AC-coupled SGs decreases, even though all the DG units are equipped with classical frequency and voltage-droop controllers. The reason is that in such an isolated grid, the total inertia of the system decreases and frequency stability problems arise. In the extreme case, if no synchronous generator is directly connected to the grid, the frequency does not vary with an active power unbalance and the voltage does not vary with a reactive power unbalance. Moreover, the classical droop controller ceases to work. The behaviours of large-inertia and low-inertia power grids during isolation are described in section 4.3.1.

4.3.1 Comparison of Large-Inertia and Low-Inertia Power Grids


A micro-grid model is used to compare the dynamic behaviour of large-inertia and lowinertia power grids, as shown in Figure 4-8. An SG is modelled as the power source of a large-inertia power grid and a VSWT is modelled for the low-inertia grid, both being equipped with droop controllers. A constant power load is modelled by a controlled VSC, and a constant impedance load is modelled by a shunt resistor and inductor. .The simulation scenario is set as Table 4-1. The droop ratio for the two power sources is also the same for the two simulations: Dp = 5% and Dq = 2% , which are typical values for a thermal power plant [Kundur'94]. The additional power produced by the generator as a result of the droop controllers at grid frequency/voltage deviations are calculated as (4.1) and (4.2): (4.1)

(4.2) where and are the grid frequency and voltage deviations, respectively; and and

are the nominal grid frequency and grid voltage, respectively; and are the nominal active and reactive power of the generator, respectively.

99

4.3 Operation of Low-Inertia Power Grid

Figure 4-8. Topology of an investigated micro-grid.

Table 4-1. Simulation scenario of an islanding micro-grid operation.

Time (second) 10 15 20

Event Circuit breaker opens, islanding operation begins Constant active power load decreases Constant reactive power load decreases

Figure 4-9 shows the dynamic simulation results of a large-inertia power grid. Before islanding, the micro-grid sends about 0.5MW and 1MVar to the main grid. At 10s, the switch opens and the micro-grid is isolated from the main grid. Because of the redundant active and reactive power, both frequency and voltage increase. The frequency of the micro-grid is measured by a PLL, as discussed in Chapter 2. Then the frequency droop controller decreases the the generated by the SG, and the voltage droop controller increases absorbed by the SG. At 15s, the constant active power load decreases, which

increases the frequency, after which the P produced by the SG decreases. At 20s, the constant reactive power load increases, which decreases the voltage, after which the absorbed by the SG decreases. All the results agree with the common knowledge of the power system, which are Pload % ! f & and Qload % ! V & . The SG droop controllers help to maintain the frequency and voltage.

100

CH.4 Grid Integration of Individual Variable Speed Wind Turbine

Q Grid(Mvar)

P Grid(MW)

1 0 -1 3 2 1 0 2 1 0 60 50 40 A B C A B 15 20 C 5 10 15 A B C A

2 1 0 -1 2 0 A -2 1 0 -1 3 2 1 5 10 A B 15 20 time s C 25 30 A B C 5 10 15 B 20 C 25 30 5 10 15 20 25 30 A

10

15

20

25

30 Q Gen(Mvar)

P Gen(MW)

20

25

30 Q Load(Mvar)

P Load(MW)

10

25

30 Vll MicroGrid(pu)

10

15

20

25

30

f MicroGrid(Hz)

10

15 20 time s

25

30

Figure 4-9. Islanding operation of a large-inertia power grid with droop controller. A: beginning of islanding, B: load active power decrease, C: reactive load power increase.

However, in low-inertia power grids, as shown in Figure 4-10, unexpected phenomenon appears that do not agree with the common knowledge of the power system. Before islanding, the micro-grid also sends about 0.5MW and 1MVar to the main grid. At 10s, the switch opens and the micro-grid is isolated from main grid. However, in contrast to the large-inertia power grid, frequency decreases and voltage increases. The frequency of the micro-grid is also measured by a PLL. Most importantly, the voltage stabilises at unacceptable values, almost 2 pu. However, it is not possible to limit the voltage by further increasing Dq , as this will cause large oscillations and eventually render the grid unstable. At 15s, the constant active power load decreases, which does not change the frequency but increases the grid voltage. At 20s, the constant reactive power load increases, which decreases the voltage and increases the frequency. The seemingly strange changes in frequency and voltage due to the load changes challenge the islanding operation of this micro-grid. The classical droop controller installed on the power source cannot respond

101

4.3 Operation of Low-Inertia Power Grid

correctly to load change, because of the strange behaviour of frequency and voltage. Actually the classical droop controller in this low-inertia power grid deteriorates the islanding operation.
1 0 -1 3 2 1 0 2 1 0 60 50 40 A B C A 5 10 15 B 20 C 25 30 Vll MicroGrid(pu) 5 10 15 20 25 30 Q Load(Mvar) 1 0 -1 3 2 A 1 5 10 15 20 time s 25 30 B C A B C A B C A 2 1 0 -1 2 0 -2 -4 5 10 A 15 B 20 C 25 30 5 10 15 20 25 30 A

10

15

20

25

30 Q Gen(Mvar)

P Load(MW)

P Gen(MW)

Q Grid(Mvar)

P Grid(MW)

10

15

20

25

30

f MicroGrid(Hz)

10

15 20 time s

25

30

Figure 4-10. Islanding operation of a low-inertia power grid with a droop controller. A: beginning of islanding, B: load active power P decrease, C: load reactive power Q increase.

4.3.2Analysis of the Low-Inertia Power Grid


In section 4.3.1, the strange behaviour of a low-inertia power grid during islanding operation in dynamic simulation is presented in Figure 4-10. In this section, the reason for the strange behaviour will be revealed using a steady-state analysis. In order to simplify the analysis, only constant power load from a power electronic converter is considered. In such an isolated low-inertia power grid, the grid frequency is measured by a PLL, and not by the actual rotation speed of the SG. The dynamic behaviour can be understood by analysing the voltage equation of the shunt capacitance. The capacitor is from the

102

CH.4 Grid Integration of Individual Variable Speed Wind Turbine

converter filter or the parasitic capacitor from the cable and transmission line. Its smallsignal model is (4.3) and (4.4).

1 _ Vd = (id gen id load ) + !e0 Vq C 1 _ Vq = (iq C


gen

(4.3) (4.4)

iq

load )

!e0 Vd

The voltage vector is chosen to align with the q axis, thus at steady state Vd = 0 and

Vq = jV j, P = Vq iq and Q = Vq id . Then, id is the reactive current, and iq is the active


current. And the simplified small-signal model of PLL is:

_ P LL = 0 Vd _ !P LL = P LL = kp (0 Vd ) ki P LL

(4.5) (4.6)

From (4.3), it can be seen that increase in the reactive load current id load will at least transiently decrease the d axis voltage Vd. Then from (4.6) it can be seen that the measured grid frequency !P LL will increase. Similar to (4.4), the decrease of load active current

iq

load

will at least transiently increase the q axis voltage Vq , which is the actually the

magnitude of voltage jV j . To conclude, in a low-inertia power grid, the frequency ! or f is at least transiently related to reactive power load Qload % ! f %, and voltage V is transiently related to active power load Pload % ! V &. This analysis agrees with the simulation results shown in Figure 4-10. The conventional frequency and voltage-droop controllers cannot sufficiently meet the load variation due to the unconventional frequency and voltage responses.

4.3.3 Operation of Low-Inertia Power Grid with Central Control Unit


In sections 4.3.1 and 4.3.2, a low-inertia power grid is analysed. The conclusion is that the classical droop controller of VSWTs or other power sources with power electronic interface cannot meet the load variation sufficiently. This shortcoming is because of the unconventional frequency and voltage behaviour. In sections 4.3.3 and 4.3.4, operation methods of a low-inertia power grid are discussed, of which one uses a central control unit and master loop to control the frequency and voltage of the low-inertia power grid, and the other uses distributed control units. In certain situations where the generations and loads are not electrically far away and the geographic area of the grid is not very large, a central control unit can be used. In this case, the communication between units is not difficult to implement, and the central control unit 103

4.3 Operation of Low-Inertia Power Grid

can effectively control the frequency and voltage of all parts of the grid. An example of such a grid is a wind farm composed of DFIGs that is integrated with the main grid by HVDC thyristor bridges. In such a wind farm, essentially no inertia exists. A STATCOM is installed in the AC wind park to control the frequency and voltage. The DC voltage of the STATCOM is an index for the active power balance between the wind farm production and thyristor bridges transmission. The fire angle of the thyristor rectifier is used to control the DC voltage of the STATCOM, which uses a PI controller. The topology of this wind farm is shown in Figure 4-11. In order to simplify the simulation, only two DFIGs are considered here to investigate the possible interactions between DFIGs inside the wind farm.

Figure 4-11. Topology of a DFIG wind farm with thyristor bridge integration.

The simulation results are shown in Figure 4-12 to Figure 4-13. In Figure 4-12, the wind speed and power of each DFIG is plotted. In Figure 4-14, the power and DC voltage of the STATCOM is plotted. In Figure 4-13, the power and DC voltages of the thyristor bridge are plotted.

104

CH.4 Grid Integration of Individual Variable Speed Wind Turbine

DFIG 01 15 wind (m/s) wind (m/s) 15

DFIG 02

10

10

5 3 Ps (MW)

20

40

5 3 Ps (MW)

20

40

1 0.4 Pconv (MW)

20

40

1 0.4 Pconv (MW)

20

40

0.2

0.2

0 0

20

40

0 0

20

40

slip

-0.1

slip 0 20 time(s) 40

-0.1

-0.2

-0.2

20 time(s)

40

Figure 4-12. Wind speed, active, reactive power and slip of the DFIG.

105

4.3 Operation of Low-Inertia Power Grid

Thyristor Rectifier 15 fire angle ( ) fire angle ( ) 170

Thyristor Inverter

10

165

5 45.5 Vdc(kv)

20

40

160

20

40

45

Vdc(kv) 0 20 40

44.8 44.6 44.4

44.5 5 P (MW)

0 5 P (MW) 4 3 2 1.5 Q (MVar)

20

40

4 3 2

20

40

20

40

1.5 Q (MVar)

0.5

20 time(s)

40

0.5

20 time(s)

40

Figure 4-13. Fire angle, DC voltage, active and reactive power of thyristor bridges.

106

CH.4 Grid Integration of Individual Variable Speed Wind Turbine

STATCOM P,Q 0.5 P,Q (MVA) STATCOM P STATCOM Q 0 1.05

STATCOM Vdc

Vdc (pu)

-0.5

20 time(s)

40

0.95

20 time(s)

40

Figure 4-14. Active, reactive power and DC voltage of the STATCOM.

From the simulation results, it can be seen that although the active power of the DFIGs changes significantly following a change in wind speed, the DC voltage of the STATCOM is virtually constant. The fire angle of the thyristor rectifier follows the change in DC voltage, and all the active power produced by the wind farm is transmitted through the thyristor bridge. Furthermore, a simple feed-forward loop is added onto the DFIGs reactive power control loop so that the DFIG is used to produce reactive power in proportion to its own active power. The reactive power reference depends on the power factor of the thyristor rectifier and equals PDF IG sin , where is the fire angle of the thyristor rectifier. In this way, the reactive power consumed by the thyristor rectifier will not be provided by the STATCOM, but by all the DFIGs instead. In all, the STATCOM is only required to balance the active power transiently and to provide part of the reactive power consumed by the thyristor rectifier. Thus the required power rating of STATCOM is limited. In the above case study, a STATCOM is required to control the DFIG wind farm with thyristor bridge. Control methods without using STATCOM are also studied. Two new controllers are developed through that case study, namely self terminal voltage controller and self flux controller[Zhou, et al.'06]. Both controllers treat the thyristor bridge as a constant impedance load, where the fire angle of rectifier of thyristor bridge is kept as constant. Both new controllers do not need energy storage device - STATCOM in the wind farm, and can control the active power transferred through the thyristor bridge. However for the new controllers, reactive power flow between DFIGs is a concern. Additional reactive power controller is designed to limit the unwanted reactive power flow. Both methods still have transient problems, which need further studies:

107

4.3 Operation of Low-Inertia Power Grid

Reactive power flow between DFIGs can only be controlled slowly, otherwise the active power flow will be influenced; Low frequency oscillations exist inside the wind farm and are difficult to be damped out.

4.3.4Operation of Low-Inertia Power Grid with Distributed Control Unit


In section 4.3.3, a special low-inertia power grid (DFIG) wind farm with thyristor bridge integration is analysed which can easily use a central control unit to control the frequency and voltage. However, in many situations it is not possible or realistic to use only one central control unit. The nature of distributed generators and loads requires distributed control units. A new type of controller is developed for this consideration, which is called the inverse frequency and voltage droop controller[Engler'04, Engler and Osika'03, Engler and Soultanis'05, Lasseter and Piagi'04]. The name inverse frequency and voltage droop controller comes from the control strategy of the grid VSC. The grid VSC measures its output active power and reactive power, and controls its output AC voltage magnitude and frequency according to its active and reactive power, explained as Figure 4-15. The grid-side VSC controls the AC grid frequency and voltage actively, and the unit side VSC controls the DC voltage. The new controller emulates the behaviour of a synchronous generator, thus actively contributing to frequency and voltage stability. The benefits of this controller are: (1) it does not require mode switching between the grid connection and islanding operation, thus islanding detection and resynchronization are not required; and (2) it does not require a master system controller and it only uses local signals, thus allowing more DG units to easily be added into the micro-grid. The concept of the inverse droop controller is shown in Figure 4-15. The droops are based on equations (4.7) and (4.8):

f = f f0 = Dp (P P0 ) V = V V0 = Dq (Q Q0 )

(4.7) (4.8)

The rate limiter used in the inverse frequency droop emulates the inertia constant of a synchronous generator. The rate limiter is based on the swing equation (4.9). Generators with larger inertia constants have a great amount of kinetic energy stored on the shaft and therefore they can be used to meet larger load variations transiently.

d! 1 = (Pmech Pelec ) dt 2H

(4.9)

108

CH.4 Grid Integration of Individual Variable Speed Wind Turbine

A low-inertia micro-grid with multiple-power electronic-interfaced generators and load is simulated, as shown Figure 4-16.

1 1 + Tf S

1 S

_ f =

P 22H

1 1 + Tf S

|V |

Figure 4-15. Topology of inverse droop controller.

Figure 4-16. Low-inertia power grid containing multiple inverse droop-controlled VSCs.

The general operation of the VSC-interfaced generator is also illustrated in Figure 4-16, where the grid VSC is controlled in inverse frequency and voltage droop mode, and the generator VSC controls the DC link voltage and generator terminal AC voltage. For wind

109

4.3 Operation of Low-Inertia Power Grid

turbine applications such as Type D wind turbines, the grid VSC active power P0 is calculated from the power-speed curve of the wind turbine. Additionally, if wind turbines dominate the micro grid, in order to ensure power balance during islanded operation, the wind turbine has to be operated in de-loaded mode, during which its active power output is decreased from its level at maximum available power. During grid-connected operation, because grid frequency f = f0 , the active power produced from the inverse frequency droop-controlled grid VSC equals its reference power P = P0. During islanded operation, the active power output is determined by the grid frequency deviation, the droop ratio, and the emulated inertia of the inverse frequency droop control, as shown in Figure 4-15. The grid VSC of Generator 1 emulates a SG with the inertia constant H = 0:25, and the grid VSC of generator 2 emulates a SG with the inertia constant H = 0:5. The droop ratios are Dp = 5% and Dq = 2%. The simulation scenario is set up as indicated in Table 4-2; the simulation results are shown in Figure 4-17.
Table 4-2. Simulation scenario of islanding operation of a micro-grid.

Time (seconds) 10 15 20

Event Circuit breaker opens, islanding operation begins Constant active power load increases Constant reactive power load increases

Figure 4-17 shows the dynamic simulation results of a low-inertia power grid with multiple inverse droop controlled power sources. Before islanding, the micro-grid sends about 0.5MW and 1MVar to the main grid. At 10s, the switch opens and the micro-grid is isolated from the main grid, the frequency and voltage of which increase and stabilise at adequate values. At 15s, the constant active power load increases, which decreases the frequency. At 20s, the constant reactive power load increases, which decreases the voltage. All the changes in frequency and voltage coincide with the behaviour of a large-inertia power grid. Moreover, the changes in generation correctly follow the changes in load. Another important feature is that with the different H (emulated inertia constants) in the grid VSC, two power sources provide different active power support during transients. Generator 2, which emulates a larger inertia SG, temporarily provides more active power than generator 1 does with low inertia. This implies that the inverse droop-controlled VSWT can make use of the turbine inertia, as long as the rate limiter is set in accordance with the turbines inertia constant. The active and reactive power produced by grid VSCs are finally stabilised at values according to their droop ratio. However, the over-current

110

CH.4 Grid Integration of Individual Variable Speed Wind Turbine

protection of inverse droop-controlled grid VSCs is not considered here but is recommended for future research.
Q Grid(Mvar) 1 P Grid(MW) A 0 -1 P Gen1(MW) 2 1 0 2 1 0 2 P Load(MW) 1 0 51 50 49 A B C A B C A B C A B C 2 1 0 -1 2 0 -2 2 0 -2 2 0 -2 1.1 1 0.9 A B C A B C 5 10 15 20 25 30 A

10

15

20

25

30 Q Gen1(Mvar)

10

15

20

25

30 Q Gen2(Mvar)

10

15

20

25

30

P Gen2(MW)

10

15

20

25

30 Q Load(Mvar)

10

15

20

25

30

10

15

20

25

30 Vll MicroGrid(pu)

10

15

20

25

30

f MicroGrid(Hz)

10

15 20 time s

25

30

10

15 20 time s

25

30

Figure 4-17. Islanding operation of low-inertia power grid with an inverse droop-controlled VSC. A: beginning of islanding, B: load active power P increase, C: reactive load power Q increase.

