You are on page 1of 110

Power Systems

Adel Abdelbaset
Yehia S. Mohamed
Abou-Hashema M. El-Sayed
Alaa Eldin Hussein Abozeid Ahmed

Wind Driven Doubly


Fed Induction
Generator
Grid Synchronization and Control
Power Systems
More information about this series at http://www.springer.com/series/4622
Adel Abdelbaset Yehia S. Mohamed

Abou-Hashema M. El-Sayed
Alaa Eldin Hussein Abozeid Ahmed

Wind Driven Doubly Fed


Induction Generator
Grid Synchronization and Control

123
Adel Abdelbaset Abou-Hashema M. El-Sayed
Faculty of Engineering, Electrical Faculty of Engineering, Electrical
Engineering Department Engineering Department
Minia University Minia University
El-Minia El-Minia
Egypt Egypt

Yehia S. Mohamed Alaa Eldin Hussein Abozeid Ahmed


Faculty of Engineering, Electrical Faculty of Engineering, Electrical
Engineering Department Engineering Department
Minia University Al-Azher University
El-Minia Qena
Egypt Egypt

ISSN 1612-1287 ISSN 1860-4676 (electronic)


Power Systems
ISBN 978-3-319-70107-3 ISBN 978-3-319-70108-0 (eBook)
https://doi.org/10.1007/978-3-319-70108-0
Library of Congress Control Number: 2017957658

Mathematics Subject Classification (2010): 111, 000

© Springer International Publishing AG 2018


This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part
of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations,
recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission
or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar
methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this
publication does not imply, even in the absence of a specific statement, that such names are exempt from
the relevant protective laws and regulations and therefore free for general use.
The publisher, the authors and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the
authors or the editors give a warranty, express or implied, with respect to the material contained herein or
for any errors or omissions that may have been made. The publisher remains neutral with regard to
jurisdictional claims in published maps and institutional affiliations.

Printed on acid-free paper

This Springer imprint is published by Springer Nature


The registered company is Springer International Publishing AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Contents

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Book Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Scope of the Book . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2 Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 7
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 7
2.2 Synchronous Generators Driven by a Fixed Speed Turbine . . . . .. 8
2.2.1 Wound Field Synchronous Generator Driven by a Wind
Turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 8
2.2.2 Permanent Magnet Synchronous Generator (PMSG)
Driven by a Wind Turbine . . . . . . . . . . . . . . . . . . . . . . .. 9
2.3 Induction Generators Driven by a Variable Speed
Wind Turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 9
2.3.1 Squirrel-Cage Induction Generator (SCIG) Driven by a
Wind Turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 9
2.3.2 Doubly Fed Induction Generator (DFIG) Driven by a
Wind Turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.4 Modeling of a Wind Turbine-Generator System . . . . . . . . . . . . . . 12
2.4.1 Aerodynamic Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.4.2 Drive Train Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.4.3 DFIG Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.4.4 Power Converter Modeling . . . . . . . . . . . . . . . . . . . . . . . . 14
2.5 Control Strategies for a Wind Turbine-Generator System . . . . . . . 14
2.5.1 Pitch Angle Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.5.2 Maximum Power Point Tracking Control . . . . . . . . . . . . . 15
2.5.3 DFIG Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.6 Power Converter Topologies for a Wind Turbine-Generator
System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.6.1 Multi-level Converters . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.6.2 Matrix Converters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.7 DFIG Grid Synchronization . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

v
vi Contents

3 A Modified MRAS Observer for Sensorless Control of a Wind


Driven DFIG Connected to Grid . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.2 Description of the System Under Study . . . . . . . . . . . . . . . . . . . . 22
3.3 Dynamic Modeling of a Wind Turbine DFIG System . . . . . . . . . . 22
3.3.1 Aerodynamic Wind Turbine Model . . . . . . . . . . . . . . . . . . 22
3.3.2 Dynamic Model of a DFIG Taking Iron Losses and
Magnetic Saturation into Consideration . . . . . . . . . . . . . . . 24
3.3.3 DC Link Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.4 Vector Control of a DFIG for Grid-Connected Operations . . . . . . 27
3.5 Speed Estimation Based on the Modified MRAS Observer . . . . . . 30
3.6 Configuration of Complete System . . . . . . . . . . . . . . . . . . . . . . . 31
3.7 Results and Discussions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4 Grid Synchronization Enhancement of a Wind Driven DFIG Using
Adaptive Sliding Mode Control . . . . . . . . . . . . . . . . . . . . . . . . . . .. 41
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 41
4.2 Description of the DFIG System . . . . . . . . . . . . . . . . . . . . . . . .. 41
4.3 Dynamic Model of a DFIG Taking Iron Losses into
Consideration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
4.4 Direct Relationship Between Stator Voltage and Rotor Voltage . . . 42
4.4.1 External Disturbances and Parametric Uncertainty . . . . . . . 44
4.5 Design of Proposed Adaptive Sliding Mode Control . . . . . . . . . . . 45
4.6 Complete System of a Wind Generation with Proposed SMC . . . . 49
4.7 Results and Discussions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
5 Adaptive Sliding Mode Control for Grid Synchronization of a
Wind Driven DFIG Under Unbalanced Grid Voltage . . . . . . . . . ... 57
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ... 57
5.2 DFIG Model During Grid Synchronization Under Unbalanced
Grid Voltage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ... 57
5.3 Positive Sequence Controller Design . . . . . . . . . . . . . . . . . . . ... 60
5.3.1 Controlled Plant Analysis . . . . . . . . . . . . . . . . . . . . . . ... 60
5.3.2 Direct Relationship Between Positive Sequence Stator
and Rotor Voltages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
5.3.3 Parametric Uncertainty . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
5.3.4 Adaptive Sliding Mode Control Design . . . . . . . . . . . . . . . 62
5.4 Negative Sequence Controller Design . . . . . . . . . . . . . . . . . . . . . 66
5.5 Configuration of the Complete Control Scheme . . . . . . . . . . . . . . 68
5.6 Results and Discussions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
6 Conclusions and Suggestions for Future Work . . . . . . . . . . . . . . . . . 81
6.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
6.2 Suggestions for Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
Contents vii

Appendix A: Parameters of DFIG Wind Turbine System and


Laboratory Measurement of Magnetizing Inductance . . . . 83
Appendix B: Adaptation Mechanisms and Stability of MRAS . . . . . . . . 85
Appendix C: Parameters of DFIG as Listed in Reference [135] . . . . . . . 87
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
Symbols

MRAS Model reference adaptive system


DFIG Doubly Fed induction generator
GSC Grid side converter
RSC Rotor side converter
SFO Stator flux-oriented
SVO Stator voltage-oriented
SMC Sliding mode control
PWM Pulse width modulation
Vw Wind speed (m/s)
xT Turbine speed (rad/s)
R Rotor radius of wind turbine (m)
q Air density (Kg/m3)
JT Total inertia
TT ,Te Aerodynamic, electromagnetic torque
B Damping coefficient
Vds ; Vqs d-q axis stator voltage components
Vdr ; Vqr d-q axis rotor voltage components
Vas ; Vbs Stationary axis stator voltage components
Vgd ; Vgq d-q axis grid side converter voltage components
ids ; iqs d-q axis stator current components
idr ; iqr d-q axis rotor current components
idm ; iqm d-q axis magnetizing current components
idFe ; iqFe d-q core-loss current components
igd ; igq d-q axis grid side converter current components
ias ; ibs Stationary axis stator voltage components
p ¼ d=dt  Differential operator
r ¼ 1  L2m ðLs Lr Þ Leakage coefficient
Rs ; Rr ; RFe Stator, rotor, core-loss resistances
Lls ,Llr Stator and rotor leakage inductances

ix
x Symbols

Lr ,Ls Stator, rotor self-inductance


Lm Magnetizing inductance
kdm ; kqm d-q axis magnetizing flux components
kam ; kbm Stationary axis magnetizing flux components
km Magnetizing flux vector
xe ; xr Supply, rotor angular frequency
Ke ,Kh Hysteresis, eddy current coefficients
np Number of pole pairs
Vdc D-C link voltage
c D-C link capacitance
i1 ; i2 Grid and rotor converter D-C currents
Pg ; Pr ; Ps Grid side converter, rotor, stator active power
Qg ; Qr ; Qs Grid side converter, rotor, stator reactive power
he ,hs Stator voltage, stator flux vector position
Lf ,Rf Inductance and resistance of the grid filter
ema ; emb Stationary axis counter e.m.f components
Tr ¼ Lr =Rr Rotor time constant