4.4 Conclusions
In this chapter, several control problems of VSWTs are discussed. Firstly, the dualsequence current controller is developed to control VSWT during unbalanced grid voltage. Simulation of a DFIG during an unbalanced grid-voltage dip shows that both the electric torque oscillation and DC voltage ripple can be limited satisfactorily by the dual-sequence

111

4.4 Conclusions

current controller. Finally, the problems of low-inertia power grids are studied, for which two control methods are proposed. One method is to operate the grid with a central unit. An example for this method is a wind farm composed of DFIGs and integrated with main power grid by HVDC thyristor brdige. A STATCOM is installed inside the wind farm and performes as the central unit. A PI controller is used to control the DC voltage of the STATCOM by changing the fire angle of the thyristor rectifier. The other method uses distributed units with inverse droop-controllers to control the frequency and voltage. An example of this method is a micro-grid composed of power electronics interfaced DGs and static load. Simulation results show that both methods work efficiently: the generations and loads are balanced, and frequencies and voltages are controlled at acceptable values. The over-current protection of inverse droop controllers is recommended for future research.

112

CH.4 Grid Integration of Individual Variable Speed Wind Turbine

4.5 References
[Aka84] H. Akagi, Y. Kanazawa, and A. Nabae, "Instantaneous Reactive Power Compensators Comprising Switching Devices without Energy Storage Components," IEEE Transactions on Industry Applications, Page(s). 625-630, 1984. V. Akhmatov, Analysis of dynamic behavior of electric power systems with large amount of wind power, PhD thesis, Electrical Power Engineering Department, Technical University of Denmark, Copenhagen, Denmark, 2003. V. Akhmatov, "Modelling and ride-through capability of variable speed wind turbines with permanent magnet generators," Wind Energy, vol. 9, issue 4, 2005. P. M. Anderson and A. A. Fouad, Power system control and stability, ISBN0-471-23862-7 Wiley-Interscience, New York, 2003. J. Bendl, M. Chomat, and L. Schreier, "Independent control of positive- and negative-sequence current components in a doubly fed machine," European Transactions on Electrical Power, vol. 15, issue 3, Page(s). 191-202, 2005. T. Brekken and N. Mohan, "Control of a Doubly Fed Induction Wind Generator Under Unbalanced Grid Voltage Conditions," IEEE Transaction on Energy Conversion, vol. 22, issue 1, Page(s). 129-135, 2007. C. Chompoo-inwai, C. Yingvivatanapong, K. Methaprayoon, and W. J. Lee, "Reactive compensation techniques to improve the ride-through capability of wind turbine during disturbance," IEEE Transactions on Industry Applications, vol. 41, issue 3, Page(s). 666-672, 2005. A. Engler, "Device for equal-rated parallel operation of single-or three-phase voltage sources," US Patent 6,693,809, 2004. A. Engler and M. S. O. Osika, "Simulation of Inverter Dominated Minigrids," in 2nd European PV-Hybrid and Mini-Grid Conference, Kassel, Germany, 2003. A. Engler and N. Soultanis, "Droop control in LV-Grids," in 2005 International Conference on Future Power Systems, 2005, Page(s). 1-6. J. M. Guerrero, J. Matas, L. G. de Vicuna, N. Berbel, and J. Sosa, "Wirelesscontrol strategy for parallel operation of distributed generation inverters," in IEEE International Symposium on Industrial Electronics, 2005. ISIE 2005, Stockholm, Sweden, 2005. J. M. Guerrero, J. Matas, L. G. de Vicuna, N. Berbel, and J. Sosa, "Decentralized Control for Parallel Operation of Distributed Generation Inverters Using Resistive Output Impedance," IEEE Transactions on Industrial Electronics, vol. 54, issue 2, Page(s). 994-1004, 2007. P. Karlsson, J. Bjornstedt, and M. Strom, "Stability of Voltage and Frequency Control in Distributed Generation Based on Parallel-Connected Converters Feeding Constant Power Loads," in 2005 European Conference on Power Electronics and Applications, EPE2005, 2005, Page(s). 10.

[Akh03]

[Akh05]

[And03] [Ben05]

[Bre07]

[Cho05]

[Eng04] [Eng03]

[Eng05] [Gue05]

[Gue07]

[Kar05]

113

4.5 References

[Kat05]

[Kun94] [Las04]

[Lev08]

[Lyo54]

[Mol08]

[Mor05]

[Nab96]

[Paa00]

[Pen07]

[Pet05]

[Pet05]

[Pia06] [Sac01]

[Sem06]

F. Katiraei and M. Iravani, "Transients of a Micro-Grid system with multiple distributed energy resources," International conference on Power System Transients, IPST05, 2005. P. Kundur, Power System Stability and Control, ISBN-0-07-035958-x, McGraw-Hill Professional, 1994. R. H. Lasseter and P. Piagi, "Microgrid: A Conceptual Solution," in 35th Annual IEEE Power Electronics Specialists Conference, PESC04, Aachen, Germany, 2004. S. Leva, "Dynamic Stability of Isolated System in the Presence of PQ Disturbances," IEEE Transactions on Power Delivery, vol. 23, issue 2, Page(s). 831-840, 2008. W. V. Lyon, Transient Analysis of Alternating Current Machinery. An Application of the Methods of Symmetrical Components., Cambridge, and John Wiley, 1954. M. Molinas, J. A. Suul, and T. Undeland, "Low voltage ride through of wind farms with cage generators: STATCOM versus SVC," IEEE Transactions on Power Electronics, vol. 23, issue 3, Page(s). 1104-1117, 2008. J. Morren and S. W. H. de Haan, "Ride through of wind turbines with doubly-fed induction generator during a voltage dip," IEEE Transactions on Energy Conversion, vol. 20, issue 2, Page(s). 435-441, 2005. A. Nabae and T. Tanaka, "A new definition of instantaneous active-reactive current and powerbased on instantaneous space vectors on polar coordinates in three-phasecircuits," IEEE Transactions on Power Delivery, vol. 11, issue 3, Page(s). 1238-1243, 1996. G. C. Paap, "Symmetrical components in the time domain and their application topower network calculations," IEEE Transactions on Power Systems, vol. 15, issue 2, Page(s). 522-528, 2000. R. Pena, R. Crdenas, E. Escobar, J. Clare, and P. Wheeler, "Control System for Unbalanced Operation of Stand-Alone Doubly Fed Induction Generators," IEEE Transactions on Energy Conversion, vol. 22, issue 2, Page(s). 544-545, 2007. A. Petersson, Analysis, modeling and control of doubly-fed induction generators for wind turbines, Energy and Environment Department, Chalmers University, 2005. A. Petersson, S. Lundberg, and T. Thiringer, "A DFIG wind turbine ridethrough system. Influence on the energy production," Wind Energy, vol. 8, issue 3, 2005. P. Piagi and R. H. Lasseter, "Autonomous control of microgrids," in IEEE Power Engineering Society General Meeting, Montreal, Canada, 2006. G. Saccomando and J. Svensson, "Transient operation of grid-connected voltage source converterunder unbalanced voltage conditions," in Conference Record of the 2001 IEEE Industry Applications Conference, 36th IAS Annual Meeting. , 2001. S. Seman, J. Niiranen, and A. Arkkio, "Ride-Through Analysis of Doubly Fed Induction Wind-Power Generator Under Unsymmetrical Network

114

CH.4 Grid Integration of Individual Variable Speed Wind Turbine

[Xia06]

Disturbance," IEEE Transactions on Power Systems, vol. 21, issue 4, Page(s). 1782, 2006. D. Xiang, L. Ran, P. J. Tavner, and S. Yang, "Control of a doubly fed induction generator in a wind turbine during grid fault ride-through," IEEE Transaction on Energy Conversion, vol. 21, issue 3, Page(s). 652662, 2006.

115

4.5 References

116

CH.5 Wind Park Conceptual Design and Control

Chapter 5
Wind Park Conceptual Design and Control 5.1 Introduction
Wind turbine technology has evolved from CSWTs to VSWTs. This development has mainly been caused by the fact that VSWTs have better power quality and low-voltage ride through (LVRT) capability [Akhmatov'05, Blaabjerg and Chen'06, Bollen, et al.'05]. The fact that VSWTs are able to increase energy production is not deterministic for the evolution of wind turbine technology, because the improvement in the energy production of VSWTs is very small [Hoffmann'02]. CSWTs still have some advantages, such as being low-cost and having a simple design which requires less maintenance. A hybrid wind park composed of both CSWTs and VSWTs is proposed in this thesis, as this combination can provide similar power quality and LVRT capability with better cost benefits. VSWTs in a wind park work as distributed STATCOMs, using reactive power to limit flicker and improve LVRT capability. A multi-objective optimisation method is used to determine the optimal configuration of a hybrid wind park which has low levelised production cost, high power quality, high reliability and high LVRT capability. The wind park has to be connected to the main power grid. The possible integration options are an HVAC system, a classical HVDC system using a line-commutated converter a thyristor bridge (HVDC LCC), and an HVDC system using a self-commutated converter - an IGBT bridge (HVDC VSC). The ultimate choice is made by comparing the transmission costs of the three systems, including both investment and energy loss. In order to be integrated successfully into a power grid, the hybrid wind park has to be controlled so as to meet the grid-code requirements in a flexible way. For example, the reactive power capability of variable speed wind turbines is used to control voltage at points of common coupling to limit the flicker level and to improve LVRT capability of CSWTs in the wind park. A central controller of the hybrid wind park using an antiwindup PI controller and distribution function is designed in this thesis. In this chapter, the design and control aspects of a wind park are discussed. Firstly, the integration choices between an HVAC, HVDC LCC and HVDC VSC in a wind park are compared, then a multi-objective optimisation method is used to optimise a hybrid wind

117

5.2 Wind Park Integration Choice

park, and finally a central controller of the hybrid wind park to control the wind park is proposed as an integral wind power plant.

5.2 Wind Park Integration Choice


The three integration choices, namely HVAC, HVDC LCC, and HVDC VSC, are compared with regard to their transmission costs. The total transmission costs include investment and energy loss. When the power output of a wind park reaches beyond its transmission capability, some energy has to be spilled. Here, the spilled energy is included in the total energy loss. Thus in order to calculate the total transmission costs, the transmission capabilities and power loss of the three integration choices have to be calculated first. In [Kundur'94] and [Dunlop, et al.'79], the transmission capability of HVAC overhead lines is analysed using its distributed transmission line model. However, the offshore wind parks use submarine cables, which have different transmission capability from the HVAC overhead lines. The power loss of HVAC cable system can be found in [Brakelmann'03, '03, Todorovic'04] and loss of HVDC system can be found in [Negra'05, Zhao'08]. A detailed comparison of different integration choices is found in [Lazaridis'05], where it is concluded that HVDC LCCs are the most economically efficient due to omitting the very expensive offshore platforms. Some general conclusions can also be found in [Ackermann'05, Wright, et al.'02]. In this thesis, the method used to compare the three integration choices is presented through case studies. A Matlab program is developed to implement the comparison method, including the transmission capability calculation of an HVAC cable system, and the power loss calculation of HVAC and HVDC systems. In this comparison, the influences on the transmission capability from different aspects such as reactive power compensation, the short-circuit ratio at the point of common coupling are evaluated; moreover, the cost of using expensive offshore platforms is taken into account. From [Ackermann'05, Lazaridis'05, Wright, et al.'02], it can be seen that the total transmission cost of a wind park is related to many aspects, such as power rating, voltage rating, and also the transmission distance. A complete database should include the parameters and costs of all components with different power ratings and different voltage ratings.It is difficult to make a very thorough comparison of the three integration choices due to the lack of a complete database. However, it is possible to simplify the comparison according to the observation of results from [Ackermann'05, Lazaridis'05, Wright, et al.'02]. The transmission cost of the HVAC cable solution is sensitive to the transmission distance, due to the large reactive power produced by the shunt capacitor. Transmission capability decreases and power loss increases drastically with the increase in cable length. The

118

CH.5 Wind Park Conceptual Design and Control

HVDC VSC and HVDC LCC are both less sensitive to transmission distance. HVDC VSC is assumed to be sensitive to the power rating of the wind park, because the existing maximum converter is 550MW. The HVDC LCC is less sensitive to the power ratings, as the maximum converter reaches 6000MW. Because of the above reasons and the unavailability of a complete components database, the comparisons are simplified and only two cases are studied in this thesis. One compares an HVAC cable system and an HVDC VSC system of a 300MW wind park at all transmission distances, and the other compares an HVDC VSC and an HVDC LCC system with a fixed transmission distance and varying wind park power ratings. These two case studies are summarised in Table 5-1. The two case studies do not cover all the power ratings, voltage levels and transmission distances. However with a complete components database, the method presented in this thesis can be used to analyse the transmission costs of wind parks at different transmission distances with different voltage levels.
Table 5-1. Case studies of integration choices. Description case 1 case 2 HVAC cable and HVDC VSC compared HVDC VSC and HVDC LCC compared power rating (MW) 300 300 - 1200 transmission distance (km) 50 250 150

0.1 probability

0.05

0 wind farm produced power (MW)

10

15 wind speed (m/s)

20

25

30

300 200 100 0

10

15 wind speed (m/s)

20

25

30

Figure 5-1. Wind park characteristics. (a) probability of wind speeds at the wind park, (b) wind park power production at different wind speeds.

In Figure 5-1, the wind speed in Weibull distribution and power production of a 300MW wind park are plotted. In case study 2 when comparing two HVDC systems, the power 119

5.2 Wind Park Integration Choice

production of the wind park is scaled up until its power rating reaches 1200MW. The AC and DC cables data used in these case studies are listed in Table 5-2 [Lazaridis'05, Negra, et al.'06]. Parameters and costs of other components can be found in [Lazaridis'05, Lundberg'03, Pierik, et al.'01].
Table 5-2. AC and DC cable parameters. 220kv AC cable 380

150kv DC cable
350

150kv DC cable
550

rated power (MW ) resistance (=km) inductance (H=km) capacitance (F=km) nominal current (kA) cable price (MEuro=km) installing cost (MEuro )

4:8 103 3:7 103 1:8 107


1.055 1.65 0.1

1:38 102

9 103

1.3 0.4453 0.2

1.677 0.6086 0.2

5.2.1 HVAC Cable Transmission Capability


In [Kundur'94] and [Dunlop, et al.'79], a St. Clair curve is used to determine the transmission capability of an HVAC overhead line. The cables transmission capability is very different from the overhead line due to its large parasitic capacitance. In this section, the cable transmission capability is calculated from its distributed transmission model, similarly to [Dunlop, et al.'79, Kundur'94]. Furthermore, a more realistic (i.e. less theoretically-based) case is considered here, where the output power from the wind park and the voltage at the PCC are known. This case usually requires iterative power-flow calculation. Here a method which eliminates the iterative process is demonstrated. Additionally, this method also shows all the operation points of an HVAC cable with a given length. It is found that HVAC cable also has an inherent maximum-allowed transmission power at a given length, as shown in Figure 5-3, which is similar to the HVAC overhead line from [Kundur'94]. However, unlike the HVAC overhead line, the transmission capability of HVAC cables is determined by the thermal limitation, which is much smaller than its inherent maximum value.

5.2.1.1 Transmission Model


The transmission model of an HVAC cable system, including a shunt inductor, transformer and short-circuit impedance at PCC, is shown as Figure 5-2, where Zt1 , Zt2 are impedances of offshore and onshore transformers, respectively; YL1, YL2 are impedances of offshore and onshore shunt inductors, respectively; and Zsc is the short-circuit impedance at PCC.

120

CH.5 Wind Park Conceptual Design and Control

Figure 5-2. Model of HVAC cable transmission system.

The cable has a distributed transmission model, the equations for which can be found in [Brakelmann'03, '03, Kundur'94]. The influence of offshore and onshore transformers, shunt inductors, etc., are considered in this thesis. The complete system is represented by an ABCDtot equivalent model, which is calculated from (5.1):

Atot = (1 + Zt1 ) A + Zt1 YL2 B Btot = [Zt2 + Zsc + Zt1 YL1 (Zt2 + Zsc )] A + [YL2 (Zt2 + Zsc ) + Zt1 YL1 YL2 (Zt2 + Zsc ) + Zt1 YL1 + 1] B + Zt1 (Zt2 + Zsc ) C + [Zt1 YL2 (Zt2 + Zsc ) + Zt1 ] D Ctot = YL1 A + YL1 YL2 B + C + YL2 D Dtot = (Zt2 + Zsc ) YL1 A + [(Zt2 + Zsc ) YL1 YL2 + YL1 ] B + (Zt2 + Zsc ) C + [1 + (Zt2 + Zsc ) YL2 ] D

(5.1)

The parameters A, B , C , D , Zt1, Zt2, YL1, YL2, and Zsc in (5.1) are explained in Figure 5-2. The known variables of the equivalent ABCDtot model (5.1) are the sending-end active and reactive power (i.e., the wind park power Pwp ; Qwp ), and the receiving-end

~ ~ voltage Vgrid. The unknown variables are the sending-end voltage Vwp , the sending-end ~ ~ current Iwp , and the receiving-end current Igrid . The relations between known and
unknown variables are represented by the following equations from (5.2) to (5.4).

~ ~ ~ Vwp = Atot Vgrid + Btot Igrid ~ ~ ~ Iwp = Ctot Vgrid + Dtot Igrid ~ ~ Pwp j Qwp = conj(Vwp ) Iwp

(5.2) (5.3) (5.4)

121

5.2 Wind Park Integration Choice

After combining the equations from (5.1) to (5.4), the case of a wind park integrated into the main grid through an HVAC cable system can be solved.

5.2.1.2 Transmission Capability


1.2 sending end voltage Vs (pu) 1 0.8 0.6 0.4 0.2 0 0
Realistic values Unrealistic values

sending end voltage phase angle (deg)

150

Realistic values Unrealistic values

250km 200km 150km

100km

100

50

250km 200km 150km

100km

500 1000 1500 2000 2500 3000 3500 sending end power Ps (Mw) (a)

500 1000 1500 2000 2500 3000 3500 sending end power Ps (Mw) (b)

2 Maximum cable voltage (pu) 1.15 1.1 1.05 1 0.95 250km 200km 150km 100km Maximum cable current (pu)

1.5 250km 200km 0.5 150km 100km

500 1000 1500 2000 2500 3000 3500 sending end power Ps (Mw) (c)

200 400 600 800 sending end power Ps (Mw) (d)

1000

Figure 5-3. Characteristics of an HVAC 220kv cable with two end shunt inductors at the transmission distances: 100km, 150km, 200km and 250km. (a) sending-end (wind park) voltage magnitude; (b) sending-end (wind park) voltage angle; (c) maximum cable voltage magnitude (d) maximum cable current magnitude.

In Figure 5-3, the voltage-power and current-power characteristics of a 220kv HVAC cable are plotted from the solutions of equations (5.1) to (5.4). Reactive power is compensated at both ends of the cable by the shunt inductors. As in [Kundur'94], there are two solutions for each sending-end power, because the voltage and current equations are quadratic equations. One group consists of realistic solutions, which are shown as solid curves in Figure 5-3, where both voltages and currents are within a realistic range. The other group consists of unrealistic solutions and unstable operation points, which are shown as dotted curves in Figure 5-3. In the unrealistic solutions, the voltage magnitudes are too small and voltage angles are too large. The low voltage implies high cable current. The large angle

122

CH.5 Wind Park Conceptual Design and Control

difference between the cables sending and receiving end causes transient stability problems. The results observed in Figure 5-3 are summarised here: The voltage swing from no load to full load at the sending end is below 10%; The phase angle difference between the sending end and receiving end is below 50; The maximum cable voltage magnitude (in the middle of cable) is below 1.1pu; The thermal limitation of the cable is the critical limitation.

Thus the transmission capability of AC cable is only determined by its thermal limitation, i.e., the maximum-allowed cable current. The other limitations, such as maximum-allowed voltage magnitude and angle difference, are not critical. The transmission capability of a 220kv HVAC cable is shown in Figure 5-4. The influences of onshore and offshore transformers, influences of onshore and offshore shunt inductors, and influences of the short-circuit power limitation at the PCC are included.
450
cable + Onshore and Offshore Inductor cable + trafo + Onshore and Offshore Inductor

400

cable + trafo + Onshore Inductor cable + trafo + Onshore and Offshore Inductor & SC=3000MVA simple lumped model

maximum allowed sending end power (MW)

350

300

250

200

150

100

50

50

100

150 200 transmission distance (km)

250

300

350

Figure 5-4. Transmission capability of a 220kv AC cable with reactive power compensation, determined by the maximum-allowed cable current.

123

5.2 Wind Park Integration Choice

The results are summarised here: The transformers and limited short-circuit power at the PCC decrease the cables transmission capability; One-end reactive power compensation (onshore shunt inductor only) drastically decreases the cables transmission capability to nearly half, compared with the situation of double-end reactive power compensation; The simple lumped model cannot be used to estimate the HVAC cables transmission capability.

5.2.2 Offshore Platform


The cost of an offshore platform is the most difficult data to obtain. The experience with offshore wind farms using an HVDC VSC connection is limited and there is no existing offshore wind farm with an HVDC LCC connection at this moment. In the comparison studies of integration methods in [Lazaridis'05, Negra, et al.'06], the costs of an offshore platform are omitted, and it is concluded that the HVDC LCC connection is always more economical than HVDC VSC connection. However, the footprint of the offshore platform of HVDC LCC is much larger than that of HVDC VSC, and an economic comparison between HVDC LCC and HVDC VSC will be strongly influenced by the costs of the offshore platforms. The cost of an offshore platform is influenced by several factors: Platform area, which is the footprint of the offshore station; Weight to support; Site data, including water depth, maximum water level, and water current.

In this section, the footprints of offshore platforms for different connections are taken from available online data, which is listed as [ABB, Siemens, Wensky, et al.'06].
Table 5-3. Footprint of an offshore station. Type HVAC HVDC VSC HVDC LCC Size W*L (m) 10*10 30*40 40*100 Weight to support (tons) 700 2000 8000

The location for the offshore wind farm for this study is in the North Sea, where two existing ongoing offshore wind farm projects are planned: Q7 and Mangrove. The site data for these two offshore wind projects are listed in Table 5-44.

124

CH.5 Wind Park Conceptual Design and Control

Table 5-4. Site data for Q7 and Mangrove offshore wind farms. Offshore distance (km) 30 100 Water depth (m) 30 30 100 year max wave height (m) 14.32 15.60 Design wave period (s) 9.83 12.1 Max water level (m+LAT) 4.05 4.00 Current (m/s) 0.90 1.00

From Table 5-44, it can be seen that the site data in the North Sea for the 30 km offshore distance is similar as that for 100km. Thus the costs of an offshore platform here will mainly vary with footprint and weight to support, which are determined by the integration choice. Final results for the total support structure mass are listed in Table 5-5.
Table 5-5. Costs of an offshore platform for different integration choices. Integration choice HVAC HVDC VSC HVDC LCC Steel price (/kg) 4-6 4-6 4-6 Support structure mass (ton) 350 4000 15000 Low price (M) 1.4 16 60 High price (M) 2.1 24 90

5.2.3 Case Study 1 Comparison of HVAC cable and HVDC VSC


The transmission costs of a 300MW wind park using HVAC 220kv cable and a 150kv HVDC VSC system at transmission distances from 50 to 270 km are calculated and compared. The interest rate of the transmission investment is assumed to be 3%, and the life time is 30 years. The total transmission costs include the following parts: Investment in the wind park transmission system, cable, transformer, shunt inductor, converter, etc. [Lazaridis'05, Lundberg'03, Pierik, et al.'01]; Total cost of energy loss, including spilled energy when the wind park production is larger than its transmission capability, and the energy loss in the transformer,

cable, converter, etc. [Zhao'08]. The levelised transmission costs of a 300MW wind park using an HVAC-cable and an HVDC-VSC transmission system are plotted in Figure 5-5, where situations with and without the costs of an offshore platform are also plotted. As the price of an offshore platform for HVAC transmission is low, its influence on the levelised transmission cost (LTC) is very limited. From Figure 5-5, it can be seen that the LTCs of an HVAC cable transmission system are almost the same with and without considering the costs of an offshore platform. However, the LTCs of an HVDC-VSC transmission are clearly different with and without offshore platforms. The critical distance is 85 km without an offshore platform, and 105 km when taking into account an offshore platform. Above the critical 125

5.2 Wind Park Integration Choice

distance, the HVAC cable systems AC energy loss increases fast and its total transmission cost becomes larger than the HVDC VSC. Thus the HVDC VSC becomes more economical than HVAC cable system above the critical distance 105 km.
compare levelised transmission cost of HVAC-cable and HVDC-VSC

0.06
HVAC-cable including Offshore Platform levelised transmission cost (euro/kwh)

0.05 0.04 0.03 0.02

HVDC-VSC including Offshore Platform HVAC-cable excluding Offshore Platform HVDC-VSC exclude Offshore Platform

105km include offshore platform

0.01
85km exclude offshore platform

0 50

100

150

200

250

300

transmission distance (km) Figure 5-5. Comparison of the levelised transmission costs of a wind park using an HVAC 220kv cable and an HVDC-VSC system at a fixed 300MW wind park power rating, while the transmission distance increases from 50 km to 300 km.

5.2.4 Case Study 2 Comparison of HVDC VSC and HVDC LCC


The transmission costs of a wind park using a 150kv HVDC-VSC and a 150kv HVDC-LCC are calculated and compared. The transmission distance is kept constant at 150km, while the power rating of the wind park increases from 300MW to 1300MW. The interest rate of the transmission investment is assumed to be 3%, and the life time is 30 years. The total transmission costs include the following parts: Investment in the wind park transmission system, cable, transformer, converter, reactive power compensation, etcs [Lazaridis'05, Lundberg'03, Pierik, et al.'01]; Total cost of energy loss in transformer, cable, converter, etc. [Zhao'08].

126

CH.5 Wind Park Conceptual Design and Control

compare levelised transmission cost of HVDC-VSC and HVDC-LCC 10


HVDC VSC including offshore platform HVDC LCC including offshore platform

transmission cost (euro/Mwh)

HVDC VSC excluding offshore platform HVDC LCC excluding offshore platform

Critical pow er rating of w ind park 450MW

5 200

400

600 800 1000 wind farm rated power (MW)

1200

1400

Figure 5-6. Comparison of the levelised transmission costs of a wind park using an HVDC-VSC system and an HVDC-LCC system at a fixed transmission distance of 150km, while the power rating increasing from 300MW to 1300MW.

The levelised transmission costs of an HVDC-VSC and an HVDC-LCC are plotted in Figure 5-6, where situations with and without considering offshore platforms are compared. The curve of the levelised transmission costs is not continuous because the power rating of the HVDC cables is discontinuous in this case study. For HVDC-VSCs, the cases considered when increasing the wind park power rating are: For wind park power rating < 350MW, only one 350MW converter station and one pair of 350MW HVDC cables are used; For 350MW < wind park power rating < 550MW, only one 550MW converter station is used, and there are three possible choices of HVDC cables: one pair of 350MW HVDC cables, two pairs of 350MW HVDC cables, and one pair of 550MW HVDC cables; For 550MW < wind park power rating < 1100MW, two 550MW converter stations are used, and there are two possible choices of HVDC cables for each

127

5.3 Hybrid Wind Park Optimisation

converter station: two pairs of 350MW HVDC cables, and one pair of 550MW HVDC cables; For wind park power rating > 1100MW, three 550MW converter stations have to be used, and there are two possible choices of HVDC cables for each converter station: two pairs of 350MW HVDC cables, and one pair of 550MW HVDC cables. For HVDC-LCCs, the cases considered when increasing the wind park power rating are simpler, because the maximum power rating of HVDC-LCC is not limited: For all situations, only one converter station is used; The choices of HVDC cable are: one pair of 350MW HVDC cables, two pairs of 350MW HVDC cables, three pairs of 350MW HVDC cables, one pair of 550MW HVDC cables, two pairs of 550MW HVDC cables, and three pairs of 550MW HVDC cables. From Figure 5-6, it can be seen that without considering offshore platform, the HVDCLCC is always more economical than the HVDC-VSC. This conclusion is same as that of [Lazaridis'05, Negra, et al.'06]. However, when taking the offshore platform into account, the critical power rating of the wind park is found to be 450MW. Only beyond this power rating, the HVDC-LCC becomes more economical than the HVDC-LCC. It must be pointed out that the HVDC-VSC has a better electrical performance compared to the HVDC-LCC, such as better LVRT capability, better controllability, black start capability. These aspects are not addressed in this comparison.

5.3 Hybrid Wind Park Optimisation


The wind turbine technology evolved from CSWTs to VSWTs. This development was mainly caused by the fact that VSWTs have better power quality and LVRT capability [Akhmatov'05, Blaabjerg and Chen'06, Bollen, et al.'05]. It is true that VSWTs are able to increase energy production because the rotation speed can be optimally controlled for a larger range than CSWTs. However, the improvement in the energy production of VSWTs is small [Hoffmann'02]. Despite the above-mentioned disadvantages, the CSWT still has some advantages which are highlighted in the offshore wind energy, such as low costs, and having a robust construction which requires less maintenance. The advantages and disadvantages of CSWTs compared to VSWTs are summarised in the following table.

128

CH.5 Wind Park Conceptual Design and Control

Table 5-6. CSWT vs. VSWT [Akhmatov'03, Blaabjerg and Chen'06]. Cost + Maintenance + Flicker + LVRT Capability +

CSWT VSWT + positive effect - negative effect

The disadvantage of CSWTs can be removed by proper control of reactive power flow between the wind park and the power grid with additional reactive power compensation equipment [Larsson, et al., Marei, et al.'03, Sun, et al.'04]. However, the mechanical switched shunt capacitor and/or inductor banks are not appropriate. Firstly, the shunt capacitor cannot provide reactive power smoothly. Secondly, it has to be switched on and off very frequently due to the frequent changes in the output power of CSWTs. Finally, its reactive power compensation is poor during grid disturbance because its reactive power is in proportion to V 2 . An offshore wind park composed of CSWTs requires a SVC or STATCOM to be installed at the PCC. Large wind park capacity requires one or more large-sized SVC or STATCOM, which are both expensive. A hybrid wind park composed of both CSWTs and VSWTs seems more promising, as it can provide similar power quality and LVRT capability at a better cost. VSWTs can control active and reactive power independently. Thus the reactive power can be used to limit flicker and improve LVRT capability. In this section, the economic features and electrical performances of the hybrid wind park are investigated. The objective is to determine an optimal hybrid wind park configuration which has low levelised production cost (LPC), high power quality, high reliability and high LVRT capability.

5.3.1 Optimisation Method


The objective of optimisation is to reach a hybrid wind park configuration which has low LPC, high power quality, high reliability and high LVRT capability. In this optimisation problem, the complete design objectives are: Small steady-state voltage deviations at the PCC Low investment Low energy loss High energy yield High reliability of wind park

129

5.3 Hybrid Wind Park Optimisation

Low flicker High LVRT capability Type and number of different wind turbine concepts, including Type A, Type C, and Type D wind turbines Topology of the AC wind park (star or string connection)

The design variables are selected by:

Power rating of the VSC This multi-objective optimisation problem is solved by a design process know as Progressive Design Methodology [Kumar'08, Kumar and Bauer'08], which essentially consists of three stages. In the first stage, the goal is to select a group of optimal hybrid wind park solutions which have low LPC, small voltage deviation at the PCC, and high reliability with relatively simple steady-state models. In the intermediate stage, Pareto optimal set of solutions obtained from the first stages optimization is screened to reduce the number of feasible solutions. In the final stage, all the objectives and constraints that could not be considered in the first and second stages are taken into consideration. In this case, the dynamic performances of the intermediate configurations are evaluated through dynamic simulations. The final solution has the lowest LPC with an acceptable flicker level and LVRT capability. In the first stage, the multi-objective optimisation is performed. A simplified wind park model is developed in this stage which is used to evaluate the steady-state performance of the hybrid wind park in terms of investment, total energy loss and energy yield, steadystate voltage deviation, and reliability. The design objectives are to keep: LPCs as low as possible Steady-state voltage deviation at PCC as low as possible Reliability of wind park as high as possible.

After the first phase, designation of a large set of feasible solutions is possible (Pareto optimal solutions[Kumar'08, Kumar and Bauer'08]). In the intermediate stage, the set of Pareto optimal solutions obtained from the multi-objective optimisation is screened to reduce the list of feasible solutions. Additional design objectives and constraints that cannot be expressed explicitly in mathematical form are taken into consideration. These objectives and constraints are evaluated based on the experience of the design engineers. The additional criteria here are: At least one Type D wind turbine at each string, which can provide reactive power support even when there is no wind speed;

130

CH.5 Wind Park Conceptual Design and Control

Available reactive power ratio that is as high as possible, which can provide reactive power to improve LVRT capability of the wind park;

Levelised production costs that are as low as possible. After the intermediate phase, the number of solutions will be reduced to less than 10. Because of the limited number of feasible solutions, all the objectives and constraints that could not be considered in the first phase are taken into consideration. In this case, the dynamic performances of the intermediate configurations will be evaluated through dynamic simulations. The final solution will have the lowest production costs with acceptable flicker levels and LVRT capability: Lowest levelised production cost; Wind park has an acceptable flicker level of Pst < 0:2; All wind turbines can stay connected to power grid during grid-voltage dips. The voltage dip percentage and duration follow the LVRT requirement of the grid code. In the first and intermediate stages, the simple steady-state model of the wind park is used, while in the final stage, the complex and detailed dynamic model of the wind park is used. These models will be discussed in detail below.

5.3.2 Simple Steady-State Model 5.3.2.1 Parameters


The parameters of Type A, Type C, and Type D wind turbines are listed in appendix C. Other parameters such as power capacity, nominal voltage rating, short circuit ratio, and wind speed in Weibull distribution of the studied wind park are listed in Table 5-7. All the parameters of the generators in appendix C. refer to the stator. K , A and Vw
avg

are the

parameters of wind speeds Weibull distribution function, as stated in equation (2.2) in Chapter 2They are used in steady-state models to calculate the design objectives values with different sets of design variables..
Table 5-7. Wind park parameters. Number of turbines 20

Pnom
55MW

Vnom
34kV

Offshore cable length 15km

Short circuit ratio at PCC 10

Short circuit impedance angle 65

Wind speed Weibull factors K=2.336, A=9.643, Vw_avg=10.5m/s

131

5.3 Hybrid Wind Park Optimisation

5.3.2.2 Energy Production and Loss


The annual energy production and loss of a hybrid wind park is calculated by integrating the power production and power loss of the wind turbines at different wind speeds following the annual Weibull wind speed distribution. The wind turbine power production is calculated using the following equation:

P=

1 R2 Cp (; )(Vw )3 2

(5.5)

where the power coefficient Cp (; ) for different wind turbines is taken from ECNs wind turbine database [Pierik, et al.'04]. The power production at different wind speeds of Type A, Type C, and Type D wind turbines are plotted in Figure 5-7. The passive stall control of the Type A wind turbine decreases its captured wind power slightly at high wind speeds.
3000 Type A Type C Type D

Pwt (kW)

2000

1000

10

15 20 wind speed (m/s)

25

30

Figure 5-7. Wind turbine power vs. wind speed.

Generator power loss includes iron loss, gearbox loss, stator and rotor copper loss, converter loss, and transformer loss. They are calculated according to the following equations [Polinder, et al.'06, Slemon'92] and plotted in Figure 5-8. Iron loss is equal to:

Piron = 2MF e PF e0h (

fe BF e 2 fe BF e 2 )( ) + 2MF e PF e0e ( )2 ( ) f0 B0 f0 B0

(5.6)

where fe is the actual frequency of magnetic field, BF e is the actual magnetic flux density in the iron, PF e0h and PF e0e are the hysteresis loss and eddy current loss at nominal flux density B0 and nominal frequency fe, respectively. Gearbox loss is equal to:

Pgear = Pgear

nom

n nnom

(5.7)

132

CH.5 Wind Park Conceptual Design and Control

Type A Loss (kW)

100
Iron Gearbox

50

Stator Winding Rotor Winding Transformer

10 15 wind speed (m/s)

20

25

Type C Loss (kW)

100
Iron Gearbox

50

Stator Winding Rotor Winding Transformer Converter

10 15 wind speed (m/s)

20

25

Type D Loss (kW)

400
Iron Stator Winding Transformer Converter

200

10 15 wind speed (m/s)

20

25

Figure 5-8. Power losses of Type A, Type C and Type D wind turbines.

Copper loss is equal to:

Pcu = 3I 2 R
Converter loss is equal to:

(5.8)

Pconv = Pconv

nom (k1

+ k2

Iconv Iconv
nom

+ k3 (

Iconv Iconv
nom

)2 )

(5.9)

The power losses of different wind turbine concepts at different wind speeds are calculated and plotted in Figure 5-8. From Figure 5-8, it can be seen that the power production of the

133

5.3 Hybrid Wind Park Optimisation

VSWT is higher than the CSWT, however its power loss is also higher than CSWT. The final annual energy yield of the VSWT is not much larger than the CSWT.