Suffixes, Superscripts
d-q d-q axis
s-r Stator, rotor
List of Figures

Fig. 2.1 Squirrel-cage induction generator (SCIG) driven by a wind


turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 10
Fig. 2.2 Doubly Fed wound rotor induction generator driven by a wind
turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 11
Fig. 3.1 Configuration of a DFIG driven by a wind turbine connected
to a power grid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 22
Fig. 3.2 Wind turbine characteristics [8] . . . . . . . . . . . . . . . . . . . . . . . .. 24
Fig. 3.3 d-q equivalent circuit of the DFIG in synchronous
coordinates. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 24
Fig. 3.4 Block diagram of MRAS observer. . . . . . . . . . . . . . . . . . . . . .. 30
Fig. 3.5 Vector-control scheme based on SVO control for GSC
of the wind driven DFIG system . . . . . . . . . . . . . . . . . . . . . . .. 32
Fig. 3.6 Vector-control scheme based on SFO control for RSC
of the wind driven DFIG system . . . . . . . . . . . . . . . . . . . . . . .. 33
Fig. 3.7 Grid voltage and current of GSC during subsynchronous
speed operation of the DFIG system . . . . . . . . . . . . . . . . . . . .. 34
Fig. 3.8 Grid voltage and current of GSC during supersynchronous
speed operation of the DFIG system . . . . . . . . . . . . . . . . . . . .. 34
Fig. 3.9 Calculated and measured [8] values of q-axis GSC current . . .. 34
Fig. 3.10 Supply phase voltage and line current when q-axis of GSC
current reference stepped form −4 to 4 A . . . . . . . . . . . . . . . .. 35
Fig. 3.11 Calculated and measured [8] d-q axis rotor current
component, Dc-link voltage and rotor phase current when
q-axis rotor current reference component is stepped from
0 to 12 A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 36
Fig. 3.12 Calculated and measured [8] d-q axis rotor current and rotor
phase current when d-axis rotor current component reference
is stepped from 0 to 7 A . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 37
Fig. 3.13 Stator voltage and stator current when d-axis rotor current
reference component is stepped from 0 to 7 A . . . . . . . . . . . .. 37

xi
xii List of Figures

Fig. 3.14 Measured [8] and calculated d-q axis rotor current
and generator speed for step increase (top) and step decrease
(bottom) in wind speed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 38
Fig. 3.15 Steady-state optimum speed tracking . . . . . . . . . . . . . . . . . . . .. 38
Fig. 4.1 DFIG-based WECS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 42
Fig. 4.2 d–q equivalent circuit of the DFIG in synchronous rotating
reference frame taking iron loss into account. . . . . . . . . . . . . .. 43
Fig. 4.3 Block diagram of a wind driven DFIG with proposed
adaptive SMC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 50
Fig. 4.4 Measured [135] and calculated waveforms of the grid and
stator line voltages during supersynchronous speed operation
of the DFIG system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 51
Fig. 4.5 Measured [135] and calculated direct and quadrature
components of the grid and stator voltages during
supersynchronous speed of the DFIG system. . . . . . . . . . . . . .. 51
Fig. 4.6 Calculated waveforms of the grid and stator line voltages
during subsynchronous speed operation of the
DFIG system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 52
Fig. 4.7 Calculated direct and quadrature components of the grid
and stator voltages during subsynchronous speed operation
of the DFIG system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 52
Fig. 4.8 Measured [135] and calculated waveforms of the grid and
stator line voltages under disturbances using adaptive
SMC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 53
Fig. 4.9 Measured [135] and calculated direct and quadrature
components of the grid and stator voltages under disturbances
using adaptive SMC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 53
Fig. 4.10 Measured [135] and calculated waveforms of the grid and
stator line voltages under disturbances using conventional PI
control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 53
Fig. 4.11 Measured [135] and calculated direct and quadrature
components of the grid and stator voltages under disturbances
using conventional PI control . . . . . . . . . . . . . . . . . . . . . . . . .. 54
Fig. 4.12 Calculated grid, stator line voltages, and the corresponding
direct and quadrature voltages with rotor self-inductance
variation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 55
Fig. 4.13 Calculated grid, stator line voltages, and the corresponding
direct and quadrature voltages with rotor resistance
variation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 56
Fig. 5.1 Relationships between as bs ; ar br ; dq þ , and r dq reference
frames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 58
Fig. 5.2 Block diagram of proposed synchronization scheme of the
DFIG under unbalanced grid voltages . . . . . . . . . . . . . . . . . . .. 68
List of Figures xiii

Fig. 5.3 Calculated and measured [139] d-q positive sequence grid and
stator voltages during supersynchronous speed . . . . . . . . . . . .. 70
Fig. 5.4 Calculated and measured [139] d-q negative sequence grid
and stator voltages during supersynchronous speed . . . . . . . . .. 70
Fig. 5.5 Calculated and measured [139] waveforms of the grid and
stator line voltages during supersynchronous speed . . . . . . . . .. 71
Fig. 5.6 Calculated and measured [139] grid and stator voltage vectors
at steady state during supersynchronous speed . . . . . . . . . . . .. 72
Fig. 5.7 Calculated d-q positive sequence grid and stator voltages
during subsynchronous speed . . . . . . . . . . . . . . . . . . . . . . . . .. 72
Fig. 5.8 Calculated d-q negative sequence grid and stator voltages
during subsynchronous speed . . . . . . . . . . . . . . . . . . . . . . . . .. 72
Fig. 5.9 Calculated waveforms of the grid and stator line voltage
during supersynchronous speed . . . . . . . . . . . . . . . . . . . . . . . .. 73
Fig. 5.10 Calculated grid and stator voltage vectors at steady during
subsynchronous speed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 74
Fig. 5.11 Calculated and measured [139] d-q grid and stator
voltages with PI control scheme during unbalanced
grid voltages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 74
Fig. 5.12 Calculated and measured [139] waveforms of the grid and
stator line voltages with PI control scheme during unbalanced
grid voltage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 75
Fig. 5.13 Calculated and measured [139] values of grid and stator
voltage vector at steady state under unbalanced grid voltage
with PI control scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 75
Fig. 5.14 Calculated d-q positive sequence grid and stator voltages
during when the grid frequency is 48 Hz . . . . . . . . . . . . . . . .. 76
Fig. 5.15 Calculated d-q negative sequence grid and stator voltages
when the grid frequency is 48 Hz . . . . . . . . . . . . . . . . . . . . . .. 76
Fig. 5.16 Calculated waveforms of the grid and stator line voltages
when the grid frequency is 48 Hz . . . . . . . . . . . . . . . . . . . . . .. 76
Fig. 5.17 Calculated d-q positive sequence grid and stator voltages
during when the grid frequency is 52 Hz . . . . . . . . . . . . . . . .. 77
Fig. 5.18 Calculated d-q negative sequence grid and stator voltages
when the grid frequency is 52 Hz . . . . . . . . . . . . . . . . . . . . . .. 77
Fig. 5.19 Calculated waveforms of the grid and stator line voltages
when the grid frequency is 52 Hz . . . . . . . . . . . . . . . . . . . . . .. 77
Fig. 5.20 Calculated d-q positive and negative sequence grid and stator
voltages with rotor resistance variation . . . . . . . . . . . . . . . . . .. 78
xiv List of Figures

Fig. 5.21 Calculated waveforms of the grid and stator line voltages with
rotor resistance variation . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 79
Fig. 5.22 Calculated d-q positive and negative sequence grid and stator
voltages with rotor self-inductance variation . . . . . . . . . . . . . .. 79
Fig. 5.23 Calculated waveforms of grid and stator line voltages with
rotor self-inductance variation . . . . . . . . . . . . . . . . . . . . . . . . .. 80
Fig. A.1 Magnetizing curve of the induction machine
used in simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 84
List of Tables

Table A.1 Parameter and Data of DFIG wind driven system [8] . . . . . . . 84
Table C.1 Parameter of DFIG [135] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87

xv
About the Author

Adel Abdelbaset was born in Nag Hamadi,


Qena-Egypt, on October 24, 1971. He received his B.S.,
M.Sc., and Ph.D. from Faculty of Engineering,
Department of Electrical Engineering, Minia University,
Egypt, in 1995, 2000 and 2006, respectively. Now, Dr.
Adel is an full Professor in Power Electronics at the
Faculty of engineering, Minia University. He is a
Member of the Faculty of engineering, Minia University,
Egypt since 1996 until now. Also, Prof. Adel currently
works as Executive Manger of Advanced Lab. for
Electric Power Systems, Minia University, Faculty of
Eng., Electrical Eng. Dept., and Prof. Adel was a Head
of the Department of Science and Renewable Energy
Engineering, Faculty of Postgraduate Studies for
Advanced Science, Beni-Suef University. His research
interests are in the area of renewable energy sources,
power electronics, power system protection and control,
power quality and harmonics, neural network, fuzzy
systems.

xvii
Abstract

This book aims to develop a modified model reference adaptive system (MRAS)
observer for sensorless vector control of a wind driven Doubly Fed induction
generator (DFIG) connected to grid. This observer is proposed to estimate the
generator shaft speed using the stator voltage and current measurements. The
control system behaviors of grid side converter (GSC) and rotor side converter
(RSC) are described based on stator flux-oriented (SFO) and stator voltage-oriented
(SVO) controllers, respectively. The GSC ensures the regulation of the actual value
of Dc-link voltage to its desired value, while the RSC controls the active and
reactive powers injected by the DFIG to the grid independently. A mathematical
model of the DFIG as influenced by core loss and main flux saturation is presented
to improve the theoretical prediction. Digital simulations are carried out to
demonstrate the effectiveness of the proposed schemes at different operating con-
ditions using MATLAB/Simulink software package. Moreover, to validate the
correctness and accuracy of the proposed schemes, the calculated performances are
compared with those results measured experimentally in the literature.
Besides power decoupled control, grid synchronization control is another
important issue in the application of the WECS based on the DFIG, which enables
the DFIG to be connected to the power grid with minimum impacts to both the
WECS and the grid.
Grid synchronization enhancement of a wind driven DFIG using adaptive sliding
mode control (SMC) is described and evaluated in this book. The proposed scheme
directly controls the stator terminal voltage of the DFIG to track the grid voltage
without current control loop; hence, the structure of controller is simplified.
Finally, grid synchronization of a wind driven DFIG under unbalanced grid
voltage is described and evaluated in this book.

xix
Chapter 1
Introduction

The present supply of energy is mainly dependent on the fossil energy sources such
as petroleum, natural gas, coal. These fossil fuels took thousands of years to form,
and the base material for the fossil fuels was organic substances. Hence, fossil fuels
can be described as biomass stored for a long period. A huge amount of the fossil
fuels is already depleted in the twentieth century. But due to the increasing demand,
the extraction of fossil fuels will become more risky and expensive in the future. If
the consumption of fossil fuels continues as now, all available resources of natural
gas and petroleum will be exploited by twenty-first century. Coal reserves may be
available for a longer period. Thus within a few decades, the human generations
would have completely exploited the fossil energy resources which were formed
over a thousands of years. Fossil fuels will no longer be available as an energy
supply for future generations. Even with further discovery of major fossil fuel
resources, the fact that the fossil fuel sources are limited is not going to change. The
period of availability of fossil fuels can only be extended to a few years or a very
few decades at the best.
There are also limitations in the supply of Earth’s uranium reserves for operating
nuclear power stations. The estimated global reserve amounts to less than
20 million ton of which only 12.52 million tons are usable. Hence, nuclear power
cannot be a permanent alternative for fossil fuels due to the limitations in the
availability of uranium reserves. Also, the risk incurred by the use of nuclear power
suggests that renewable energy sources are a better alternative to foster to the
increasing energy demand. Renewable energy sources are the only way by which
the Earth’s energy demand can be met without affecting the climatic conditions. Of
the available renewable energy sources, wind energy has become one of the most
important and promising sources of renewable energy all over the world.
Wind energy is a free, renewable, clean, and non-polluting source of electricity.
Since earliest recorded history, wind power has been used to move ships, grind
grains, and pump water. Wind energy was used to propel boats along the Nile River
as early 5000 B.C. within several centuries before Christ; simple windmills were
used in china to pump water.
© Springer International Publishing AG 2018 1
A. Abdelbaset et al., Wind Driven Doubly Fed Induction Generator,
Power Systems, https://doi.org/10.1007/978-3-319-70108-0_1
2 1 Introduction

All electric-generating wind turbines, no matter what size, are comprised of a


few basic components: the part that actually rotates in the wind, the electrical
generator, a speed control system, and a tower. Some wind machines have fail-safe
shutdown system so that if part of the machine fails, the shutdown system turns the
blades out of the wind or puts brakes. Just like solar electric system, wind-powered
system can be used in two ways: off-grid or on-grid is when your home or business
is entirely disconnected from electric utility company, and we generate absolutely
all of the electricity we need. Usually these systems cost about 30% more than an
on-grid (or grid-tie system).
Wind power generation uses either fixed speed or variable speed turbines.
Variable speed wind turbine system based on DFIG has become the most popular
configuration in wind energy conversion system due to its merits of variable speed
constant frequency operation, decoupled active/reactive power control, maximum
power capture capability, reduced mechanical stress, low converter VA rating
(usually 30% of the generator capacity), and reduced power loss compared to other
solutions such as fixed-speed induction generators or fully rated converter systems.
All of those above-mentioned advantages of the DFIG are possible because of the
control scheme that can be implemented in the back to back converters of the DFIG.
Hence, the method of controlling back to back converter plays a significant role in
achieving better performance of the DFIG system. A speed sensor is usually needed
for vector control of the DFIG scheme. The use of position encoder has several
drawbacks in term of robustness, cost, cabling, and maintenance, so sensorless
operation is desirable.
Besides power decoupled control, grid synchronization control is another
important issue in the application of the wind energy conversion system (WECS)
based on the DFIG, which enables the DFIG to be connected to the power grid with
minimum impacts to both the WECS and the grid.
However, WECS is usually located in rural areas with weak grid connection, in
which grid voltage unbalance may arise even during normal operation. The
unbalanced grid voltage may be caused by unbalanced transmission line impe-
dance, three-phase unbalanced load, and single-phase high-power load. The grid
synchronization control of the DFIG under unbalanced grid voltage is also
important for the protections of both the WECS and the grid. If the grid voltage
unbalance is not taken into account in grid synchronization control of the DFIG, the
differences between stator voltages and grid voltages will become significant, which
will cause large current, torque, and power impact at the time of connecting.

1.1 Book Objectives

1. Development of a WECS model including a DFIG taking core loss and main
flux saturation into account as an electrical power generation unit. This is aimed
at improving the theoretical prediction and reducing the discrepancies between
the calculated and measured performance values in the literature.
1.1 Book Objectives 3

2. Modeling of sensorless vector control of a wind driven DFIG connected to grid


based on a modified model reference adaptive system (MRAS).
3. Design of a robust controller based on adaptive sliding mode control (SMC) for
grid synchronization of DFIG. Assessment of effectiveness of this controller by
comparing the calculated results, in the presence of core loss in the DFIG model,
with those measured in the literature.
4. Design of improved DFIG grid synchronization controller under unbalanced
grid voltage. Assessment of effectiveness of this controller by comparing some
calculated results with those measured in the literature.
This book has presented an analysis and discussion of sensorless control of the
wind driven DFIG connected to grid using a modified model reference adaptive
system MRAS. The dynamic model of the DFIG taking core loss and saturation into
account is presented. The vector control schemes use SFO control for the RSC to
provide independent control of active and reactive power. On the other hand, the
vector control approach based on SVO control is used for the GSC to keep the
Dc-link voltage constant. Digital simulations are carried out to evaluate the per-
formance of the proposed schemes. Comparison between the calculated perfor-
mances with those results measured experimentally in the literature is presented to
demonstrate the validity of the proposed schemes.
Also, in this book a synchronization method using adaptive SMC algorithm for
grid connection of a Doubly Fed induction generator (DFIG) in a wind generation
system is presented. The proposed scheme directly controls the stator terminal
voltage of the DFIG to track the grid voltage without current control
loop. A mathematical model of the DFIG as influenced by core loss is included to
improve the theoretical prediction. Comparisons between some simulated results
are compared with those measured experimentally in the literature.
Finally, this book presents a synchronization method for grid connection of a
DFIG in a wind generation system under unbalanced grid voltage. The improved
control scheme includes a positive sequence controller and a negative sequence
controller. The positive sequence controller controls positive sequence stator
voltages to follow positive sequence grid voltages, and the negative sequence
controller controls negative sequence stator voltages to follow negative sequence
grid voltages without current control loop. Comparisons between some simulated
results are compared with those measured experimentally in the literature.