5.3.2.3 Generator Reactive Power Capability


The reactive power capabilities of different wind turbine concepts at different operation points are calculated in this section. The Type A wind turbine always absorbs reactive power. The impedance of induction generator can be calculated if its rotation speed is known.

Rt = Rs +

!slip (( !Rr )2 slip

2 Rr X m + (Xm + Xr )2 )

(5.10)

Xt = Xs +

Xm ( !Rr )2 + Xm Xr (Xm + Xr ) slip ( !Rr )2 + (Xm + Xr )2 slip

(5.11)

where !slip is the slip of the Type A wind turbine expressed as a percentage, a dimensionless value. The total reactive power absorbed by the Type A wind turbine is calculated by adding the no-load reactive power compensation from its shunt capacitor bank as (5.12).

Qs =

Vt2 Xt V2 + t 2 R2 + Xt Xc t

(5.12)

The Type C wind turbines reactive power capability is influenced by physical limitations, such as stator and rotor currents. Stator reactive power limitation by maximum stator current limitations is equal to: r r Ps Ps 2 2 Is max ( )2 Qs Is max ( )2 (5.13) Vt Vt Stator reactive power limitation by maximum rotor current limitations is equal to:
Vt2 Vt L m !s (Ls + Lm ) Ls + Lm Qs Vt2 Vt L m + !s (Ls + Lm ) Ls + Lm s
2 Irmax (

2 Irmax (

Ps (Ls + Lm ) 2 ) Vt Lm

(5.14)

Ps (Ls + Lm ) 2 ) Vt Lm

From (5.13) to (5.14), the reactive power capability curves of a 2.75 MW DFIG (NM92) with a constant terminal voltage can be calculated. Because the stator copper winding has more over-current capability than the rotor VSC, it is the the rotor current limitation which determines the stator reactive power limitation of the Type C wind turbine.

134

CH.5 Wind Park Conceptual Design and Control

The Type D wind turbines reactive power capability is only determined by its current limitations of the grid-side VSC. r r Ps 2 Ps 2 2 Iconv max ( ) Qs Iconv max ( )2 (5.15) Vt Vt The above reactive power capabilities of the Type A wind turbine, Type C wind turbine and Type D wind turbine are implemented in the steady-state model to calculate the steady-state voltage deviation at PCC point.

5.3.2.4 Reliability Model


The reliability of wind turbines differ. For example, the CSWT has fewer components than the VSWTs with DFIG or the PMG. Their stators are similar, but the CSWT does not have a VSC, as is the case for Type C wind turbines and Type D wind turbines. Thus it can be concluded that CSWTs have a higher reliability. However CSWT blades endure greater stress than those of VSWTs =, thus its blades have less reliability. The reliability of different types of wind turbines will have an influence on the reliability of the entire hybrid wind park. Reliability of different topologies of wind parks also differ. There are many configurations of wind parks [Pierik, et al.'01], but they are all composed of two basic topologies, such as star connection and string connection, as shown in Figure 5-9. The reliability of these two topologies apparently differ: the topology (a) star conncetion has a much higher reliability compared to the topology (b) string connection. The influence of topology on reliability is calculated as follows. Firstly, the assumption of N-1 rule is borrowed here, which means that at any time only one AC cable is out of service. Secondly, an index which is called unavailable generation UG is used to evaluate reliability of the wind park based on the assumption of N-1 rule [Zhao, et al.'06, '07]. UG is the total power that cannot be transmitted to the grid when one component of the wind park (AC cable) falls out of service. Finally, it is assumed that there is no relation between component reliability and power flow/wind speed. UG can be calculated using U G = Punavailable f ailure, wherePunavailable is the power influenced by the failure of the component, which is apparently the calculated power rating of that cable; and failure is the probability of the component failure. Because of the assumption of N-1 rule used, failure of the AC cable can be calculated as:

f ailure = Ucable (1 Ucable )N 1


unavailability of the AC cable, and N is the total number of AC cable. Unavailability of cable Ucable is calculated as:

(5.16)

which means only one cable fails, and all other cables are still in service. Ucable is the

135

5.3 Hybrid Wind Park Optimisation

Ucable = 1

1 1+f

(5.17)

where f is the failure rate of the cable, and is the repair duration of the cable.

Figure 5-9. Topologies of the wind park (a) string connection (b) star connection.

5.3.2.5 Cost Model


The cost of a wind park is derived from the database of the ERAO I project [Bauer, et al.'00, Pierik, et al.'01]. As the cost database is confidential, all the prices listed here are normalised to the cost of wind turbine blades. They are listed in the following table.
Table 5-8. Cost breakdown of wind turbine components. AC cable (per km per MVA) 0.0095

Blades

Active pitch control 0.4975

Tower and other components 0.8403

Gearbox

Induction generator

PM Generator

VSC (per MVA) 0.182

AC transformer (per MVA) 0.0255

1.0

0.4

0.109

0.567

5.3.2.6 Wind Park Topology Simplification


As shown in Figure 5-10 and Figure 5-11, wind parks can have different topologies, which make it difficult to calculate the power flow of a wind park. The power flow inside Wind Park can be calculated power flow programs, such as MatPower in Matlab. It can also be

136

CH.5 Wind Park Conceptual Design and Control

calculated with simplified equivalence impedance under certain conditions. In this hybrid wind farm optimization, only AC grid is considered, and the offshore cable length is limited. The loss of offshore grid is small, compared with the loss inside the generator, and the power electronic converter. Thus the power flow calculation can be greatly simplified to use an equivalent impedance to represent the wind parks grid.
Z1 Z2 Z3 Z4

I1

I2

I3

I4

Is

Is

Zs

Figure 5-10. Equivalent impedance of a wind park (a) actual string connection (b) equivalent representation.

The equivalent impedance Zs can be used if the P and Q power losses on the string are the same, which is:
2 2 Is Zs = I1 Z1 + (I1 + I2 )2 Z2 + (I1 + I2 + I3 )2 Z3 + (I1 + I2 + I3 + I4 )2 Z4

(5.18)

If I1 = I2 = I3 = I4 is assumed, then the equivalent impedance is: Pn m2 Zm Zs = m=1 2 n where n is the number of wind turbines on one string.

(5.19)

where nm is the number of wind turbines connected in one cable. equivalent impedance can be further simplified as: Pn Zm Zp = m=1 n

A similar simplification can be done on the star connection, as shown in Figure 5-11. The equivalent impedance of the star connection as shown in Figure 5-11 can simplified as: Pn 2 m=1 nm Zm Zp = Pn (5.20) ( m=1 nm )2 The simplest star connection is that one AC cable only has one wind turbine, therefore the

(5.21)

137

5.3 Hybrid Wind Park Optimisation

Ip
I3 I4

I1

I2

Z1

Z2

Z3

Z4

n1

n2

n3

n4

Ip

Zp

Figure 5-11. Equivalent impedance of a wind park (a) actual star connection (b) equivalent representation.

The steady-state voltage deviation at PCC The steady-state voltage deviations at the PCC can be calculated with simplified wind park topology.

Vpcc =

Ppcc Rsc + Qpcc Xsc Vnom

(5.22)

After simplification, a detailed calculation with a power flow program (MatPower) is used to verify the simplified offshore grid of the hybrid wind park. The case comparison is shown in Table 5-9.
Table 5-9. Simulated case to compare detailed model and simplified model. Total turbines 20 Type A wind turbine 10 Type C wind turbine 1 Type D wind turbine 9 Number of WT per string 4 Number of strings 5 P output Full power

The calculated active and reactive power at the PCC is listed in Table 5-10.
Table 5-10. Simulation results with the two models.

Simplified model Detailed model

P at PCC (MW) 50.12 50.16

Q at PCC (MVar) -18.06 -17.99

138

CH.5 Wind Park Conceptual Design and Control

It can be seen that the differences in active power P and reactive power Q between simplified model and detailed mode of the offshore grid are very small. In the later section, the simplified model is used to calculate both the loss of the offshore grid and voltage deviation at PCC.

5.3.3 Optimisation Results 5.3.3.1 First Stage


Results of the first optimisation stage are plotted from Figure 5-12 to Figure 5-15. In Figure 5-12, the levelised production cost, unavailable generation ratio and steady-state voltage deviation of about 100 optimised configurations are shown in a 3D plot. The Pareto front surface in the three objective optimisation is not as clear as that of a twoobjective optimisation [Kumar'08], because some design objectives are not strongly correlated. This can be understood better in the 2D plots from Figure 5-13 to Figure 5-15. In Figure 5-13, the Pareto front curve is clear because the levelised production cost is strongly opposed to the steady-state voltage deviation. The VSWT concept has better reactive power capability and can control the voltage better at the PCC, but the VSWT is more expensive than the CSWT. The reliability of the CSWT is better than the VSWT because the failure rate of the gearbox is much lower than the failure rate of power electronic converters and sensors, as stated by database [Durstewitz'06, Eggersglss'00]. Thus in Figure 5-14, the levelised production cost is positively related to the unavailable generation ratio. However, because they are both influenced by the wind park topology, the Pareto front curve is not as clear as in Figure 5-13. Finally in Figure 5-15, the steady-state voltage deviation is influenced mainly by the type and converter rating of the wind turbine; the reliability - unavailable generation ratio is influenced by both the wind turbine type and wind park topology. Thus the relation between voltage deviation and unavailable generation ratio is not very strong and the Pareto front curve is not as clear as in Figure 5-13.

139

5.3 Hybrid Wind Park Optimisation

3D plot of Cost, Reliability and Performance

Performance - Voltage Dev

0.03 0.02 0.01 0 0.03 0.035 0.025 0.025 Reliability - UGR 0.02 0.02 Cost - LPC 0.03

Figure 5-12. Levelised production cost, unavailable generation ratio, voltage deviation at PCC

Cost Vs. Reliability 0.03

0.028

0.026 UGR 0.024 0.022 0.02 0.02

0.025 LPC

0.03

0.035

Figure 5-13. Levelised production cost and voltage deviation at PCC.

140

CH.5 Wind Park Conceptual Design and Control

Performance Vs. Reliability 0.03

0.028

0.026 UGR 0.024 0.022 0.015 0.02 0.025 Voltage Dev Figure 5-14. Levelised production cost and unavailable generation ratio. 0.02

0.005

0.01

0.03

Cost Vs. Performance 0.03 0.025 0.02 0.015 0.01 0.005 0 0.02

0.03 LPC Figure 5-15. Voltage deviation and unavailable generation ratio.

Voltage Dev

0.025

0.035

141

5.3 Hybrid Wind Park Optimisation

5.3.3.2 Intermediate Stage


Results from the multi-objective optimisation obtained in the first stage are screened to reduce the number of the feasible solution set. Additional design objectives are formulated in this stage: At least one Type D wind turbine at each string, which can provide reactive power support even when there is no wind speed; Available reactive power ratio
Qmax Swf

as high as possible, which can provide

reactive power to improve LVRT capability of wind park; Levelised production cost as low as possible. The importance of the additional design objectives is listed.in Table 5-11. The importance gives the weight of the corresponding criteria based on the design engineers experience, where VH means very high, H means high, M means medium, L means low, and VL means very low. The direction defines the beneficial direction of the corresponding criteria, where H means the criteria should be as high possible, and L means as low as possible.
Table 5-11. Importance of design criteria in the intermediate optimisation phase. Additional design objectives Number of Type D wind turbine per string 1 Q capability of wind park
Qmax Swf

Importance M VH VH VL M

Direction H H L L H

Levelised production cost Steady state voltage deviation at PCC Reliability of wind park

After this intermediate optimisation phase, the solutions are reduced from approximately 100 to 3. They are shown in Table 5-12.
Table 5-12. Results of the intermediate optimisation phase. Config No. 1 2 3 Number Type D 6 6 7 Number Type C 0 5 9 Number Type A 14 9 4 Turbines per string 4 4 4 S_Type D wind turbine 1.5 1.8 1.9 S_Type C wind turbine 1.5 1.5 1.0 LPC 0.026 0.028 0.030 GRA 0.978 0.978 0.978 DEV 0.003 0.002 0.002 Qcap 0.37 0.73 0.81

5.3.3.3 Final Stage


In the final stage, dynamic simulation is used to calculate flicker and LVRT capability of the limited solutions from the intermediate phase. In this stage, complex dynamic models are required [Morren, et al.'05, '05, Zhou, et al.'08] and the central wind park controller as

142

CH.5 Wind Park Conceptual Design and Control

discussed in section 5.4 is designed to control the reactive power at PCC, both to limit the flicker level and improve the LVRT capability.. The flicker level of three solutions from the intermediate optimisation stage is calculated with the offline flicker meter, and the results are shown as Figure 5-16.
Flicker level of Hybrid WindFarm, normal operation, grid impedance angle 65o 0.16 config 1 config 2 0.15 config 3

Instantaneous flicker

0.14

0.13

0.12

10

20

30

40 50 60 time second

70

80

90

100

Figure 5-16. Flicker level of three hybrid wind park configurations..

The LVRT capability of this hybrid wind park is indicted by the over-speed of the induction generator of the CSWT [Akhmatov'03] and the residual voltage at the PCC during grid-voltage dips. A 0.5pu grid-voltage dip is applied and lasts for 5 seconds. The results are shown as Figure 5-17 and Figure 5-18. generation speed of CSWT in cluster 1 of hybrid WF 120 config 1 115 config 2 config 3 110
generator speed (rad/s) 105 100 85 90 time (s)
Figure 5-17. Voltage at the PCC of three hybrid wind park configurations.

95

100

143

5.4 Wind Park Controller

voltage at PCC of hybrid WF voltage at PCC(pu) 1 0.8 0.6 0.4 85 config 1 config 2 config 3

90 time (s)

95

100

Figure 5-18. Generator rotation speed of a CSWT of three hybrid wind park configurations.

From results shown in Figure 5-17 and Figure 5-18, it is clear that the LVRT capability of configuration 1 cannot meet the requirements. The rotation speed of the CSWT accelerates so much that it has the possibility of rising above its stability margin [Akhmatov'03]. It is also seen that the voltage at the PCC cannot recover to its pre-disturbance level when the grid-voltage dip is already removed at 95 seconds. The low voltage is caused by the large reactive power drawn from the CSWT, because its rotation speed increases to a very large value, as indicated by equations (5.11) and (5.12). Finally, configuration 2 is chosen as the best configuration because it has the lowest levelised production cost with an acceptable flicker level and an acceptable LVRT capability, as Figure 5-18. The configuration is: six Type D wind turbines, zero Type C wind turbine, fourteen Type A wind turbines, and four wind turbines per string, as shown in Table 5-13.
Table 5-13. Final choice of the optimal wind park configuration. Number of Type D 6 Number of Type C 5 Number of Type A 9 Turbines per string 4 Converter Rating of Type D 1.8 Converter Rating of Type C 1.5

5.4 Wind Park Controller


In section 5.3, the configuration of a hybrid wind park is optimised through a multiobjective optimisation method. In order to be integrated successfully with the power grid, the hybrid wind park has to be controlled to meet the grid code requirements. For example, the reactive power capability of variable speed wind turbines has to be used to control voltage at the point of common coupling to limit the flicker level and to improve LVRT capability of the nearby CSWTs. The reactive power control can be implemented by the

144

CH.5 Wind Park Conceptual Design and Control

voltage-droop control of individual wind turbine. However, because each wind turbine or each cluster of wind turbines faces different wind speeds, in some aspects a central reactive power controller is superior to an individual droop controller. Firstly, individual droop control implies that the total reactive power of the wind park is not controlled precisely. Secondly, reactive power consumed by the transformer and produced by AC cable makes it even more difficult to meet certain grid codes, such as the unity power factor at the point of common coupling. Finally, with individual droop controllers, the reactive power demand of the whole wind park is not dispatched among wind turbines according to their reactive power capability. To improve transient stability during a grid fault, a certain amount of reactive power has to be reserved. For this consideration, a centralised control of the hybrid wind park is designed in Figure 5-19. In this thesis, only the centralised reactive power controller is used as an example. The centralised active power controller is similar to the reactive power controller and is only necessary when the wind park is operated in de-loaded mode (active power reserve mode). It is not described in detail here. The effectiveness of the central reactive power controller will be demonstrated through dynamic simulation of a hybrid wind park which includes both CSWTs and VSWTs.

5.4.1 Central Controller Concept


The concept of a central wind park controller is shown as Figure 5-19. It has three main function blocks: active and reactive power demand determination; an anti-windup PI controller; and a dispatching function. Either the maximum active power production mode or the active power reserve mode can be chosen for the active power controller of the wind park. For the reactive power controller, besides the standard constant power factor requirement, the reactive power of the VSWTs inside the wind park can also be controlled to limit flicker mitigation, and to improve transient voltage stability during a grid fault. The anti-windup PI controller is necessary because of the possible limitation of active and reactive power of the wind park. When the active or reactive power reference exceeds the limitation, the error between demand and actual values will accumulate through the integral part of the PI controller. This will cause large oscillations in the output of the PI controller when the reference power is less than the limitation again. With the anti-windup block, the negative effects of the PI controller with limitation can be eliminated. The third block is the dispatching function, which has the functions as:

Pwt =

Pwt max Pwf Pwf max

ref

(5.23)

145

5.4 Wind Park Controller

Qwt =

Qwt max Qwf Qwf max

ref

(5.24)

Figure 5-19. Generic wind park central controller.

The active and reactive power limitation of the wind park is the sum of those of individual wind turbines. The reactive power capability of each type of wind turbines is calculated in section 5.3.2.3. The dispatching function makes sure that each wind turbine will output the active and reactive power proportional to its capabilities. It can be understand by the following example of the reactive power. When a Type C wind turbine faces large wind speed, it will output large active power, thus its reactive power capability decreases, which can be calculated from equations from (5.13) to (5.14). According to dispatching function (5.24), its reactive power demand will be smaller compared to Type C wind turbines which face lower wind speed.

5.4.2 Case Study

Figure 5-20. Topology of the hyrid wind park.

146

CH.5 Wind Park Conceptual Design and Control

A hybrid wind park which contains both CSWTs and VSWTs is modelled. Its topology is shown as Figure 5-20 and its main characteristics are the following: 20 wind turbines, 2.75 MW per turbine 4 clusters, 5 turbines in each cluster Type C wind turbines and Type D wind turbines work as distributed STATCOMs with reactive power limitation Short circuit at the PCC is 1100MW, short-circuit ratio is 20 at PCC, grid impedance angle is 65 The reactive power limitations of the VSWT are implemented in a 3D reference table in the wind park controller. Using the distribution function, each wind turbine only contributes to part of the reactive power demand and its reactive power output stays far away from its operating limitation.