1.2 Scope of the Book

The present book is organized in six chapters and three appendices. Apart from the
introduction and conclusion chapters, four chapters form the body of book.
Chapter 1: presents the introduction, the objectives, and the contents of the book.
Chapter 2: presents a brief review of types of wind generation systems and the
types of generator used in each system. The literature review describing
4 1 Introduction

DFIG-based wind turbine-generator systems will be presented. More specifically,


the related previous studies and researches on the modeling, the control strategies,
DFIG grid synchronization, and the state-of-the-art converter topologies applied in
DFIG-based wind turbine-generator systems will be presented.
Chapter 3: presents a modeling and control approach of the sensorless wind
driven DFIG connected to grid taking core loss and main flux saturation into
consideration. A modified MRAS observer is proposed to estimate the generator
shaft speed using the stator voltage and current measurements. The vector control
schemes use SFO control for the RSC to provide independent control of active and
reactive power and SVO control for the GSC to keep the Dc-link voltage constant.
Digital simulations are carried out to evaluate the performance of the proposed
schemes using MATLAB/Simulink software package. The performances of the
proposed schemes are investigated under different operating conditions. Simulated
results are compared with those published results measured experimentally in the
literature to demonstrate the accuracy and validity of the proposed schemes.
Chapter 4: presents a robust control scheme based on adaptive sliding mode for
grid synchronization of DFIG in wind generation system. The stator voltage is
directly controlled instead of passing through a chain of cascaded loops, and hence,
the structure of controller is simplified. In order to enhance the robustness of the
system against parametric uncertainty and external disturbances, the sliding mode
control with an integral action is introduced into the control loop. The machine
model tacking core loss is included to improve the theoretical prediction.
Simulation results are presented to validate effectiveness and robustness of the
proposed algorithm. Comparisons between some simulations results with those
measured experimentally in the literature are also presented.
Chapter 5: proposes a robust control scheme based on adaptive sliding mode for
grid synchronization of DFIG in wind generation system under unbalanced grid
voltage. The proposed scheme consists of positive sequence controller and negative
sequence controller. The positive sequence controller based on adaptive SMC
directly controls the positive sequence stator voltage to track the positive sequence
grid voltage. Whereas, the negative sequence controller based on integral control
directly controls the negative sequence stator voltage to track the negative sequence
grid voltage. Thus, no extra current control loops are required, thereby simplifying
the design of the controller. Digital simulations are carried out to demonstrate the
effectiveness and robustness of the proposed scheme. Moreover, some calculated
performances are compared with those results measured experimentally in the
literature.
Chapter 6: summarizes the main conclusions drawn from the research work
reported in this book.
In addition to these chapters, a quite useful list of references pertinent to the
topics treated in the book is given.
For related details, the book is ended with three appendices summarized as
follows:
Appendix I: Lists parameters of DFIG wind turbine system and laboratory
measurement of magnetizing inductance.
1.2 Scope of the Book 5

Appendix II: Presents proofing the stability of the modified MRAS using
Lampooner’s stability theorem.
Appendix III: Lists parameters of DFIG for the study of grid synchronization in
Chaps. 4 and 5.
Chapter 2
Literature Review

2.1 Introduction

Electrical power is the most widely used source of energy for our homes, work-
places, and industries. Population and industrial growth have led to significant
increases in power consumption over the past three decades. Natural resources like
coal, petroleum, and gas which drive our power plants, industries, and vehicles for
many decades are becoming depleted at a very fast rate. This serious issue has
motivated nations across the world to think about alternative forms of energy which
utilize inexhaustible natural resources. Wind plants have benefited from steady
advances in technology made over past 15 years. Much of the advancement has
been made in the components dealing with grid integration, the electrical machine,
power converters, and control capability. The days of the simple induction machine
with soft start are long gone. We are now able to control the real and reactive power
of the machine, limit power output, and control voltage and speed [1]. There is a lot
of research going on around the world in this area, and technology is being
developed that offers great deal of capability. It requires an understanding of power
systems, machines, and applications of power electronic converters and control
schemes put together on a common platform. Unlike a conventional power plant
that uses synchronous generators, a wind turbine can operate as fixed speed or
variable speed. In a fixed speed wind turbine, the stator of the generator is directly
connected to the grid. However, in a variable speed wind turbine, the machine is
controlled and connected to the power grid through a power electronic converter.
There are various reasons for using a variable speed wind turbine [2, 3]:
• More power is achieved from the variable speed generator in comparison with
its constant speed counterpart system.
• Reactive power compensator and soft starter system could be omitted because of
power electronic devices existence.
• Simple pitch control is available with a feasible cost.

© Springer International Publishing AG 2018 7


A. Abdelbaset et al., Wind Driven Doubly Fed Induction Generator,
Power Systems, https://doi.org/10.1007/978-3-319-70108-0_2
8 2 Literature Review

• Mechanical stresses are reduced by absorbing and reducing torque pulsations in


variable speed.
• Torque pulsation reduction improves power quality by eliminating power
variation and reducing flickers.
• Maximum power point tracking systems are employable, which improve vari-
able speed systems efficiency.
• Acoustic noise is also reduced by working at a lower speed in lower wind gusts.
The use of renewable energy sources for electric power generation is gaining
importance in order to reduce global warming and environmental pollution, this is
in addition to meeting the escalating power demand of the consumers. Generally,
wind power generation uses either fixed speed or variable speed turbines, the main
configurations of generators and converters used for grid connected variable speed
wind power system (WPS) are presented in the following sections.

2.2 Synchronous Generators Driven by a Fixed Speed


Turbine

A synchronous generator usually consists of a stator holding a set of three-phase


windings, which supplies the external load, and a rotor that provides a source of
magnetic field. The rotor may be supplied either from permanent magnetic or from
a direct current flowing in a wound field.

2.2.1 Wound Field Synchronous Generator Driven


by a Wind Turbine

The stator winding is connected to network through a four-quadrant power con-


verter comprised of two back-to-back sinusoidal PWM. The machine side converter
regulates the electromagnetic torque, while the grid side converter regulates the real
and reactive power delivered by the WPS to the utility. The wound field syn-
chronous generator has some advantages that are
The efficiency of this machine is usually high because it employs the whole
stator current for the electromagnetic torque production [4].
The main benefit of the employment of wound field synchronous generator with
salient pole is that it allows the direct control of the power factor of the machine,
consequently the stator current may be minimized at any operation circumstances.
The existence of a winding circuit in the rotor may be a drawback as compared
with permanent magnet synchronous generator. In addition, to regulate the active
and reactive power generated, the converter must be sized typically 1.2 times of the
WPS rated power [5].
2.2 Synchronous Generators Driven by a Fixed Speed Turbine 9

2.2.2 Permanent Magnet Synchronous Generator (PMSG)


Driven by a Wind Turbine

Many configuration schemes using a permanent magnet synchronous generator for


power generation had been adopted. In one of them, a permanent magnet syn-
chronous generator was connected to a three-phase rectifier followed by boost
converter. In this case, the boost converter controls the electromagnet torque. The
supply side converter regulates DC link voltage as well as control the input power
factor. One drawback of this configuration is the use of diode rectifier that increases
the current amplitude and distortion of the PMSG [6]. As a result, this configuration
has been considered for small size wind power system (WPS) (smaller than
50 kW).
In another scheme using PMSG, the PWM rectifier is placed between the gen-
erator and the DC link, while another PWM inverter is connected to the network.
The advantage of this system regarding the use of field-orientation control (FOC) is
that it allows the generator to operate near its optimal working point in order to
minimize the losses in the generator and power electronic circuit. However, the
performance is dependent on the good knowledge of the generator parameter that
varies with temperature and frequency. The main drawbacks, in the use of PMSG,
are the cost of permanent magnet that increases the price of machine, demagneti-
zation of the permanent magnet material, and it is not possible to control the power
factor of the machine [7].

2.3 Induction Generators Driven by a Variable Speed


Wind Turbine

The AC generator type that has most often been used in wind turbines is the
induction generator. There are two kinds of induction generator used in wind
turbines that are squirrel-cage and wound rotor.

2.3.1 Squirrel-Cage Induction Generator (SCIG) Driven


by a Wind Turbine

Three-phase squirrel-cage induction generators are usually implemented in stan-


dalone power systems that employ renewable energy resources, like hydropower
and wind energy. This is due to the advantages of these generators over conven-
tional synchronous generators. The main advantages are reduced unit cost, absence
of a separate DC source for excitation, ruggedness, brushless rotor construction, and
ease of maintenance. A three-phase induction machine can be operated as a
self-excited induction generator if its rotor is externally driven at a suitable speed,
10 2 Literature Review

and a three-phase capacitor bank of a sufficient value is connected across its stator
terminals. The stator winding in this generation system is connected to the grid
through a four-quadrant power converter comprised of two PWM VSI connects
back-to-back trough a DC link voltage, this can be shown in Fig. 2.1.
The control system of the stator side converter regulates the electromagnetic
torque and supplies the reactive power to maintain the machine magnetized. The
supply side converter regulates the real and reactive power delivered from the
system to the utility and regulates the DC link, but the uses of SCIG have some
drawbacks as following [8]:
Complex system control, whose performance is dependent on the good
knowledge of the generator parameter that varies with magnetic saturation, tem-
perature, and frequency.
The stator side converter must be oversized 30–50% with respect to rated power,
in order to supply the magnetizing requirement of the machine.