5.4.3 Simulation Results


In the simulation, each cluster faces different wind speed, and wind turbines in the same cluster face the same wind speed as shown in Figure 5-21. Cluster Effective Wind Speed
20 Vw (m/s) 15 10 5 0 20 40 60 time (s) 80 100 120 Cluster 1 Cluster 2 Cluster 3 Cluster 4

Figure 5-21. Wind speed of four clusters.

A 120s period is simulated, and the following five simulation scenario events are arranged:. Region I unity power factor of Type C wind turbine and Type D wind turbine Region II unity power factor at PCC Region III flicker mitigation at PCC Region IV voltage droop control, grid-voltage drop to 0.9 pu Region V grid voltage recovered to 1.0 pu

Simulation results can be seen in Figure 5-22 to Figure 5-25, which are voltage at the PCC, reactive power flow at the PCC, active and reactive power of the Type A wind turbine, and active and reactive power of the Type C wind turbine, respectively. In Region I, the unity

147

5.4 Wind Park Controller

power factor operation of each Type C wind turbine and Type D wind turbine does not guarantee a unity power factor at the PCC, because of the reactive power consumed by the Type A wind turbine and transformer and produced by the offshore cable. In Region II, the power factor at the PCC is controlled to zero. However, due to the relatively small X/R ratio (the grid impedance angle is 65), which is typical for a low-voltage level distribution network, the active power flow causes large voltage deviation at the PCC. Then in Region III, the steady-state voltage at PCC is controlled at a constant by using the reactive power of the VSWT as in equation (5.25).

Qpcc =

Ppcc Rsc Xsc

(5.25)

In Region IV, the grid-voltage drops and reactive power of wind park increases to improve the LVRT capability of the Type A wind turbine. Finally in Region V, the grid voltage recovers and the reactive power control strategy of wind park returns to a constant steadystate voltage control at the PCC.
Line to line Voltage at PCC 35 34 Vll (kV) 33 32 31 30 0 20 40 60 time (s) 80 100 120

Figure 5-22. Line-to-line voltage at the PCC.


Reactive power flow at PCC 20 Qpcc (MVar) 10 0 -10 -20 0 20 40 60 time (s) 80 100 120 Q reference Q actual

Figure 5-23. Reactive power flow at the PCC.

The active and reactive power of Type A and C wind turbines are plotted in Figure 5-24 and Figure 5-25. The active power of both Type A (CSIG) and Type C (DFIG) increase

148

CH.5 Wind Park Conceptual Design and Control

with wind speed. It can be seen that the wind speed fluctuation is directly reflected in the active power output of the Type A wind turbine. Thus the flicker will be a problem if the voltage at the PCC is not controlled by the reactive power of Type C and Type D wind turbines. The reactive power of VSWTs in Figure 5-25 is controlled to fulfil the requirements in different regions: Region I unity power factor of VSWTs: The VSWTs reactive power is controlled to be zero; Region II unity power factor at PCC. The VSWTs reactive power is controlled to compensate the reactive power consumed by the CSWTs, consumed by the transformer, and generated by the cable; Region III flicker mitigation at the PCC. The VSWT reactive power is controlled to compensate the flicker produced by the active power fluctuation of the CSWTs; Region IV voltage-droop control when grid-voltage drops to 0.9 pu. The VSWT reactive power is controlled to limit the voltage drop at the PCC, thus improving the LVRT capability of the CSWTs; Region V grid voltage recovered to 1.0 pu. The VSWT reactive power controls the unity power factor at the PCC.

In short, the reactive power flow of the hybrid wind park is controlled satisfactorily and total reactive power demand is dispatched to the VSWTs according to their reactive power capability.

5.5 Conclusion
In this chapter, the conceptual design and control of a wind park are presented. Firstly, the transmission capability of a 220kv HVAC cable is calculated. The integration choices of an HVAC, HVDC LCC and HVDC VSC are compared based on their transmission costs. The critical integration distance between an HVAC 220kv cable and HVDC VSC is shown to be around 120km. Secondly, a hybrid wind park concept is proposed which contains both CSWTs and VSWTs. VSWTs in the hybrid wind park work as distributed STATCOMs to limit the voltage deviation at the PCC and to improve the LVRT capability of the wind park. A multi-objective multi-criteria optimisation method is used to optimise the hybrid wind park to reach the best configuration. The optimisation objective is to have low levelised production costs, high reliability, an acceptable flicker level and LVRT capability. Finally, the individual wind turbine controller and the central wind park controller are compared. Benefits of the central controller for the wind park are addressed. A generic

149

5.5 Conclusion

central controller using an anti-windup PI controller and distribution function is designed to control the hybrid wind park to work as an integral wind power plant. Its effectiveness is proved by the simulation results.

150

CH.5 Wind Park Conceptual Design and Control

Type A Active Power 4


Cluster 1

Ps (MW)

3 2 1 0 0

Cluster 2 Cluster 3 Cluster 4

20

40

60 80 time (s) Type A Reactive Power

100

120

1 Qs (MVar) 0 -1 -2 -3 0 20 40 60 time (s) 80 100 120

Figure 5-24. Active and reactive power of the Type A wind turbine.
Type C Active Power 4
Cluster 1

Ps (MW)

3 2 1 0 0

Cluster 2 Cluster 3 Cluster 4

20

40

60 80 time (s) Type C Reactive Power

100

120

3 2 Qs (MVar) 1 0 -1 -2 0 20 40 60 time (s) 80 100 120

Figure 5-25. Active and reactive power of the Type C wind turbine.

151

5.6 References

5.6 References
[ABB09] [Ack05] [Akh03] ABB, "ABB library for HVDC," last accessed: June, 2009, available online: http://www.abb.com/industries/us/9AAF400196.aspx. T. Ackermann, Wind power in power systems, ISBN-0-470-85508-8, John Wiley & Sons, 2005. V. Akhmatov, Analysis of dynamic behavior of electric power systems with large amount of wind power, PhD thesis, Electrical Power Engineering Department, Technical University of Denmark, Copenhagen, Denmark, 2003. V. Akhmatov, "Modelling and ride-through capability of variable speed wind turbines with permanent magnet generators," Wind Energy, vol. 9, issue 4, 2005. P. Bauer, S. W. H. De Haan, C. R. Meyl, and J. T. G. Pierik, "Evaluation of electrical systems for offshore windfarms," in 2000 IEEE Industry Applications Conference., 2000. F. Blaabjerg and Z. Chen, Power Electronics for Modern Wind Turbines, Morgan & Claypool Publishers, 2006. M. H. J. Bollen, G. Olguin, and M. Martins, "Voltage Dips at the Terminals of Wind Power Installations," Wind Energy, vol. 8, issue 3, Page(s). 307-318, 2005. H. Brakelmann, "Efficiency of HVAC power transmission from offshorewindmills to the grid," in IEEE Power Tech Conference Proceedings, 2003, Bologna, Italy, 2003. H. Brakelmann, "Loss determination for long three-phase high-voltage submarine cables," European Transactions on Electrical Power, vol. 13, issue 3, Page(s). 193-197, 2003. R. D. Dunlop, R. Gutman, and P. P. Marchenko, "Analytical Development of Loadability Characteristics for EHV and UHV Transmission Lines," IEEE Transactions on power apparatus and systems, vol. 98, issue 2, Page(s). 606617, 1979. M. Durstewitz, "Wind Energy Report Germany 2001-2006; Annual Evaluation of WMEP," 2006. W. Eggersglss, "Windenergie X, XI, XII, XIII Praxisergebnisse 1997, 1998, 1999, 2000 Landwirtschaftskammer Schleswig-Holstein.," 2000. R. Hoffmann, a comparison of control concepts for wind turbines in terms of energy capture, Elektrotechnik und Informationstechnik Department, Technischen University at Darmstadt, Darmstadt, Germany, 2002. P. Kumar, A frame work for multi-objective optimization and multi-criteria decision making for design of electrical drives, Department, Delft University of Technology, 2008. P. Kumar and P. Bauer, "Progressive design methodology for complex engineering systems," Soft Computing, vol. online version, September, 2008 2008. P. Kundur, Power System Stability and Control, ISBN-0-07-035958-x, McGraw-Hill Professional, 1994.

[Akh05]

[Bau00]

[Bla06] [Bol05]

[Bra03]

[Bra03]

[Dun79]

[Dur06] [Egg00] [Hof02]

[Kum08]

[Kum08]

[Kun94]

152

CH.5 Wind Park Conceptual Design and Control

[Lar99]

[Laz05]

[Lun03] [Mar03]

[Mor05]

[Mor05]

[Neg05]

[Neg06]

[Pie01]

[Pie04]

[Pol06]

[Sie09]

[Sle92] [Sun04]

[Tod04]

T. Larsson, C. Poumarede, A. Syst, and S. Vasteras, "STATCOM, an efficient means for flicker mitigation," IEEE 1999 Power Engineering Society Winter Meeting, vol. 2. L. P. Lazaridis, Economic Comparison of HVAC and HVDC Solutions for Large Offshore Wind Farms under Special Consideration of Reliability, Department of Electrical Engineering Department, Royal Institute of Technology of Sweden, Stockholm, 2005. S. Lundberg, "Performance comparison of wind park configurations," Chalmers University of Technology, Goteborg 2003. M. I. Marei, T. H. M. El-Fouly, E. F. El-Saadany, and M. M. A. Salama, "A flexible wind energy scheme for voltage compensation and flicker mitigation," Power Engineering Society General Meeting, 2003, IEEE, vol. 4, 2003. J. Morren, J. T. G. Pierik, S. W. H. de Haan, and J. Bozelie, "Grid Interaction of Offshore Wind Farms. Part 1. Models for Dynamic Simulation," Wind Energy, vol. 8, issue 3, Page(s). 265-278, 2005. J. Morren, J. T. G. Pierik, S. W. H. de Haan, and J. Bozelie, "Grid Interaction of Offshore Wind Farms. Part 2. Case Study Simulations," Wind Energy, vol. 8, issue 3, Page(s). 279, 2005. N. B. Negra, Losses Evaluation of HVDC solutions for Large Offshore Wind Farms, Department of Electrical Engineering Department, Royal Institute of Technology of Sweden, Stockholm, 2005. N. B. Negra, J. Todorovic, and T. Ackermann, "Loss evaluation of HVAC and HVDC transmission solutions for large offshore wind farms," Electric Power Systems Research, vol. 76, issue 11, Page(s). 916-927, 2006. J. T. G. Pierik, M. E. C. Damen, P. Bauer, and S. W. H. de Haan, "Electrical and Control Aspects of Offshore Wind Farms, Phase 1: Steady State Electrical Design, Power Performance and Economic Modeling," ECN and TUDelft, Petten ECN-CX-01-083, 2001. J. T. G. Pierik, J. Morren, E. J. Wiggelinkhuizen, S. W. H. de Haan, T. G. van Engelen, and J. Bozelie, "Electrical and Control Aspects of Offshore Wind Farms II (Erao II) - Volume 1: Dynamic Models of Wind Farms," Energieonderzoek Centrum Nederland (ECN), Petten 2004. H. Polinder, F. F. A. van der Pijl, G. J. de Vilder, and P. Tavner, "Comparison of direct-drive and geared generator concepts for wind turbines," IEEE Transactions on Energy Conversion, vol. 21, issue 3, Page(s). 725-733, 2006. Siemens, "High-voltage DC Transmission Systems," June,2009, https://www.energyportal.siemens.com/static/de/de/products_solutions/1650_kn03011202.html G. R. Slemon, Electric machines and drives, ISBN: 0201578859, AddisonWesley, 1992. T. Sun, Z. Chen, and F. Blaabjerg, "Flicker mitigation of grid connected wind turbines using STATCOM," in 2nd International Conference on Power Electronics, Machines and Drives, PEMD 2004., Edinburgh,UK, 2004. J. Todorovic, Losses Evaluation of HVAC Connection of Large Offshore Wind Farms, Master Thesis thesis, Department of Electrical Engineering Department, Royal Institute of Technology of Sweden, Stockholm, 2004.

153

5.6 References

[Wen06]

[Wri02]

[Zha08]

[Zha06]

[Zha07]

[Zho08]

D. Wensky, J. Hanson, P. Sandeberg, and R. Grnbaum, "FACTS and HVDC for grid connection of large offshore wind farms," in European Wind Energy Conference EWEC, 2006, 2006. S. D. Wright, A. L. Rogers, J. F. Manwell, and A. Ellis, "Transmission Options for Offshore Wind Farms in the United States," in AWEA 2002, USA, 2002, Page(s). 112. M. Zhao, Optimization of Electrical System for Offshore Wind Farms via a Genetic Algorithm Approach, Department of Electrical Engineering Department, Aalborg University, Aalborg, Denmark, 2008. M. Zhao, Z. Chen, and F. Blaabjerg, "Generation Ratio Availability Assessment of Electrical Systems for Offshore Wind Farms," in 12th International Power Electronics and Motion Control Conference, EPE-PEMC 2006. , Portoro, Slovenia, 2006, Page(s). 561-566. M. Zhao, Z. Chen, and F. Blaabjerg, "Generation Ratio Availability Assessment of Electrical Systems for Offshore Wind Farms," IEEE Transactions on Energy Conversion, vol. 22, issue 3, Page(s). 755-763, 2007. Y. Zhou, P. Bauer, and J. A. Ferreira, "Reactive power limitation of doubly fed induction generator and its application on coordinated hybrid wind farm control," in European Wind Energy Conference 2008, EWEC2008, Brussels, Belgium, 2008.

154

CH.6 Power System with Large Wind Power Installation

Chapter 6
Power System with Large Wind Power Installation

6.1 Introduction
The prime mover governing system of classical power plants (i.e., thermal power plants and hydro power plants) provides a means of controlling power and frequency of the power system, which are referred to as primary frequency and secondary frequency supports. Today both total wind power installation and single wind park power ratings are increasing. The operation of wind power plants is very different from the operation of classical power plants. The generator rotation speed of VSWTs is completely decoupled from the power-grid frequency. The VSWT is often working in maximum-power production mode, and its primary source (i.e., wind speed) cannot be controlled. Thus the influence of a large wind power installation on the frequency response of the power system has to be studied. Small-signal stability is the ability of the power system to maintain synchronism when subjected to small disturbances. The small-signal stability of a traditional power system is the insufficient damping of one or several modes of system oscillations. The high-response exciter and weak grid connection of the SG are often the reasons for low and even negative damping [Anderson and Fouad'03, Kundur'94, Rogers'00]. A PSS is installed on the exciter of the SG to improve the small-signal stability. As both the structure and operation of wind power plants are very different from classical power plants, a quantitative analysis of the small-signal stability of a power system with a high wind power installation is required. In this chapter, the behaviour of a power system with 20% wind-power penetration is investigated. A four-machine power system is chosen as the case study, which is used as the benchmark of the multi-machine power system model in [Anderson and Fouad'03, Kundur'94, Rogers'00]. The frequency responses of power systems with and without windpower penetration are compared. The small-signal stability of power systems with and without a wind power plant is studied. The possibilities for using the inertia of a wind power plant for short-term primary frequency support and power-system oscillation damping are investigated. The fundamental limitations of inertia usage are discussed.

155

6.2 Frequency Response

6.2 Frequency Response


One important task of thermal power plants is frequency support, which includes primary frequency and secondary frequency support. Primary frequency support stabilises the frequency and prevents any further frequency deviation during grid disturbances, such as sudden load change or unplanned large power plant outage. Its time scale is from 0 to 30 seconds. Secondary frequency support restores the frequency to the reference value, and keeps the power flow on the tie lines between different power sections to the preset values. Its time scale is from 30 seconds to 30 minutes.

6.2.1 Frequency Response in Traditional Power Plants


The traditional power plants are thermal power plants and hydro power plants. Large thermal power plants use steam turbine with a re-heater, whereas hydro power plants have penstock. Both of which have time delays when increasing their input mechanical power. Traditional power plants often have a synchronous generator directly connected to the AC power grid. The active power ramp-up speed and time delay of thermal plants and hydro power plants are different. The influences of both on the frequency response of the entire system are shown in Figure 6-1. It can be seen that the thermal power plant without a reheater has the best frequency response because its mechanical power can be changed very fast. The re-heater introduces a time delay of about 6s [Anderson and Fouad'03, Kundur'94]. The frequency response of hydro power plant is the poorest, because its mechanical power first decreases and then increases very slowly when its valve is opened. Nowadays, the large steam turbine is often equipped with a re-heater to improve efficiency [Anderson and Fouad'03, Kundur'94]. Thus the frequency of the power system first has an over-dip, and then stabilises at the final value. The final stabilised frequency is determined by the active power change P , the load frequency coefficient Dload , and the droop ratio

Rdroop of power plant.


Because wind power plants have a different structure and operation mode than classical power plants, the frequency response of power systems with large wind power installations is different from that of the traditional power system and will be studied in the next section.

156

CH.6 Power System with Large Wind Power Installation

t(s)
Figure 6-1. Frequency responses of different energy sources [Anderson and Fouad'03].

157

6.2 Frequency Response

6.2.2Frequency Response in Wind Power Plants

Pmsys +

Pm0sys +

T
Pload

1 2Hsys s

!sys

Pewt +

Pe
+
=0

Pe0

Pe
Vw

Cp (; )

Pmwt +

T
1 2Hwt s

!wt

Figure 6-2. Simplified power system model for frequency response calculation.

Wind power plants are often not required for primary frequency support, because they work in maximum-power production mode and the primary energy source (i.e., wind speed) cannot be controlled. However, when wind power penetration increases to a certain level, for example 20%, the frequency response of wind power plants without primary frequency support deteriorates. A simplified multi-machine system is modelled here which has both a thermal power plant and a wind power plant. The thermal power plant with SGs and a re-heater is modelled in [Kundur'94]. The pitch control and dynamic inflow effect of the turbine blades are omitted in the wind power plant model. The inertia constant of the SG is Hsys , which is the equivalent inertia constant of the entire power system, with the exception of the load and 158

CH.6 Power System with Large Wind Power Installation

wind power plant. The simplified system model for frequency response calculation is shown in Figure 6-2, and the simulation results are shown in Figure 6-3.
50.1 50.05 50 49.95 49.9 49.85 49.8 50 55 60 65 70 75 time (s) 80 85 90 95 100 0% wind power 20% wind power without primary frequency support 20% wind power with primary frequency support 20% wind power with primary frequency support and deloading

Figure 6-3. Simulation results of the frequency response of a power system with and without wind power penetration.

In Figure 6-3, the frequencies of a power system with and without wind power penetration at sudden load increase are shown. Without wind power penetration, the lowest frequency is below 49.85 Hz, which is often a threshold for low-frequency load-shedding. With 20% wind power penetration and no primary frequency support from wind power plant, the frequency drops to as low as 49.85 Hz and triggers the low-frequency load shedding protection. Thus without primary frequency support from the wind power plant, the larger frequency dip will trigger low-frequency load-shedding protection. The large inertia of the wind turbine can be used to provide short-term primary frequency support [de Almeida and Lopes'07, Morren, et al.'06]. The capabilities of inertia usage of a wind power plant will be discussed in section 6.4. The wind power plant can use more of its inertia energy for primary frequency support, because its rotation speed is allowed to be changed in a larger range than the traditional power plant. Because of this reason, the power system frequency with primary frequency support from the wind power plant is better than the situation with only a thermal power plant, as shown in Figure 6-3. The lowest frequency with primary frequency support from the wind power plant is higher. However, as the kinetic energy of VSWT is limited and the rotation speed has to be recovered to the optimal rotation speed, after the first 10 seconds the frequency decreases and stabilises at the same value as the situation which has no primary frequency support from wind power plant, as shown in Figure 6-3.

159

6.2 Frequency Response

If the power system has other energy storage devices such as batteries, the system can then be arranged to coordinate with the wind power plant and eliminate the frequency dip caused by the speed recovery of the wind turbine [Morren, et al.'06]. For systems without any energy storage devices, one way to improve the frequency response between 10s and 30s is to de-load the VSWT. The de-loaded VSWT works in the active power reserve mode and it can increase its mechanical power input during a grid frequency dip, similar to a thermal power plant. With a de-loaded wind power plant, the frequency response of the power system becomes the best for time periods ranging from 0 to 30 seconds, as shown in Figure 6-3.

Figure 6-4. Electric power from thermal power plant and wind power plant with primary frequency control.

Figure 6-5. Conceptual power speed curves of de-loaded wind turbine.

160

CH.6 Power System with Large Wind Power Installation

The output electric power changes of the thermal power plant and wind power plant with short-term primary frequency support are illustrated in Figure 6-4. The conceptual power speed curves of the VSWT in maximum power production mode and de-loaded mode are shown in Figure 6-5. The de-loaded wind power plant will not only provide primary frequency support, but it can also provide secondary frequency support. The conceptual control loops of primary and secondary frequency support of a VSWT are shown in Figure 6-6.

Figure 6-6. Conceptual primary and secondary frequency controllers of a VSWT.

6.3 Small-signal Stability


Small signal stability of power system with synchronous generator is well studied [Anderson and Fouad'03, Kundur'94, Rogers'00]. The stabilities of a single wind turbine or other power electronic interfaced energy sources are studied in [Banakar, et al.'06, Coelho, et al.'02, Congwei, et al.'01, Gurbuz and Akpinar'02, Karlsson, et al.'05, Lee and Wang'08]. However, the small signal stability of a power system with large wind power penetration, especially variable speed wind turbine, is seldom studied. In this section, the small-signal stability of a multi-machine power system with and without wind power penetration is analysed. The small-signal model of a Type C wind turbine (DFIG) is derived from Chapter 3. The small-signal model of synchronous generators with excitation control can be found in [Anderson and Fouad'03, Kundur'94, Rogers'00]. A PSS is designed for the wind power plant to damp out the inter-area oscillation of the multi-machine power system.

161

6.3 Small-signal Stability

6.3.1 Multi-machine Power System


The small-signal stability of a power system with and without 20% wind power penetration is studied with the analysis of its eigen values. For 0% wind power penetration, the system is a four-synchronous generator system used as a benchmark multi-machine power system model in [Kundur'94, Rogers'00]. For the situation of 20% wind power penetration, G4 is replaced by a wind power plant which is modelled as an aggregated DFIG model in per unit system.
G1 - 25%

G3 - 20%

G2 - 25%

G4 - 30%

Figure 6-7. Topology of a multi-plant system [Kundur'94, Rogers'00].

6.3.2 Model Implementation


As was discussed in Chapter 3, an non-interconnected power system with generator is modelled as:

_ x = Ax + Bv V + Bu u

(6.1) (6.2)

y = Cx + Dv V + Du u

where V is change in node voltage, which should be influenced by both x and u. For example, a reactive power reference change in one unit should not only influence its own terminal voltage, but also other node voltages in an interconnected system. A power grid model is required to eliminate states x and inputs u. The power grid is composed of transmission lines and transformers, which connect all the units together. In Chapter 3, the small-signal model of an individual DFIG with an infinite bus system is derived. As the system is small and number of states is limited, the power grid is modelled with detailed derivative equations. Its transmission lines and transformers, 162

CH.6 Power System with Large Wind Power Installation

which connect all the units together, are modelled as series RL circuits. Small-shunt capacitance is modelled at each node (terminator of the unit) to convert the node voltages

Vds and Vqs to the state variables. However, this modelling method greatly increases
the number of state variables and is not efficient for a large system. As the high frequency phenomenon in the power grid is usually not interesting for researchers, the power grid can be modelled by algebraic equations. The changes in node voltages V can be calculated with the changes in injected currents i from admittance matrix as in (6.3):

i = Yn V

(6.3)

where Yn is the incremental admittance matrix of the power grid. Assuming that the power grid is a linear system, Yn is the same as the admittance matrix used in power flow calculation.

i are often state variables, or at least they can be expressed by state variables and inputs
as in (6.4):

i = Ci x + Dv V + Du u
Combining equations (6.4) and (6.3), the node voltages can be calculated as:

(6.4)

V = (Yn Dv )1 (Ci x + Du u)

(6.5)

Substituting (6.5) in (6.1) and (6.2), the interconnected state space model is completed:

_ x = [A + Bv (Yn Dv )1 Ci ]x + [Bu + Bv (Yn Dv )1 Du ]u y = [C + Dv (Yn Dv )1 Ci ]x + [Du + Dv (Yn Dv )1 Du ]u


(6.7).

(6.6) (6.7)

where only state variables x and inputs u exist on the right hand of equations (6.6) and

6.3.3 Low Frequency Oscillation Modes


The low-frequency oscillation modes of a four-machine system with 0% wind power penetration are shown in Figure 6-8. The excitation system of a synchronous generator introduces low-frequency damping problem, which is traditionally solved by installing a PSS on the synchronous generators. From the eigen values shown in Figure 6-8, it is clear that this system has poorly-damped local modes, which are oscillations between G1 and G2, and between G3 and G4. This system also has only one marginally stable inter-area mode, which is an oscillation between the pairs G1,G2 and G3,G4. The damping, frequency and participation factor of the low-frequency oscillation modes are listed in Table 6-1.

163

6.3 Small-signal Stability

The eigen values of a power system with 25% wind power penetration is calculated by replacing G4 with an aggregated wind power plant model in a multi-machine system. The low frequency oscillation modes are shown as in Figure 6-9. The aggregated wind power plant here is simulated as a single DFIG wind turbine with large power capacity. The aggregation of models of wind speed and mechanical parts is not taken into account in this thesis. Only the aggregation of electrical system is implemented. The damping, frequency and participation factor of the low-frequency oscillation modes are listed in Table 6-2. One of the local oscillation modes disappears, which is the oscillation between G3 and G4. From the participation factors of the inter-area mode in Table 6-2, it can be seen that the wind power plant does not participate in both the local mode and inter-area mode. By comparing the oscillation modes of the multi-machine power system with 0% wind power penetration and 25% wind power penetration, as shown in Figure 6-8 and Figure 6-9, some conclusions on the small signal stability of wind power are drawn: The wind power plant based on VSWTs does not participate in both the local mode and inter-area mode.
modes for d2asbeg 3 2.5 frequency Hz 2 1.5 1 0.5 0 -0.2

two local oscillations: G1-G2, G3-G4 inter-area oscillation

0.2

0.4 0.6 dampiing ratio

0.8

1.2

Figure 6-8. Low frequency oscillation modes of multi-machine power system with 0% wind power penetration.

The wind power plant introduces several low-frequency oscillation modes, such as active and reactive power controller modes, but they are well damped. It is also interesting to notice that the damping of the inter-area oscillation mode increases from 0.0012 to 0.0044, when the wind power plant replaces the thermal power plant G4.

164

CH.6 Power System with Large Wind Power Installation

Thus even though the wind power plant does not participate in the inter-area oscillation mode, it is still considered beneficial for the small-signal stability of the multi-plant system.

Table 6-1 Inter-area and local oscillation modes of 0% wind power penetration system No. Damp Freq(Hz) Participation Factors G1 0.9410 + 0.0500i 0.9756 + 0.0734i G2 0.6319 + 0.0119i 0.6557 + 0.0268i G3 0.9614 + 0.0574i 0.9967 + 0.0815i G4 0.6471 + 0.0081i 0.6715 + 0.0232i G3 0.7414 + 0.1630i 0.7405 + 0.1711i G4 0.9763 + 0.2104i 0.9753 + 0.2210i G1 0.8681 + 0.1776i 0.8672 + 0.1875i G2 0.9802 + 0.1908i 0.9794 + 0.2019i
modes for d2asbegDFIG 3 2.5 frequency Hz 2 1.5 DFIG controller 1 0.5 0 -0.2 local oscillation G1-G2 inter-area oscillation 0 0.2 0.4 0.6 dampiing ratio 0.8 1 1.2

Modes Inter-area oscillations between two pairs of synchronous generators G1+G2, G3+G4

0.0012

0.651

0.101

1.161

Local oscillations between G3 and G4

0.105

1.1453

Local oscillations between G1 and G2

Figure 6-9. Low frequency oscillation modes of the multi-machine power system with 25% wind power penetration.

Table 6-2. Inter-area and local oscillation modes of 25% wind power penetration system.

165

6.3 Small-signal Stability

No.

Damp

Freq(Hz)

0.0044

0.7195

0.1148

1.1486

Participation Factors G1 0.7562 + 0.1761i 0.7728 + 0.1983i G2 0.2230 + 0.0336i 0.2336 + 0.0463i G3 0.9464 - 0.0299i 0.9999 + 0.0145i G1 0.9961 + 0.0708i 0.9966 + 0.0818i G2 0.6089 + 0.0510i 0.6092 + 0.0645i

Modes Inter-area oscillations between two groups of synchronous generators G1+G2 and G3

Local oscillations between G1 and G2

However, the inter-area oscillation mode between the two groups of synchronous generators, G1+G2 and G3, still exists and is still marginally stable. Traditionally, this can be solved by the following methods: Decrease the power flow through the inter-area transmission line, or increase the strength of the transmission system;

Install a PSS on the synchronous generator G1,G2 or G3. The first method is usually expensive. For the second method, installing a PSS on the synchronous generator is complex because the electric power output or speed of the synchronous generator cannot be directly controlled. The output electric power of SG can only be controlled indirectly by changing the excitation voltage. Because there is considerable phase delay between the excitation voltage and electrical power, small-signal modelling of the whole system is required. And the on-line tuning of the PSS is necessary, as the change in grid topology will influence the small-signal model of the whole system. Moreover, the PSS also has a negative effect on the transient stability of the synchronous generators [Anderson and Fouad'03, Kundur'94, Rogers'00]. On the other hand, it is relatively easy to install a PSS onto the wind power plant, because the active power output of the VSWT can be directly controlled by utilising the inertia of the VSWT. The effectiveness of the PSS on the wind power plant will be shown in the next section.

6.3.4 Power System Stabiliser of Wind Power Plant


In previous section, it is found that the local oscillation mode between the wind power plant and the thermal power plant disappears. However, the inter-area oscillation mode between two groups of generators still exists. In order to improve the damping of the system or to limit the inter-area oscillation, a power system stabiliser is installed onto the wind power plant [Jauch'07, Jauch, et al.'07]. The stabiliser detects the power flow of the

166

CH.6 Power System with Large Wind Power Installation

inter-connection lines and controls its own active power output in reverse proportion to the power variations of the transmission lines. The active power is extracted from or stored on the wind turbine shaft temporarily using the inertia. A generic PSS of a wind park is shown in Figure 6-10. The change in the active power reference of the wind park can be distributed to individual wind turbines by the central controller of the wind park, as discussed in Chapter 5. In the PSS of the synchronous generator, the phase compensation is used to compensate the phase delay between the exciter input and generator electric torque. In the PSS of the wind park, the phase compensation is used to compensate the phase delay caused by the band-pass filter (washout block) and/or signal communication delay.

Two S 1+Two S

1+T1S 1+T2S

Figure 6-10. Generic power system stabiliser of wind power plant.

The PSS of the wind power plant used in the simulation is tuned with the following parameters.
Table 6-3. Parameters of PSS for wind power plant.

10 s 0.8 s 0.2 s PSS Gain 0.33 pu

The effect of installing a PSS onto the wind power plant can be seen in Figure 6-13. At 100s, one synchronous generator has a 0.2 pu power reference step-up, and the PSS on the wind power plant begins at 102s. The dynamic simulation results shown in Figure 6-11 demonstrate the effectiveness of the PSS of the wind power plant. The power oscillations on the tie lines between two group of synchronous generators damp out very fast due to the PSS on the wind power plant. The speed and power of the synchronous generator can be seen in Figure 6-12.

167

6.3 Small-signal Stability

Inter-area Power Flow with and without PSS 1000 Power (MW) 900 800 700 600 95 100 105 with PSS without PSS

110 115 120 time (s) Figure 6-11. Power flow of the inter-area tie line with and without PSS of wind power plant.

Synchronous Generator Power with and without PSS 0.9 Power(pu) 0.8 0.7 0.6 0.5 95 100 105 110 115 120 with PSS without PSS

Synchronous Generator Speed with and without PSS 1.005 Speed(pu) with PSS without PSS 1

0.995 95

100

105 time (s)

110

115

120

Figure 6-12. Power and speed of synchronous generator with and without a PSS on the wind power plant.

However, the PSS of a wind power plant adds more torque oscillations and increases the fatigue of the wind turbine itself, as shown in Figure 6-13. Another important point is that the PSS of a VSWT has to avoid the self-resonance frequency of the wind turbine, which is about 0.3Hz. It has to be noted that the steady-state speed decrease shown in Figure 6-13 is not due to the PSS of the wind power plant, but is caused by the change in aerodynamic power (wind speed).

168

CH.6 Power System with Large Wind Power Installation

Wind Park Power with and without PSS 1.5 Power (pu) with PSS without PSS 1

0.5 95

100

105

110

115

120

Wind Turbine Speed with and without PSS 0.88 Speed(pu) 0.87 0.86 0.85 100 with PSS without PSS

101

102 time (s)

103

104

105

Figure 6-13. Power and speed of the VSWT with and without a PSS on the wind power plant.

6.4 Limitation of Inertia Usage


As previously discussed, the inertia of the VSWT can be used for short-term primary frequency support and low-frequency oscillation damping. However, there are some fundamental limitations [Rawn and Lehn'08], such as speed limitation and converter power rating limitation. The optimum and 5% de-loaded operation curves are plotted in Figure 6-14. The speed limitations of Type C wind turbines 0.7 1.2 pu and Type D wind turbines 0.5 1.5 pu are also plotted. The optimum operation curve is based on the maximum power production mode instead of the maximum torque production mode, and the optimum tip-speed ratio is always larger than the tip-speed ratio of the maximum torque production.

169

6.4 Limitation of Inertia Usage

x 10

1.8

5%de-loaded Popt

1.6

1.4

active power (W)

1.2

transition from maximum power production to de-loaded mode

Paero @ Vw = 9m/s

0.8

min DFIG speed


0.6

max DFIG speed max PMG speed

min PMG speed


0.4

0.2

0.2

0.4

0.6

0.8 1 1.2 rotation speed (pu)

1.4

1.6

1.8

Figure 6-14. Maximum power production and de-loaded operation curves of a variable speed wind turbine, with speed limitation of Type C wind turbines 0.7pu to 1.2 pu and Type D wind turbines 0.5pu to 1.5pu.

The primary frequency support capability of Type C wind turbines using kinetic energy is plotted as Figure 6-15, where the results of wind speeds between 7m/s and 10m/s are shown. The capability is plotted as the maximum allowable supporting time versus the maximum output power increase , and is limited by both its converter power rating and minimum-allowed rotation speed. The possibility of over-speed is not considered a critical limitation here, because the speed can be limited by the blade pitch controller. The deloaded operation significantly improves the primary frequency support capability. For example, at 7m/s wind speed, without de-loaded operation the Type C wind turbine cannot be used to provide primary frequency support using its kinetic energy. With de-loaded operation, the wind turbine has certain primary frequency support capability. The capability of Type D wind turbines using kinetic energy is plotted as Figure 6-16. Type D wind turbines have larger primary frequency support capability than Type C wind turbines, because of the larger allowable speed range. In both figures, the influence of aerodynamic power change due to the change in its rotation speed is taken into account, but the influence of possible wind speed change during the primary frequency support period is omitted.