2.3.2 Doubly Fed Induction Generator (DFIG) Driven


by a Wind Turbine

The wind power system shown in Fig. 2.2 consists of a DFIG, where the stator
winding is directly connected to the network and the rotor winding is connected to
the network through a four-quadrant power converter comprised of two
back-to-back sinusoidal PWM. The thyristor converter can be used, but they have
limited performance.
Usually, the controller of the rotor side converter regulates the electromagnetic
torque and supplies part of the reactive power to maintain the magnetization of the

GRID

Dc-link
Wind Speed

Gear
Box SCIG

Rotor side Grid side


converter converter

Fig. 2.1 Squirrel-cage induction generator (SCIG) driven by a wind turbine


2.3 Induction Generators Driven by a Variable Speed Wind Turbine 11

Wind Speed

Gear
DFIG
Box
GRID

Dc-link
Filter

Rotor side Grid side


converter converter

Fig. 2.2 Doubly Fed wound rotor induction generator driven by a wind turbine

machine. On the other hand, the controller of the grid side converter regulates the
DC link voltage [9]. Compared to synchronous generator, this DFIG offers the
following advantages [10]:
• Reduced inverter cost, because inverter rating typically 25% of the total system
power. This is because the converters only need to control the slip power of the
rotor.
• Reduced cost of the inverter filter, because filters rated for 0.25 p.u. total system
power, and inverter harmonics represent a smaller fraction of total system
harmonics.
• Robustness and stable response of this machine facing against external
disturbances.

2.3.2.1 Operation Modes

DFIGs have two dedicated operating modes [11] as follows:


1. Operation mode, in which generator rotor rotates at a speed above the syn-
chronous speed and is called supersynchronous mode. In this mode, slip is
negative, and both stator and rotor windings deliver power to the grid.
2. Operation mode, in which generator operates under synchronous speed and is
called as the subsynchronous mode. In this mode, slip is positive, and stator
winding delivers power to both the grid and the rotor winding. Total obtained
power from stator winding does not exceed the producible power in super-
synchronous mode with respect to smaller rotational speed in this mode.
12 2 Literature Review

2.4 Modeling of a Wind Turbine-Generator System

The modeling of a wind turbine-generator system consists of the aerodynamic


modeling, the drive train system modeling, the DFIG modeling, and the power
converter modeling. Hence, this part of the study will only focus on the modeling of
such system.

2.4.1 Aerodynamic Modeling

In [12], Tao sun deduced the maximum energy that a wind turbine system can
extract from the air system under ideal conditions. In [13], the authors derived the
relationship between the mechanical power input and the wind speed passing
through a turbine rotor plane, which can be expressed by the power coefficient of
the turbine. There are three most commonly used methods to simulate the power
coefficient which is provided by the wind turbine manufacturer. The first two
methods are given in Refs. [12, 14, 15]. The third method is the lookup table
method and given in Refs. [16, 17]. There are two other methods to approximate the
power efficiency curve, but they are not commonly used. Interested readers can find
them in [18, 19].

2.4.2 Drive Train Modeling

For the drive train system modeling, the work in Ref. [20] elaborately explained the
reduced mass conversion method and compared a six-mass model with reduced
mass models for transient stability analysis. In [21], Stavros A. Papathanassiou used
a six-mass drive train model to analyze the transient processes during faults and
other disturbances. In [22], three different drive train models and different power
electronic converter topologies were considered to study the harmonic assessment.
Reference [23] compared the transient stabilities of a three-mass model, a two-mass
model, and a one-mass model. In addition, the effects of different bending flexi-
bilities, blade and hub inertias on the transient stabilities of large wind turbines were
also analyzed. In [24], a three-mass model, which took into account the shaft
flexibility and blade flexibility in the structural dynamics, was developed and then
used to derive a two-mass model. In [20, 25], the authors concluded that a two-mass
drive train model was sufficient for transient stability analysis of wind
turbine-generator systems. Besides, the two-mass model is widely used in Refs.
[26–31]. Other references, such as [15, 32–34] focused their study on the generator
control and modeling, where the drive train system was simply expressed by single
mass models.
2.4 Modeling of a Wind Turbine-Generator System 13

2.4.3 DFIG Modeling

The Doubly Fed induction machines can be categorized into four types. These types
are the standard Doubly-Fed induction machine, the cascaded Doubly-Fed induc-
tion machine, the single-frame cascaded Doubly-Fed induction machine and the
brushless Doubly-Fed induction machine [35]. However, only the standard type and
brushless type of Doubly-Fed induction machines have been applied in wind
turbine-generator systems. In Ref. [36], the authors developed the brushless
Doubly-Fed induction generator by employing two cascaded induction machines to
eliminate the brushes and copper rings and used a closed-loop stator flux-oriented
control scheme to achieve active and reactive power control. In [37], Yongchang
Zhang proposed a direct power control (DPC) strategy for cascaded brushless
Doubly Fed induction generators which featured quick dynamic responses and
excellent steady-state performances. The DFIG model can be expressed in the
stationary stator reference frame, the reference frame rotating at rotor speed, and the
synchronously rotating reference frame. In [33, 38], the authors adopted the syn-
chronously rotating reference frame in order to simplify the controller design
because of the fact that all the currents and voltages expressed under this reference
frame will be of a DC nature. While, in [9], both stator and rotor variables were
referred to their corresponding natural reference frames, and the machine model
expressed in such reference frame is called the “Quadrature-Phase Slip-Ring”
model.
The DFIG model can usually be expressed by reduced order models, which can
yield a third-order model by neglecting the derivative terms of the stator flux and
first-order model by neglecting both the derivative terms of the stator flux and rotor
flux [39]. But in [38], the authors proposed an enhanced third-order model which
considered the DC-components of the stator currents and gave a comparison
between a full-order model and the proposed model for wind ramp conditions.
Alvaro Luna, in [40], deduced a new reduced third-order model by ignoring the
stator resistances and inductances through applying the Laplace transformation and
compared the proposed model with a full-order model for transient analysis.
There are many references which made the comparison between the full-order
model and reduced order models [41–43]. In [44], the authors even considered the
saturated conditions and made a detailed comparison among these unsaturated and
saturated full-order models and reduced order models. Pablo Ledesma, in [45],
compared a third-order model with a full-order model in two extreme operation
points under short-circuit fault conditions. These points are subsynchronous speed
and supersynchronous speed, respectively. As known, the difference between the
model of a squirrel-cage induction generator and a Doubly-Fed induction generator
is the rotor input. Hence, the simplified models of squirrel-cage induction genera-
tors may be helpful for understanding the reduced order models of DFIGs.
Interested readers can find them in [46, 47].
14 2 Literature Review

2.4.4 Power Converter Modeling

The traditional power converter used in wind turbine-generator systems is a


back-to-back two-level PWM converter. The three-phase voltage source PWM
converter model can be expressed in the abc reference frame and the d-q syn-
chronous reference frame which is deduced for control purposes. The mathematical
model based on space vectors expressed in the abc reference frame was derived in
[48]. In [49–51], the authors showed the detailed work about the transformation of a
PWM converter model from the abc reference frame to the d-q synchronous ref-
erence frame. For wind turbine applications, some researchers simplified the power
converter model by employing an equivalent AC voltage source that generates the
fundamental frequency [34]. In [52], José R. Rodríguez gave the detailed
description for the working principles, control strategies and made comparisons for
three-phase voltage source and current source PWM converters.

2.5 Control Strategies for a Wind Turbine-Generator


System

The control schemes for a wind turbine-generator system include the pitch angle
control, maximum power point tracking control, and the DFIG control. The tradi-
tional control techniques and advanced control techniques for wind
turbine-generator systems are reviewed in this section.

2.5.1 Pitch Angle Control

The pitch angle control is a mechanical method of controlling the blade angle of the
wind turbine when the captured wind power exceeds its rated value or wind speed
exceeds its rated value. In this way, pitch angle control is enabled to limit the
maximum output power to be equal to the rated power, and thus protect the gen-
erator when the wind speed experiences gusts. The pitch angle controller is only
activated at high wind speeds.
There are numerous pitch angle regulation techniques described in the literature
[5, 53–58]. The conventional pitch angle control usually uses PI controllers [5, 53,
54]. However, several advanced pitch control strategies were proposed. A new
approach for the pitch angle control, which worked well for unstable and noisy
circumstance, was presented in [55]. Besides, a fuzzy logic pitch angle controller
was developed in [56], which did not need much knowledge about the system.
Furthermore, a pitch angle controller using a generalized predictive control was
presented in [57], whose strategy was based on the average wind speed and the
standard deviation of the wind speed. Another pitch control scheme was proposed
2.5 Control Strategies for a Wind Turbine-Generator System 15

in [58], in which a self-tuning regulator adaptive controller that incorporated a


hybrid controller of a linear quadratic Gaussian neuro controller and a linear
parameter estimator was developed for the pitch angle control. In [59], the authors
only applied a fuzzy logic pitch angle controller in a wind turbine-generator system
to achieve the maximum power point tracking control and power control.

2.5.2 Maximum Power Point Tracking Control

In order to achieve the maximum power point tracking (MPPT) control, some
control schemes have been presented. The maximum power point tracking control
can be mainly divided into two types. They are the conventional control schemes
and intelligent control schemes.

2.5.2.1 Conventional Control Schemes

The conventional control schemes can also be divided into current mode control
and speed mode control, which depends on the setting of reference values. The
reference values are the active power and electromagnetic torque for current mode
control [60–62] and the rotational speed for the speed mode control [63]. In [64],
the author compared these two control strategies for dynamic transient analysis and
concluded that the current mode control has slow response with simple construc-
tion, while the speed mode control has fast response with complex construction.
The discussions and limitations of these two control schemes were presented in
[65].
In fact, the wind speeds in above conventional control schemes need to be
exactly measured. However, the anemometer cannot precisely measure the wind
speed because of the flow distortion, complex terrain, and tower shadow influence
[66]. Hence, some studies on maximum wind energy tracking without wind
velocity measurement had been developed in [67, 68].

2.5.2.2 Intelligent Control

The intelligent control strategies usually apply the hill-climbing control and the
fuzzy logic control to the maximum power point tracking control. The traditional
hill-climbing control uses a fixed-step speed disturbance optimal control method to
determine the speed, perturbation size, and direction according to the changes in the
power before and after sampling [69]. However, this control method is usually slow
in speed because the step disturbance is fixed. Therefore, some improved
hill-climbing control methods were proposed. For example, a method of using
variable-step wind energy perturbation method to control the captured wind power
was analyzed in [65]. Another advanced hill-climbing searching method with an
16 2 Literature Review

online training process, which can search for the maximum wind turbine power at
variable wind speeds, even without the need for knowledge of wind turbine char-
acteristics, wind speed, and turbine rotor speed, was developed in [70]. Fuzzy logic
control-based MPPT strategies have the advantages of having robust speed control
against wind gusts and turbine oscillatory torque, having superior dynamic, and
steady performances, and being independent of the turbine parameters and air
density; see [66, 71].

2.5.2.3 Other Control Strategies

In [72], the authors presented a novel adaptive MPPT control scheme in which the
wind speed was estimated by the output power and the efficiency of the generator,
and the maximum efficiency was estimated by the maximum tip-speed ratio tracker.
A novel MPPT strategy that was based on directly adjusting the DC/DC converter
duty cycle according to the results of comparisons between successively monitored
wind turbine output powers was proposed in [73], in which there was no require-
ment for the knowledge of wind turbine characteristic and measurements of the
wind speed.

2.5.3 DFIG Control

Control of the DFIG is more complicated than the control of a squirrel-cage


induction generator, because the DFIG can operate at subsynchronous speed and
supersynchronous speed by regulating the rotor terminal voltages. Through the
years, many researchers have presented various types of DFIG control strategies,
such as field-oriented control, direct torque/power control, predictive control,
sensorless control, and nonlinear control.

2.5.3.1 Field-Oriented Control

Field-oriented control (FOC) or vector control is commonly used in DFIG controls


due to its ability of controlling the motor speed more efficiently, and the low
economic cost to build an FOC system. Field-oriented control also provides the
ability of separately controlling the active and reactive power of the generator.
Currently, there are mainly two types of field-oriented control in DFIG, which are
stator voltage-oriented control and stator flux-oriented control, respectively. The
stator flux-oriented control is widely used in the DFIG control designs [26, 9], in
which the q-axis current component is used for active power control and the d-axis
component is used for reactive power control.
While for the stator voltage-oriented control, the situation is on the contrary
[74, 75], the d-axis component is used for active power control and the q-axis
2.5 Control Strategies for a Wind Turbine-Generator System 17

current component is used for reactive power control. In [76], the author compared
real and reactive power control for a DFIG-based wind turbine system using stator
voltage- and stator flux-oriented control, respectively, and the simulation results
illustrated same performances.

2.5.3.2 Direct Torque/Power Control

Recently, a new technique for directly control of the induction motors’ torque or
power was developed, which included direct torque control (DTC) and direct power
control (DPC). Direct torque control scheme was first developed and presented by I.
Takahashi and T. Nogouchi [77, 78]. Based on the principles of DTC for electrical
machines, direct power control for a three-phase PWM converter was introduced in
[79].
Direct torque control techniques do not require current regulators, coordinate
transformations, specific modulations, and current control loops [80]. Thus, direct
torque control has the ability of directly controlling the rotor flux linkage magnitude
and generator torque through properly selecting the inverter switching states [10].
To show the advantages of DTC, the comparison between the field-oriented control
and direct torque control was made in [81]. Direct torque control using space vector
modulation technology was presented in [82]. In [83, 84], the authors applied basic
direct torque control to a Doubly-Fed induction generator. Direct torque control
which was achieved without PI controller and only required the knowledge of grid
voltages, rotor currents, and rotor position as was proposed in [48, 80]. Z. Liu, in
[85, 86], proposed a novel direct torque control scheme which was developed based
on the control of the rotor power factor.
Direct power control has the merits of being simple, requiring fewer sensors,
having low computational complexity, fast transient response, and low machine
model dependency compared with direct torque control [87]. In [88], the com-
parison between field-oriented control and direct power control for a PWM rectifier
was presented, and the simulation results showed that the virtual-flux-based direct
power control was superior to the [86] voltage-based direct power control and
field-oriented control. In [89, 90], the authors used direct power control in a
DFIG-based wind turbine system under unbalanced grid voltage conditions. A new
direct power control, which was based on the stator flux and only needed the stator
resistance values of the machine parameters, was proposed in [91].

2.5.3.3 Other Control Strategies

In recent years, increasing attention is being paid to the application of predictive


control in the field of the DFIG-based wind turbine-generator systems [92–94].
Several predictive direct power control strategies were studied and compared for
AC/DC converters in [95].
18 2 Literature Review

Sensorless operation is important for wind applications due to the need for low
cost and high reliability particularly for wind turbines which are usually installed in
harsh environment [96]. There are many studies worked on the sensorless control;
see Refs. [97–102]. Sensorless control is usually achieved by estimating the rotor
position, so that there is no need for the rotor position encoder. A common way
used for the estimation of parameters without taking any feedback is the use of
model reference adaptive system (MRAS) observer as used in [100–102].
Moreover, direct torque/power control strategies can be considered as “sensor-
less type” control techniques because direct torque/power control could obtain a
good dynamic control of the torque/power without any mechanical transducers on
the machine shaft [81]. A nonlinear control approach, which used the nonlinear
static and dynamic state feedback controllers with a wind speed estimator in a wind
turbine-generator system, was proposed in [25].

2.6 Power Converter Topologies for a Wind


Turbine-Generator System

Power electronics, being the technology of efficiently converting electric power,


plays an important role in wind power systems. In recent years, the multi-level
converters and matrix converters became main solutions for medium voltage drives.
In this section, the application of multi-level converters and matrix converters in
wind turbine-generator systems is reviewed.

2.6.1 Multi-level Converters

Compared with traditional two-level converters, multi-level converters have many


advantages, such as more sinusoidal output voltage waveforms, lower total har-
monic distortion (THD), reduced filter size and cost, reduced switching losses in the
IGBTs, lower dv/dt [103, 104]. This is due to the fact that the output voltages can
be formed using more than two voltage levels.
Generally speaking, multi-level converters can be classified into three categories
[105]:
Neutral-point-clamped (NPC) converters.
Flying capacitor converters.
Cascaded H-bridge (CHB) converters.
Multi-level neutral-point-clamped converters are most widely used in wind
turbine-generator systems. In [106], three-level NPC converters were applied in
PMSG-based wind turbine systems with field-oriented control. In [107], the author
used a three-level neutral-point-clamped PWM converter to drive a permanent
magnet synchronous generator, in which a space vector modulated direct power
2.6 Power Converter Topologies for a Wind Turbine-Generator System 19

control, was applied. In [108], a new application of the predictive direct power
control was presented for a Doubly-Fed induction machine equipping with
three-level NPC converters, in which constant switching frequency technology was
achieved. In [109], the active and passive components of a NPC converter, such as
insulated-gate bipolar-transistors, free-wheeling diodes, clamping diodes, grid fil-
ters, DC-bus capacitors, were designed for a wind turbine system equipped with a
squirrel-cage induction generator. A comparison between traditional two-level
converters and three-level NPC converters for a wind power system was made
[110]. In [111, 112], the authors made comparisons between the
neutral-point-clamped converters, flying capacitor converters, and cascaded
H-bridge converters for wind power generation. The application of cascaded
H-bridge converters in wind turbine-generator systems was developed in recent
years; interested readers can find them in [113, 114].