170

CH.6 Power System with Large Wind Power Installation

In Figure 6-15 and Figure 6-16, the capability curves are often composed of two parts. One part is a straight line which is determined by the converter power rating, and the other part is a curved shape which is determined by the minimum-allowed rotation speed. For power system stability application, the wind power plant can damp higher frequency (0.5Hz) oscillations more easily than lower frequency (0.1Hz) oscillations, because lower frequency oscillation damping requires more kinetic energy and causes larger speed deviation. Figure 6-15 and Figure 6-16 can also be used to estimate the low-frequency oscillation damping capabilities of Type C and Type D wind turbines. For example, a maximum-allowed support time of 10s at change is controlled to be less than 0.2pu ( Figure 6-16. De-loaded operation significantly improves the capabilities of Type C and Type D wind turbines when their kinetic energy is used for short-term primary frequency support and low-frequency oscillation damping. At low wind speeds, the critical limitation is the minimum-allowed rotation speed, while at high wind speed it is the converter power rating. At medium wind speeds (between 7m/s to 10m/s), both Type C (during de-loaded operation) and D wind turbines can use the kinetic energy for 10s-primary frequency support or for 0.1-Hz low frequency oscillation damping, if the power output changes are kept smaller than 0.2pu ( ). implies that the wind turbine can ). use its kinetic energy to damp out oscillations as low as 0.1Hz when its active power Some general conclusions can be drawn from analysing the results in Figure 6-15 and

6.5 Conclusion
In this chapter, the behaviour of a power system with 20% wind power penetration is studied, including the primary frequency response and small-signal stability. By only utilising the inertia for primary frequency support, in the first 10 seconds the frequency response with support from the wind power plant is better than the situation with only thermal a power plant. With a de-loaded wind power plant, the primary frequency response of the power system is improved for the whole range 0s-30s. For small-signal stability, the wind power plant using VSWTs does not participate in both the local mode and inter-area mode, and therefore the inertia can be used to damp out the existing low-frequency power system oscillations. The fundamental limitations of inertia usage of VSWTs for primaryfrequency support and low-frequency oscillation damping are also discussed.

171

6.5 Conclusion

0.8 5% de-loaded @ Vw = 7m/s Opt @ Vw = 7m/s

0.7

5% de-loaded @ Vw = 8m/s 0.6 Opt @ Vw = 8m/s

0.5 Opt @ Vw = 9m/s

5% de-loaded @ Vw = 9m/s

P (pu)

0.4 5% de-loaded @ Vw = 10m/s 0.3 Opt @ Vw = 10m/s 0.2

0.1

10

15 support time (s)

20

25

30

Figure 6-15. Primary frequency support capabilities of Type C wind turbines at wind speeds between 7m/s to 10m/s; the solid curves are the capability curves when the wind turbine is operating in maximum-power production mode, and the dotted curves are wind turbine in 5% de-loaded mode.
0.8 5% de-loaded @ Vw = 7m/s 0.7 Opt @ Vw = 7m/s

5% de-loaded @ Vw = 8m/s 0.6 Opt @ Vw = 8m/s

0.5

5% de-loaded @ Vw = 9m/s Opt @ Vw = 9m/s

P (pu)

0.4

5% de-loaded @ Vw = 10m/s

0.3 Opt @ Vw = 10m/s 0.2

0.1

10

15 support time (s)

20

25

30

Figure 6-16. Primary frequency support capabilities of Type D wind turbines at wind speeds between 7m/s and 10m/s; the solid curves are the capability curves when the wind turbine is operating in maximum-power production mode, and the dotted curves are wind turbine in 5% de-loaded mode.

172

CH.6 Power System with Large Wind Power Installation

6.6 References
[And03] [Ban06] P. M. Anderson and A. A. Fouad, Power system control and stability, ISBN0-471-23862-7 Wiley-Interscience, New York, 2003. H. Banakar, C. Luo, and B. T. Ooi, "Steady-state stability analysis of doublyfed induction generators under decoupled P-Q control," IEE ProceedingsElectric Power Applications, vol. 153, issue 2, Page(s). 300, 2006. E. A. A. Coelho, P. C. Cortizo, and P. F. D. Garcia, "Small-signal stability for parallel-connected inverters instand-alone AC supply systems," IEEE Transactions on Industry Applications, vol. 38, issue 2, Page(s). 533-542, 2002. L. Congwei, W. Haiqing, S. Xudong, and L. Fahai, "Research of stability of double fed induction motor vector control system," in The Fifth International Conference on Electrical Machines and Systems, 2001. ICEMS 2001. , 2001, Page(s). 1203-1206 R. G. de Almeida and J. A. P. Lopes, "Participation of Doubly Fed Induction Wind Generators in System Frequency Regulation," IEEE Transactions on Power Systems, vol. 22, issue 3, Page(s). 944, 2007. F. Gurbuz and E. Akpinar, "Stability Analysis of a Closed-Loop Control for a Pulse Width Modulated DC Motor Drive," Turkish Journal Electrical Engineering, vol. 10, issue 3, Page(s). 427-438, 2002. C. Jauch, "Transient and Dynamic Control of a Variable Speed Wind Turbine with Synchronous Generator," Wind Energy, vol. 10, issue 3, Page(s). 247, 2007. C. Jauch, T. Cronin, P. Sorensen, and B. Bak-Jensen, "A fuzzy logic pitch angle controller for power system stabilisation," Wind Energy, vol. 10, Page(s). 19-30, 2007. P. Karlsson, J. Bjornstedt, and M. Strom, "Stability of Voltage and Frequency Control in Distributed Generation Based on Parallel-Connected Converters Feeding Constant Power Loads," in 2005 European Conference on Power Electronics and Applications, EPE2005, 2005, Page(s). 10. P. Kundur, Power System Stability and Control, ISBN-0-07-035958-x, McGraw-Hill Professional, 1994. D. J. Lee and L. Wang, "Small-Signal Stability Analysis of an Autonomous Hybrid Renewable Energy Power Generation/Energy Storage System Part I: Time-Domain Simulations," IEEE Transaction on Energy Conversion, vol. 23, issue 1, Page(s). 311-320, 2008. J. Morren, S. W. H. de Haan, W. L. Kling, and J. Ferreira, "Wind turbines emulating inertia and supporting primary frequency control," IEEE Transactions on Power Systems, vol. 21, issue 1, Page(s). 433-434, 2006. B. Rawn and P. Lehn, "Wind rotor inertia and variable efficiency: fundamental limits on their exploitation for inertial response and power system damping," in European Wind Energy Conference 2008, EWEC2008, Brussels, Belgium, 2008.

[Coe02]

[Con01]

[Alm07]

[Gur02]

[Jau07]

[Jau07]

[Kar05]

[Kun94] [Lee08]

[Mor06]

[Raw08]

173

6.6 References

[Rog00]

G. Rogers, Power System Oscillations, ISBN-0-7923-7712-5, Kluwer Academic Pub, 2000.

174

CH.7 Conclusions and Recommendations

Chapter 7
Conclusions and Recommendations

7.1 Conclusions
The aim of this thesis is to improve the performance of wind farms and power grids with existing wind turbine concepts, both during normal operation and disturbances. Three wind turbine concepts are considered: Type A constant speed wind turbines, Type C doubly fed induction generators, and Type D permanent magnet generators with a full-scale voltage source converter. The research topics of this thesis are the following. Dynamic Modelling Dynamic models in a dq synchronously rotating reference frame are developed, including an average model and switching model for VSCs and CSCs. These models have been developed for simulation purposes such as flicker calculation, low-voltage ride through capability studies, etc. The dynamic simulations of three different dynamic models (i.e., different control and modelling approaches) are compared, which are an FOC with an average model, an FOC with a switching model, and a DTC with a switching model. The FOC with an average model is the fastest and is accurate enough for electro-mechanical simulation. The dynamic model of Type C wind turbines is verified through experiment result. Small-Signal Modelling The small-signal model is implemented in Matlab and is used for control design and stability analysis of variable speed wind turbines. Its accuracy is validated by comparing the step responses of the small-signal model and dynamic model of a single DFIG with an infinite bus system. The stability of a single DFIG integrated with an infinite bus system is analysed, which determines the stability margin of the closed-loop controlled DFIG. VSWTs during Unbalanced Grid Voltage Dips During unbalanced grid voltage, the most severe operation problem for variable speed wind turbines is not the transient over-current, but both the large electric torque oscillation, which causes wear and tear of the gearbox, and the large voltage ripple in the DC link of 175

7.1 Conclusions

back-to-back VSCs, which may decrease the lifetime of the DC capacitance. The dualsequence current controller is developed to control VSWTs during unbalanced grid voltage. Simulation of a DFIG during unbalanced grid-voltage dips shows that both the torque oscillation and DC voltage ripple can be limited satisfactorily by the dual-sequence current controller. VSWTs in a Low-Inertia Power Grid During classical power system operation, where synchronous generators dominate, the grid frequency is mainly related to active power, while grid voltage is mainly related to reactive power. However, in a low-inertia power grid, where there is no synchronous generator, the frequencies ! or f are at least transiently related to the reactive power load

Qload % ! f % , while voltage V is transiently related to the active power load

Pload % ! V &. Two control methods are proposed to operate the low-inertia power
grid with variable speed wind turbines. One method is to operate the system with a central control unit. A STATCOM is installed inside a DFIG wind farm with an HVDC LCC (thyristor bridges). The DC voltage is controlled by changing the fire angle of the thyristor rectifier. The DC capacitor of STATCOM is used to transiently balance the active power produced by the DFIGs and the active power transmitted through the HVDC LCC. The other method uses distributed units with inverse droop controllers to control the frequency and voltage in order to emulate the behaviour of the synchronous generator. Wind Park Integration Choices The three integration choicesan HVAC cable system, an HVDC LCC, and an HVDC VSCare compared with regard to their transmission costs. The total transmission costs include investment, energy loss and energy spilled. A Matlab program is developed to implement the comparison method, including the transmission capability calculation of HVAC cable systems, the power-loss calculation of HVAC and HVDC systems, etc. The influence different aspects have on the transmission capability, such as reactive power compensation, short-circuit ratio at point of common coupling, are evaluated; the costs of an expensive offshore platform are also taken into account. The critical distance between an HVAC 220kv cable and an HVDC VSC is found to be around 120km. Hybrid Wind Park Optimisation and Control A hybrid wind farm composed of both a constant speed wind turbine and a variable speed wind turbine is proposed in this thesis. The configuration of the hybrid wind park is optimised to have low levelised production cost, high reliability, an acceptable flicker level

176

CH.7 Conclusions and Recommendations

and LVRT capability. Variable speed wind turbines work as a distributed STATCOM, for which the reactive power capability is used to limit flicker and improve LVRT capability. A centralised reactive power controller of the hybrid wind farm is designed to control the hybrid wind farm as an integral unit. The central reactive power controller has several advantages when compared to the individual reactive power droop controller of a wind turbine: firstly, the total reactive power of the wind farm is controlled more precisely; secondly, the reactive power control strategy of the wind farm is more flexible, such as reaching the unity power factor at the point of common coupling (PCC), controlling voltage at the PCC, and limiting flicker. The central reactive power controller has two main functioning blocks: an anti-windup PI controller, and a dispatching function. The anti-windup PI controller prevents large oscillations in the output of the PI controller when the reactive power demand reaches the reactive power limitation. The dispatching function ensures that each wind turbine will output the reactive power proportional to its capabilities. Frequency Response of Power System with Large Wind Power Penetration Frequency responses of a power system with and without wind power penetration are compared. Without primary frequency support from the wind power plant, the power system with 20% wind power penetration has a frequency dip larger than a power system with 0% wind power; this will unnecessarily trigger the low-frequency load shedding. The large inertia of the wind turbines can be used to provide primary frequency support for a short period, for example 10s. However, as the rotation speed of VSWTs has to be return to the optimal rotation speed after 10 seconds, the frequency decreases and stabilises at the same value as it does without primary frequency support from wind power plant. For a system without an energy storage device, one way to improve the frequency response between 10s and 30s is to de-load the wind turbine, i.e., to operate the wind turbine in power-reserve mode. Local and Inter-Area Oscillation of Power System with Large Wind Power Penetration The small-signal stabilities of a four-machine power system with and without wind power penetration are analysed. From the Eigen values, it can be seen that the wind power plant does not participate in both the local mode and inter-area oscillation mode. One local mode disappears if the wind power plant replaces one thermal power plant. However, the interarea oscillation between two groups of synchronous generators does not disappear, and is still marginally stable. A power system stabiliser is installed on the wind power plant,

177

7.2 Recommendations

which detects the power flow over the inter-connection lines, and damps the active power oscillations over the transmission lines. The energy is extracted from or stored on the wind turbine shaft temporarily using the inertia of the wind turbine.

7.2 Recommendations
The following issues related to this thesis are proposed for future research. The control strategy of wind farms and DFIGs with HVDC LCC integration has to be investigated for the possibility of operating without using the central control STATCOM unit as power balancing unit. The voltage source converter is sensitive to over-current, thus over-current protection needs to be implemented in the inverse frequency and voltage-droop controller. A complete database of price and parameters of electric components, including cable, transformers, converters, etc, is required. From which comparisons between HVAC, HVDC LCCs, HVDC VSCs can be made for any voltage level and integration distance. The optimisation of hybrid wind farms can be extended to consider the cases of HVDC connections inside wind farms or between wind farms and the power grid. The wind power plant PSS in this thesis only uses the wind turbine inertia to damp out the low-frequency oscillation of the power system. The reactive power capability of the wind power plant is recommended for future research on PSSs, which is similar to using SVCs to improve small-signal stability of power system.

178

Appx.A List of Symbols and Abbreviations

Appendix. A
List of Symbols and Abbreviations A. 1 Symbols
Awb

Cp Cdc Cdr
Dload
Scaling factor of wind speed in Weibull distribution Wind turbine blade pitch angle Aerodynamic power coefficient DC capacitance Damping coefficient of shaft Load frequency coefficient Inertia Constant Equivalent inertia constant of the power system a, b, c phase currents d, q components of currents DC current Gearbox ratio Inertias of wind turbine and generator Form parameter of wind speed Weibull distribution Stiffness of shaft DC inductance Stator, rotor and magnetizing inductances Stator and rotor leakage inductance Wind turbine tip speed ratio Stator, rotor and magnetizing fluxes d, q components of fluxes Active and reactive power
[web turns] [web turns]

[deg] [F ]

[N m=rad]

H Hsys
ia ; ib ; ic id ; iq Idc
gear Jwt ; Jgen

[A] [A] [A]

[kg m2 ] [N m=rad] [H] [H] [H]

kwb Kdr Ldc Ls ; Lr ; Lm Lls ; Llr s ; r ; m d ; q P; Q

[W ]

179

Appx.A List of Symbols and Abbreviations

Pgen ; Ps ; Pr ; Pconv Ptie line Qgen ; Qs ; Qr ; Qconv Rdc Rdroop Rs ; Rr R wt Taero ; Tshaf t ; Te Ts va ; vb ; vc vd ; vq Vw Vdc !wt ; !gen

Generator, stator, rotor and grid VSC active power Active power flow on the inter-area transmission line Generator, stator, rotor and VSC reactive power DC resistance Droop ratio of primary frequency Stator, rotor resistance Air density Wind turbine radius Twisted angle of the turbine shaft Aerodynamic, shaft and electric torque Discrete time of pulse width modulation a, b, c phase voltages d, q components of voltages Effective wind speed DC voltage Rotation speeds of wind turbine and generator Probability of wind speed in Weibull distribution

[W ] [W ] [V ar] [] [] [kg=m3 ] [m] [rad] [N m] [s] [V ] [V ] [m=s] [V ] [rad=s]

A. 2 Subscripts
s r

Stator components Rotor components Grid components Synchronously rotating reference frame components Stationary reference frame components

g d; q ;

A. 3 Notations
The variables in the equations use the following nottations

x(t) X

time-dependent signal RMS value

180

Appx.A List of Symbols and Abbreviations

~ X

Vector

A. 4Abbreviations
AC CSC CSWT DFIG DG DPC DTC FOC GRA HVAC HVDC IGBT LCC LPC LTC LVRT PCC PLL PMG PSS pu PWM SCIG SG SPWM STATCOM SVC Alternating Current Current Source Converter Constant Speed Wind Turbine Doubly Fed Induction Generator Distributed Generation Direct Power Controller Direct Torque Controller Field Oriented Controller Generation Ratio Availability High Voltage Alternative Current High Voltage Direct Current Insulated Gate Bipolar Transistor Line Commutated Converter Levelised Production Cost Levelised Transmission Cost Low Voltage Ride Through Point of Common Coupling Phase Locked Loop Permanent Magnet Generator Power System Stabilizer Per Unit Pulse Width Modulation Squirrel Cage Induction Generator Synchronous Generator Sinusoidal Pulse Width Modulation Static Compensator Static Var Compensator

181

Appx.A List of Symbols and Abbreviations

SVPWM UGR VSC VSWT WRIG

Space Vector Pulse Width Modulation Unavailable Generation Ratio Voltage Source Converter Variable Speed Wind Turbine Wound Rotor Induction Generator

182

v grid

f grid

f grid

iqr

idr

Te _ ref
idr _ ref vd , vq iqr _ ref vd , vq

Te

Vdc

Tadd

idref

Pref

Vdc ref

Qs
iqref

Q
Qref

Qs _ ref

Appendix. B

Figure B-1 Structure of Type C wind turbine

Appx.B Structures of Closed Loop Controlled Type C and D VSWT

Structures of Closed Loop Controlled Type C and D VSWT

183

184
f grid

!m
Vdc

isd isq
idref
P
Pref

Te _ ref

Te

Vdc ref

Tadd

id _ ref

Appx.B Structures of Closed Loop Controlled Type C and D VSWT

r
iqref

r _ ref

vd , vq

vd , vq

Q
Qref

iq _ ref

f grid

Figure B-2 Structure of Type D wind turbine

Appx.C Wind Turbine Parameters

Appendix. C
Wind Turbine Parameters C.1 Type A Wind Turbine
Table C-1 Type A Wind Turbine Parameters Pnom
Jwt

nominal power inertia of wind turbine inertia of generator gearbox ratio diameter of rotor blades stator leakage inductance rotor leakage inductance mutual inductance stator resistance rotor resistance

2:75
1:26 10 2:4 102 70:65
7

[MW ]
[kg m2 ] [kg m2 ]

Jgen

Gear ratio
Dwt Lls Llr Lm Rs Rr

92 1:18 104 3:08 104 9:1 103 1:2 103 5:2 103

[m] [H] [H] [H] [] []

C.2 Type C Wind Turbine


Table C-2 Type C Wind Turbine Parameters Pnom
Jwt Jgen

nominal power inertia of wind turbine inertia of generator gearbox ratio diameter of rotor blades stator leakage inductance

2:75 1:26 107 2:4 102 70:65 92 1:18 104

[MW ] [kg m2 ] [kg m2 ]

Gear ratio
Dwt Lls

[m] [H]

185

C.3 Type D Wind Turbine

Llr Lm Rs Rr

rotor leakage inductance mutual inductance stator resistance rotor resistance DC capacitance of VSC converter resistance

3:08 104 9:1 10


3

[H] [H] [] [] [F ] [] [H]

1:2 103 5:2 10


3

Cdc
Rconv Lconv

1 103 1:25 102 1 104 H


2:5 102 ; 1:55 101 3:2 104 ; 3:2 103 2 102 ; 3:2 101 2 101 ; 3:2 101 2 104 ; 2 105 1 102 ; 1 101 1 102 ; 1 101

converter inductance
electric torque Te control stator reactive power Qs control d-axis rotor current id r control q-axis rotor current iq r control DC voltage Vdc control d-axis grid VSC current id conv control q-axis grid VSC current iq
conv

kp ; ki kp ; ki kp ; ki kp ; ki kp ; ki kp ; ki kp ; ki Dp ; Dq kp ; ki

control

grid frequency and voltage droop ratio phase locked loop

4%
5 101 ; 1 103

C.3 Type D Wind Turbine


Table C-3 Type D Wind Turbine Parameters Pnom
Jwt Jgen Dwt Ls Lm f Rs

nominal power inertia of wind turbine inertia of generator diameter of rotor blades stator inductance mutual inductance magnetic flux stator resistance

2:75 1:26 10 2:4 104 92 3 10


3 7

[MW ] [kg m2 ] [kg m2 ] [m] [H ] [H ] [web turns] []

9:1 103 3:8 1:47 102

186

Appx.C Wind Turbine Parameters

Cdc
Rconv Lconv kp ; ki kp ; ki kp ; ki kp ; ki kp ; ki Dp ; Dq kp ; ki

DC capacitance of VSC converter resistance

1 103 1:25 10
2

[F ] [] [H ]
1

converter inductance
d-axis rotor current id r control q-axis rotor current iq r control DC voltage Vdc control d-axis grid VSC current id conv control q-axis grid VSC current iq
conv

1 104 H
6:8 10 ; 3:4 10
0

1:45 103 ; 2:9 101 2 104 ; 2 105 1 102 ; 1 101 1 102 ; 1 101

control

grid frequency and voltage droop ratio phase locked loop

4%
5 101 ; 1 103

187

C.3 Type D Wind Turbine

188

Appx.D Two Area Four Machine System

Appendix. D
Two Area Four Machine System D.1 Topology

Figure D-1 Single line diagram - two area four machine system

D.2 Bus Data


All the parameters are converted to per unit values, with nominal power 100MVA and voltage ratings as Table D-1.
Table D-1 Bus data Bus Number 1 Bus Type PV (generator bus) Rated Voltage (kV) 22.0 P 7.0 Q 1.143

189

D.3 Transmission Line Data

2 3 4 10 11 12 13 14 20 101 110 120

PV (generator bus) PQ (load bus) PQ (load bus) PQ (load bus) PV (generator bus) PV (generator bus) PQ (load bus) PQ (load bus) PQ (load bus) Swing (infinite bus) PQ (load bus) PQ (load bus)

22.0 230.0 115.0 230.0 22.0 22.0 230.0 115.0 230.0 230.0 230.0 230.0

7.0 0.0 9.76 0.0 7.0 7.0 0.0 9.76 0.0

1.784 0.0 1.0 0.0 1.143 2.0 0.0 1.0 0.0

0.0 0.0

0.0 0.0

D.3 Transmission Line Data


Table D-2 Transmission line data from bus 1 2 3 3 3 3 10 11 12 13 13 13 to bus 10 20 4 20 101 101 20 110 120 101 101 14 R 0.0 0.0 0.0 0.001 0.011 0.011 0.0025 0.0 0.0 0.011 0.011 0.0 0.001 0.0025 X 0.0167 0.0167 0.005 0.01 0.110 0.110 0.025 0.0167 0.0167 0.110 0.110 0.005 0.01 0.025 B 0.0 0.0 0.0 0.0175 0.1925 0.1925 0.0437 0.0 0.0 0.1925 0.1925 0.0 0.0175 0.0437

13
110

120
120

190

Appx.D Two Area Four Machine System

D.4 Synchronous Generator and Excitation Data


Table D-3 Synchronous generator data Base Power generator Base Power stator leakage reactance stator resistance d-axis steady-state reactance d-axis transient reactance d-axis sub-transient reactance d-axis transient time constant d-axis transient time constant q-axis steady-state reactance q-axis transient reactance q-axis sub-transient reactance q-axis transient time constant q-axis transient time constant inertia constant 900 0.2 0.0 1.8 0.3 0.25 8.0 0.03 1.7 0.55 0.24 0.4 0.05 6.5
[MV A]

Table D-4 Excitation system Type exciter type time constant time constant time constant exciter gain exciter gain input maximum voltage input minimum voltage output maximum voltage output minimum voltage AC4a 0.015 10.0 1.0 200 0.0 10 -10 5.64 -4.53

191

D.4 Synchronous Generator and Excitation Data

Figure D-2 AC4a excitation system

192

List of Publications

List of Publications
1. Y. Zhou, P. Bauer, J. A. Ferreira, Full Order Small Signal Modeling of Doubly Fed Induction Generator with Field Oriented Current Controller, under second review, IEEE Transaction on Power Electronics, Special Issue on Modeling and Advanced Control in Power Electronics, 2009. 2. Y. Zhou, P. Bauer, J. A. Ferreira, J. Pierik, Operation of Grid-Connected DFIG under Unbalanced Grid Voltage Condition, IEEE Transaction on Energy Conversion, Vol. 24, Issue 1, Pages 240-246, March 2009. Y. Zhou, P. Bauer, J. A. Ferreira, J. Pierik, Integration of Large Offshore Wind Farm Doubly Fed Induction Generators with Classical HVDC, Acta Electrotechnica et Informatica, Vol. 9, No 2, Pages 8-14, April 2009. 4. Y. Zhou, P. Kumar, P. Bauer, J. A. Ferreira, Optimization of a Hybrid Wind Park through a Design Approach - Progressive Design Methodology, IPEMC 2009 conference, Beijing, May 17-20, 2009. Y. Zhou, J. Morren, P. Bauer, J. A. Ferreira, Wind Park and Grid Integration a Grid Friendly Wind Power Plant", 2008 Global Wind Power conference, Beijing, October 29-31, 2008. 6. Y. Zhou, P. Bauer, J. A. Ferreira, Reactive power limitation of doubly fed induction generator and its application on coordinated hybrid wind farm control, 2008 EWEC conference, Brussels Expo, Belgium, March 31 April 3, 2008. Y. Zhou, P. Bauer, J. A. Ferreira, Grid-connected and islanded operation of a hybrid power system, 2007 Power Africa conference, Johannesburg, South Africa, July 16 - 20, 2007. Y. Zhou, P. Bauer, J. A. Ferreira, J. Pierik, Control of doubly fed induction generator to ride through unsymmetrical voltage dip, 2007 38th PESC conference, Orlando, USA, June 17 21, 2007. 9. Y. Zhou, P. Bauer, J. A. Ferreira, J. Pierik, Aggregated Models of Offshore Wind Farm Components for Grid Study, 2006 37th PESC conference, Jeju,

3.

5.

7.

8.

Korea, June 18 22, 2006. 10. Y. Zhou, P. Bauer, J. A. Ferreira, J. Pierik, New Thyristor Bridge Models for Dynamic Simulation of Grid Integration of Offshore Wind Farms, 2006 International PCIM Conference, Nrnberg, Germany, May 30 June 1, 2006.

193

11. Y. Zhou, P. Bauer, J. A. Ferreira, J. Pierik, Wind Park and Grid Integration Issues, 2006 IEEE Young Researchers Symposium in Electrical Power Engineering, Ghent University, Ghent, Belgium, April 27-28, 2006.

194

Summary

Summary
Grid Integration of Wind Power: from Individual Wind Turbine to Wind Park as a Power Plant
PhD Thesis by Yi Zhou As power capacities of single wind turbine, single wind park and total wind power installation are continuously increasing, the wind power begins to challenge the safety operation of the power system. This thesis focuses on the grid integration aspects such as the dynamic behaviours of wind power during disturbances, and dynamic behaviours of power system with large wind power integration. The work in this thesis is in a down-up approach, starting with concepts for individual wind turbines, including control and modelling, followed by a conceptual wind park design and control, and finally on the highest level, wind power plant support aimed at improving power system performance.

Individual Wind Turbine


In order to analyze the interactions between wind power and power system, the first step is to develop appropriate models of individual wind turbines, including dynamic model and small-signal model. Dynamic model is used to study the integration problems of wind turbines, such as flickers, harmonics, transient stability, and low-voltage ride through capability. Small-signal model is used to study integration problems, such as stability of the closed-loop controller, and local and inter-area power system oscillations of the power system during large wind-power penetration. In this thesis, dynamic models are developed in Matlab/Simulink, including field-oriented controller with an average model, a field-oriented controller with a switching model, and a direct torque controller with a switching model. The results of dynamic simulations with the above different control methods and modelling approaches are compared. The dynamic model of a wind turbine with a doubly fed induction generator (DFIG) and a field-oriented controller during a grid-voltage dip is verified by experiments. By linearising the dynamic model around an operating point, the small-signal model is derived in an analytical way in this thesis. The small-signal model of a wind turbine with DFIG is validated by comparing its time-step responses with the dynamic model. The

195

stability of a single DFIG wind turbine integrated into an infinite bus system is analysed with its small signal model, and the stability margin is determined from the analysis.

Wind Park
The wind park has to be connected to the main power grid. The possible integration options are an high voltage alternating current (HVAC) system, a classical high voltage direct current (HVDC) system using a line-commutated converter (HVDC LCC), and an HVDC system using a self-commutated converter (HVDC VSC). The ultimate choice is made in this thesis by comparing the transmission costs of the three systems, including both investment and energy loss. Wind turbine technology has evolved from constant speed wind turbines (CSWTs) to variable speed wind turbines (VSWTs). This development has mainly been caused by the fact that VSWTs have better power quality and low-voltage ride through (LVRT) capability. The fact that VSWTs are able to increase energy production is not deterministic for the evolution of wind turbine technology. Inspired by the above facts, a hybrid wind park composed of both CSWTs and VSWTs is proposed in this thesis, as this combination can provide similar power quality and LVRT capability with better cost benefits. VSWTs in the hybrid wind park work as distributed static compensators (STATCOMs) using reactive power to limit flicker and improve LVRT capability. A multi-objective optimisation method is used to determine the optimal configuration of a hybrid wind park which has low levelised production cost, high power quality, high reliability and high LVRT capability. In order to control this hybrid wind park as an integral wind power plant, a central controller of the hybrid wind park using an anti-windup PI controller and dispatching function is designed in this thesis.

Power System with Large Wind Power Integration


In a power system, the frequency deviation is believed to be related with the active power balance. This phenomenon is due to the swing equation of synchronous generator (SG). However, with large wind power integration, the power systems frequency response will be different, because the VSWTs rotation speed is completely decoupled with power grids electrical frequency by using the power electronic converter. The influence of 20% wind power installation on the frequency response of the power system is studied in this thesis. Small-signal stability is the ability of the power system to maintain synchronism when subjected to small disturbances. The small-signal instability of a traditional power system appears as poorly damped local and inter-area oscillations between synchronous generators, which are often caused by exciter and weak grid connection of the generators. As both the

196

Summary

structure and operation of wind power plants are very different from classical power plants, the small-signal stability of a power system with 20% wind power installation is analyzed quantitatively in this thesis by using the small-signal model of DFIG wind turbine. The kinetic energy stored on the shaft of the wind turbine is used to provide short-term primary frequency support and power-system oscillation damping, and the limitations of kinetic energy usage are analyzed.

197

198

Samenvatting

Samenvatting
Grid Integration of Wind Power: from Individual Wind Turbine to Wind Park as a Power Plant
Proefschrift van Yi Zhou Het vermogen van windturbines, windparken en windenergie-installaties neemt

voortdurend toe. Daardoor neemt het effect van windenergie op het elektrische systeem toe en moeten windturbines en windparken aan strengere eisen van de netbeheerders voldoen. Dit proefschrift richt zich op de aspecten van het integreren van windenergie in het elektriciteitsnet, zoals het dynamisch gedrag van windvermogen tijdens windvlagen en tijdens storingen in het elektriciteitsnet. Ook wordt het dynamisch gedrag van het elektriciteitsnet met een hoog aandeel windenergie onderzocht. Het proefschrift volgt een bottom-up benadering, te beginnen met concepten voor individuele windturbines, inclusief regeling en modellering. Daarna volgt het ontwerp van een conceptueel windpark plus regeling, en ten slotte, op het hoogste niveau, volgt de ondersteuning van elektriciteitscentrales door windenergie, gericht op het stabiliseren van het elektriciteitsnet.

Individuele windturbine
De eerste stap in het analyseren van de interactie tussen windenergie en het elektriciteitsnet is het ontwikkelen van geschikte modellen voor individuele windturbines, zowel lineair (voor regelaarontwerp) als niet-lineair (voor evaluatie van transienten). De problemen die voorkomen bij de integratie van windturbines, zoals flikker, harmonischen, stabiliteit en het vermogen voor een Low-Voltage Ride Through, LVRT (in bedrijf blijven tijdens een spanningsdip), worden bestudeerd met behulp van het dynamisch model. Het gelineariseerde model wordt gebruikt voor regelaarontwerp en stabiliteitsanalyse van de closed-loop regeling, en de analyse van een grote hoeveelheid windvermogen op het hoogspanningsnet. De dynamische modellen in dit proefschrift zijn ontwikkeld in Matlab/Simulink, inclusief een veldgeorinteerde regeling op basis van een gemiddeld model, een veldgeorinteerde regeling met instantaan model (met schakelende componenten), als en een direct torque (DTC) regeling met instantaan model. De simulatieresultaten van de bovengenoemde

199

regelmethoden en modellen zijn vergeleken. Het dynamisch model van een type windturbine met dubbelgevoede asynchrone generator (DFIG) en een veldgeorinteerde regeling tijdens een spanningsdip in het net is geverifieerd door vergelijking met experimenten. Voor een windturbine met DFIG is het gelineariseerde model geverifieerd door het vergelijken van de tijdstapresponsie met het dynamisch model.

Windpark
Het windpark op zee moet worden verbonden met het hoogspanningsnet. Hiervoor kan gebruik worden gemaakt van een AC-verbinding (HVAC), een klassieke stroombron HVDC verbinding op basis van thyristoren (HVDC LCC), of een spanningsbron HVDC verbinding op basis van IGBTs (HVDC VSC). In dit proefschrift wordt de uiteindelijke keuze gebaseerd op de totale transportkosten, dus zowel de investeringskosten als het energieverlies. De technologie van windturbines is gevolueerd van turbines met constant toerental (CSWTs), naar turbines met een variabel toerental (VSWTs). Deze ontwikkeling is in gang gezet door het feit dat VSWTs een betere regelgedrag hebben en in principe beter aan de neteisen kunnen voldoen (met name LVRT). In dit proefschrift wordt een hybride windpark voorgesteld, bestaande uit zowel CSWTs en VSWTs. Deze combinatie heeft een vergelijkbare arbeidsfactor en capaciteit voor LowVoltage Ride Through, met een potentieel betere kosteneffectiviteit. In het hybride windpark werken de VSWTs als gedistribueerde blindstroomregelaars (STATCOMs), waarmee reactief vermogen gebruikt wordt om flikkering te verminderen en de LVRT capaciteit te verbeteren. De optimale configuratie binnen het hybride windpark is bepaald door middel van een multi-objective optimalisatiemethode. Hierin is geoptimaliseerd naar lagere productiekosten, een gunstige arbeidsfactor, hoge betrouwbaarheid en een hoge LVRT bekwaamheid. Om dit hybride windpark als integrale windenergiecentrale te regelen, is een centrale regeling ontworpen welke gebruik maakt van een anti-windup PI regelaar en een distributie functie.

Elektriciteitsnet met een hoge mate van windenergie integratie


In het elektriciteitsvoorzieningsysteem is de frequentievariatie gerelateerd aan de vermogensbalans van het actief vermogen. Dit is het gevolg van de aanwezige roterende massa (rotatietraagheid) van alle direct aan het net gekoppelde synchrone en andere generatoren. Echter, een relatief groot aandeel windenergie in het net, kan de responsie bij

200

Samenvatting

verstoringen van de vermogensbalans veranderen. Dit wordt veroorzaakt door het ontbreken van een vaste koppeling tussen de netfrequentie en het toerental van VSWT. De invloed van 20% windvermogen op de frequentierespons van het elektriciteitsnet is in dit proefschrift onderzocht. Klein-signaal stabiliteit in een traditioneel hoogspanningsnet houdt verband met de demping van verschillende oscillatiemodes, dat benvloed worden door de snelle regeling van de opwekker van synchrone machines (spanningsregeling) en een hoge waarde netreactantie. Aangezien de structuur en werking van windparken erg verschilt van klassieke elektriciteitscentrales, is de klein-signaal stabiliteit van een elektriciteitsnet met 20% windenergie kwantitatief geanalyseerd in dit proefschrift. Hierbij is gebruik gemaakt van het gelineariseerde model van de windturbine met DFIG. De kinetische energie die is opgeslagen in de as van de windturbine wordt gebruikt om de netfrequentie gedurende korte tijd te ondersteunen, en om oscillaties in het elektriciteitsnet te dempen. De fundamentele beperkingen hiervan zijn geanalyseerd.

201

202

Grid Integration of Wind Power: from Individual Wind Turbine to Wind Park as a Power Plant

Matlab/Simulink 20%

203

CSWTVSWT LVRT STATCOM

20 20%

204

Curriculum Vitae

Curriculum Vitae

Yi Zhou was born in Hefei, China on October 7, 1973. In 1995, he received the B.Sc degree from the Department of Electrical Power System of Zhejiang University in Hangzhou, China. Since 1995, he had worked seven years in Anhui Electric Power Company, branch of State Grid Corporation of China. His responsibilities include steady-state power flow analysis, transient stability analysis, and short and medium-term power system planning. In 2002, He was promoted as the head of power system stability group. In 2004, he received the M. Sc degree from the Department of Electrical Power Engineering of Delft University of Technology in the Netherlands. Since 2005, he started working towards the PhD degree at the Delft University of Technology, in the Electric Power Processing group. The objective of his PhD research is to analyze the dynamics of wind power and power system with large scale wind power integration. It includes dynamic and small signal modeling of various concepts of wind turbines, control of wind turbines and integral wind farm, analysis and control of power system with large scale wind power. His research interests include power system analysis and grid integration of renewable energy.

205

206

You might also like