2.6.2 Matrix Converters

The matrix converter concept, which was first introduced by A. Alesina and M.
G. B. Venturini [86], has become increasingly attractive for wind power applica-
tions. When compared with back-to-back two-level converters, matrix converters
have some significant advantages, such as sinusoidal input and output currents,
absence of a Dc-link capacitor, fewer IGBT switches, simple and compact power
circuit, operation with unity power factor for any load, and regeneration capability
[115, 116].
Numerous works have been published for the application of matrix converters in
wind turbine-generator systems. The application of a matrix converter for power
control of a DFIG-based wind turbine system can be found in [117]. In [118], a
wind turbine system, which was composed by a SCIG and a matrix converter, was
presented. For the applications of PMSG-based wind turbine systems, one can
easily find them in [119].

2.7 DFIG Grid Synchronization

DFIG should be synchronized with grid before connection in order to have mini-
mum impact on power system. But only a few authors studied the DFIG grid
connection control [120].
Due to the universality of vector control (VC), it has also been extended to the
grid synchronization process. In general, a cascaded structure using four PIs (two
for outer stator voltage loop and two for inner rotor current loop) is used to achieve
the equality of amplitude, frequency, and phase [121, 122], which requires the
information of stator voltage, grid voltage, rotor current, and rotor position. To
reduce the complexity and tuning work, a single loop stator voltage oriented direct
20 2 Literature Review

voltage control strategy is proposed in [123], which reduces the control loops by
half by using two PIs only and eliminates the use of rotor current. As a result, the
demand on the computation power and the number of parameters for tuning is
reduced. Although smooth grid synchronization is achieved by using VC, the
tuning effort of PI is still necessary.
In [124], DTC method is used for grid synchronization and normal condition but
using PI controller beside hysteresis one, variable switching frequency and
noticeable torque ripples are its disadvantages.
The main goal of successful synchronization is to reduce stresses on the elec-
trical and mechanical components of the wind turbine. Also, it helps in preventing
power system disturbance due to stator–grid connection. The mechanical stress is
caused by heavy transient torque at the start-up, and the electrical stress is due to
huge heavy start-up currents. The mechanical stress can damage the gearbox, shaft,
and the rotor of the machine while electrical stress can damage the insulation, and
winding of the stator and the rotor over a period of time [124].
In [125], the rotor current is controlled for grid synchronization. Having
noticeable differences between stator and grid voltages because of voltage feedback
lack is the main drawback of this method. Reference [80] presents Direct Virtual
Torque Control (DVTC) that is achieved without PI controller and requires only the
measurement grid voltage, rotor current, and rotor position. But, because of using
hysteresis controller, switching frequency is variable and ripples of the flux and
torque are high.
Chapter 3
A Modified MRAS Observer
for Sensorless Control of a Wind
Driven DFIG Connected to Grid

3.1 Introduction

The Doubly Fed induction generator (DFIG) is a popular wind turbine system due
to its high energy efficiency, reduced mechanical stress on the wind turbine, and
relatively low power rating of the connected power electronics converter of low
costs [126].
For the DFIG, sensorless operation is desirable because the use of position
encoder have several drawbacks in term of robustness, cost, cabling, and
maintenance.
In this chapter, a modified MRAS for the speed estimation of DFIG connected to
grid in order to implement the sensorless vector control is described and evaluated.
A mathematical model of DFIG as influenced by core loss and main flux saturation
is also presented. The vector control schemes of the rotor side converter (RSC) and
the grid side converter (GSC) are developed in the stator flux-oriented (SFO) frame
and stator voltage-oriented (SVO) frame, respectively. The GSC ensures the reg-
ulation of the DC voltage to the desired value, while the RSC controls the active
and reactive powers injected by the DFIG to the grid independently. Modeling and
control of the system are studied and analyzed at different operating cases. Digital
simulations, using MATLAB/Simulink software package, are carried out to
demonstrate the effectiveness of the proposed schemes. Moreover, to validate the
correctness and accuracy of the proposed scheme, the calculated values are com-
pared with those measured in the literature [8].

© Springer International Publishing AG 2018 21


A. Abdelbaset et al., Wind Driven Doubly Fed Induction Generator,
Power Systems, https://doi.org/10.1007/978-3-319-70108-0_3
22 3 A Modified MRAS Observer for Sensorless Control …

3.2 Description of the System Under Study

The basic configuration of a DFIG driven by a wind turbine is shown in Fig. 3.1 the
machine may be simulated as the wound-rotor induction machine having 3-phase
supply on the stator and 3-phase supply on the rotor. The rotor circuit is connected
through slip rings to the back-to-back converters arrangement controlled by pulse
width modulation (PWM) strategies. The back-to-back converters consist of two
voltage source converters (ac-dc-ac) having a dc link capacitor connecting them.
The RSC takes the variable frequency voltage and converts it into a dc voltage.
The GSC has the voltage conversion from the dc link as input and ac voltage at grid
as output. The vector control of the RSC and the GSC is developed in stator
flux-oriented reference frame and stator voltage-oriented reference frame. RSC acts
as a voltage source converter, while the GSC is expected to keep the capacitor
voltage constant under wind speed variations [127]. The active power, reactive
power, and voltage controllers of Fig. 3.1 are included for obtaining the rotor side
 
and grid side voltage references (Vabcr and Vabcg ).

3.3 Dynamic Modeling of a Wind Turbine DFIG System

3.3.1 Aerodynamic Wind Turbine Model

In practice, the wind turbine can be characterized by its Cp  k curve where the tip
speed ratio k is defined as [128].
Wind Speed

DFIG
GRID

RSC Dc-link GSC


Filter
Q r Pr i2 i1
V dc
* Qg
V abcr *
Vabcg
RSC based GSC based Pg
On SFO on SVO

Ps* Q s* *
Q gsc V dc*

Fig. 3.1 Configuration of a DFIG driven by a wind turbine connected to a power grid
3.3 Dynamic Modeling of a Wind Turbine DFIG System 23

xT R
k¼ ð3:1Þ
Vw

where Vw is the wind speed (m/s), xT is the turbine angular speed, and R is the rotor
radius (m).
The power coefficient Cp ðk; bÞ can be expressed as a function of blade pitch
angle b and the tip speed ratio k as [128]:
 
116
Cp ðk; bÞ ¼ 0:5176  0:4b  5 e21=ki þ 0:0068k ð3:2Þ
ki

where

1 1 0:035
¼ 
ki k þ 0:08b b3 þ 1

The mechanical turbine power PT extracted from the wind can be expressed as:

1
PT ¼ qpR2 Cp ðb; kÞVw3 ð3:3Þ
2

where q is the air density (kg/m3).


The maximum power coefficient Cp max corresponds to the optimal tip speed ratio
kopt with constant blade pitch angle. Clearly, the wind turbine speed should be
changed with wind speed so that the optimum tip speed ratio is maintained. The
mechanical turbine maximum power Popt and the corresponding turbine rotor speed
of the wind turbine xT opt are related as:

Popt ¼ Kopt x3T opt ð3:4Þ

where

R5
Kopt ¼ 0:5qpCp max
k3opt

Figure 3.2 shows the characteristic, with a fixed b, for the 7.5 kW [8]. The curve
Popt defines the maximum energy capture, and the objective of a tracking control is
to keep the turbine on this curve as the wind velocity varies.
The single mass mechanical model of the turbine may be considered as:

JT pxr ¼ TT  Te  Bxr ð3:5Þ

where JT , TT , and B are the total inertia, aerodynamic torque brought at the high
speed side and the damping coefficient, respectively.
24 3 A Modified MRAS Observer for Sensorless Control …

Fig. 3.2 Wind turbine


characteristics [8]

Power (kW)

Turbine speed referred to generator side (rpm)

3.3.2 Dynamic Model of a DFIG Taking Iron Losses


and Magnetic Saturation into Consideration

The dynamic equivalent circuit of the DFIG in d-q synchronous rotating reference
frame, taking the iron loss and main flux saturation into account, is shown in
Fig. 3.3 [129].
From the equivalent circuit of Fig. 3.3, the following set of differential equations
of the DFIG can be expressed as [129]:

Fig. 3.3 d-q equivalent circuit of the DFIG in synchronous coordinates


3.3 Dynamic Modeling of a Wind Turbine DFIG System 25

dids dkdm  
Vds ¼ Rs ids þ Lls þ  xe Lls iqs þ kqm ð3:6Þ
dt dt
diqs dkqm
Vqs ¼ Rs iqs þ Lls þ þ xe ðLls ids þ kdm Þ ð3:7Þ
dt dt
didr dkdm  
Vdr ¼ Rr idr þ Llr þ  ðxe  xr Þ Llr iqr þ kqm ð3:8Þ
dt dt
diqr dkqm
Vqr ¼ Rr iqr þ Llr þ þ ðxe  xr ÞðLlr idr þ kdm Þ ð3:9Þ
dt dt
dkdm
RFe idFe ¼  xe kqm ð3:10Þ
dt
dkqm
RFe iqFe ¼ þ xe kdm ð3:11Þ
dt
idFe þ idm ¼ ids þ idr ð3:12Þ

iqFe þ iqm ¼ iqs þ iqr ð3:13Þ

The electromagnetic back torque can be expressed as:


np    
Te ¼  ðLlr idr þ kdm Þkqm  Llr iqr þ kqm kdm ð3:14Þ
Llr

Applications of the above model require two nonlinear functions: The first one
represents the change of equivalent core-loss resistance Rfe with the stator supply
frequency f as [130]:

Ke f fn þ Kh f
Rfe ¼ ðRfe Þfn ð3:15Þ
Ke f fn þ Kh fn

where Ke and Kh are the hysteresis and eddy current coefficients, respectively. The
open circuit test is carried out to obtain the values of hysteresis and eddy current
coefficients. In this test, the three-phase induction machine is supplied from syn-
chronous generator to obtain variable voltage and variable frequency for main-
taining the flux density constant. The corresponding values of the power input Piron
were recorded. Plot of Piron =f versus f is a straight line whose slope and intersection
with the vertical axis determine the hysteresis and eddy current coefficients Ke and
Kh , respectively.
The second function represents the change of magnetizing inductance with main
flux as:
26 3 A Modified MRAS Observer for Sensorless Control …

L m ¼ F ð km Þ ð3:16Þ

where Lm ¼ kimm
Magnetizing inductance of the DFIG may vary significantly when the main
magnetic flux is saturated, this requires online identification algorithm of the
magnetizing inductance [131]:
The magnitude of magnetic flux vector is calculated from its components as:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
km ¼ k2am þ k2bm ð3:17Þ

The air-gap magnetizing flux components can be obtained in the stationary


reference frame as:
Z
kam ¼ ðVas  Rs ias Þdt  Lls ias ð3:18Þ

Z
 
kbm ¼ Vbs  Rs ibs dt  Lls ibs ð3:19Þ

where Vas , Vbs , ias , and ibs are the stator voltage and stator current components in
stationary reference frame (ab axis). The magnetizing curve of the machine is
identified offline in the laboratory from no-load test and is represented with a
suitable polynomial relating the magnetizing flux with the magnetizing current.
Since the magnetizing flux is known, it is possible to estimate the magnetizing
inductance using the known nonlinear inverse magnetizing curve. Data of the
magnetization curve which obtained from no-load test is listed in Appendix I [131].

3.3.3 DC Link Model

The DC link capacitor provides dc voltage to the RSC and any attempt to store
active power in the capacitor would raise its voltage level. To ensure stability of the
system, power flow of the grid side Pg and rotor side converters Pr , as indicated in
Fig. 3.1, should guarantee the following control objective:

Pg ¼ Pr ð3:20Þ

The differential equation of the dc link can be written as:

dVdc
c ¼ i1  i2 ð3:21Þ
dt

where Vdc is the DC voltage at the converter output terminals and c is the smoothing
capacitor.
3.3 Dynamic Modeling of a Wind Turbine DFIG System 27

Assuming no power losses for the converters, i1 and i2 can be derived as:

Pg
i1 ¼ ð3:22Þ
Vdc
Pr
i2 ¼ ð3:23Þ
Vdc

3.4 Vector Control of a DFIG for Grid-Connected


Operations

The DFIG wind turbine control system generally consists of two parts.
1. Control of the RSC.
2. Control of the GSC.
The objective of the RSC is to allow the DFIG wind turbine for decoupled
control of active power (mechanical input power) and reactive power (rotor exci-
tation current).
In the SFO reference frame, the d-axis is aligned with the stator flux linkage
vector ks , namely, kds ¼ ks and kqs ¼ 0, using equations describing DFIG model in
d-q rotating reference frame model disregarding core loss and main flux saturation,
the following relationships can be obtained [26].
 
Lm
iqs ¼  iqr ð3:24Þ
Ls
 
kds  Lm idr
ids ¼ ð3:25Þ
Ls
D
Vdr ¼ Vdr  ðxe  xr ÞrLr iqr ð3:26Þ

Vqr ¼ VqrD þ ðxe  xr ÞðLo ims þ rLr idr Þ ð3:27Þ


28 3 A Modified MRAS Observer for Sensorless Control …

where

D didr
Vdr ¼ Rr idr þ rLr ;
dt
diqr
VqrD ¼ Rr iqr þ rLr ;
dt
L2
r¼1 m
Ls Lr
L2m ks
Lo ¼ and ims ¼
Ls Lm

The stator active and reactive powers are obtained as follows:


 
3 Lm
Ps ¼  Vqs iqr ð3:28Þ
2 Ls
 
3 kds  Lm idr
Qs ¼ Vqs ð3:29Þ
2 Ls

Equations (3.28) and (3.29) clearly show that the stator real and reactive power
can be independently controlled by regulating the rotor q- and d-axis currents,
respectively.
The stator flux components in stationary reference frame (a  b) can be obtained
through the integration of the difference between the phase voltage and the voltage
drop in the stator resistance:
R
ksa ¼ R ðVsa  Rs isa Þdt
 ð3:30Þ
ksb ¼ Vsb  Rs isb dt

The stator flux vector position (hs ) can be calculated using stator flux compo-
nents as:
 
ksb
hs ¼ tan1 ð3:31Þ
ksa

The stator flux vector position (hs ) is used to calculate the slip angle
(hslip ¼ hs  hr ) which is used to estimate the unit vector (cos hslip and sin hslip ). The
estimated unit vector is used as input to the rotor coordinate transformation.
Also, the stator flux magnitude ks is to be calculated from its components as:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ks ¼ k2as þ k2bs ð3:32Þ

The objective of the GSC is to keep the Dc-link voltage constant regardless of
the magnitude and direction of the rotor power. A vector control approach as used,
with a reference frame oriented along the stator (or the grid) voltage vector position,
3.4 Vector Control of a DFIG for Grid Connected Operations 29

enabling independent control of the active and reactive power flowing between the
grid and the GSC.
In the d-q reference frame, the voltage balance across the grid filter shown in
Fig. 3.1 is [132]:

digd
Vds ¼ Rf igd þ Lf  xe Lf igq þ Vgd ð3:33Þ
dt
digq
Vqs ¼ Rf igq þ Lf þ xe Lf igd þ Vgq ð3:34Þ
dt

where igd , igq , Vgd and Vgq are d and q axis components of the GSC current and
output voltage, respectively. Lf and Rf are the inductance and resistance of the grid
filter.
In the SVO reference frame, the d-axis is aligned with the stator voltage vector
Vs , so that, Vqs ¼ 0, and Vds ¼ Vs ¼ constant.
The active power and reactive power of GSC can be expressed as:

3
Pg ¼ Vds igd ð3:35Þ
2
3
Qg ¼  Vds igq ð3:36Þ
2

So that active and reactive powers are proportional to d- and q-axis components
of converter output currents, respectively.
The d- and q-axis components of the GSC current are regulated to obtain the d-
and q-axis of the GSC output voltage. This voltage is developed by rewriting
Eqs. (3.33) and (3.34) as:
D
Vgd ¼ Vgd þ xe Lf igq þ Vds ð3:37Þ

D
Vgq ¼ Vgq  xe Lf igd ð3:38Þ

where
D digd
Vgd ¼ Rf igd þ Lf
dt
D digq
Vgq ¼ Rf igq þ Lf
dt

The angular position of the stator voltage (he ) is calculated as:


Z  
Vbs
he ¼ xe dt ¼ tan1 ð3:39Þ
Vas
30 3 A Modified MRAS Observer for Sensorless Control …

This angular position he is used to estimate the unit vector (cos he and sin he ) in
the stator coordinate transformation.

3.5 Speed Estimation Based on the Modified MRAS


Observer

A speed sensor is usually needed for vector control of the DFIG scheme. The use of
speed sensor has several drawbacks in term of robustness, cost, cabling, and
maintenance. A MRAS is one of the most popular adaptive control methods used in
sensorless control applications.
The classical MRAS scheme requires pure integration of the back e.m.f. This
leads to problems with initial conditions and drift. To avoid these problems, a
modified MRAS scheme is proposed. This scheme is based on the comparison
between the counter e.m.f obtained from stator equations (reference model) and
rotor equations (adjustable model) and adjusting the value of speed (x^ r ) in the rotor
equations for minimizing the resulting e.m.f error as shown in Fig. 3.4.
The stator equation used to obtain the counter e.m.f. in the stationary reference
frame (a  b) is given as [133]:
  " #
ema Vas Rs þ rLs ddt 0 ias
¼  ð3:40Þ
emb Vbs 0 Rs þ rLs ddt ibs

Vαs iαs
Actual machine model
Vβs iβs

Reference model em
(stator equation)

ω̂r +
-
iαs êm ew
Adjustable model
(rotor equation)
iβs
ω̂r
Adaptive mechanism

Fig. 3.4 Block diagram of MRAS observer


3.5 Speed Estimation Based on the Modified MRAS Observer 31

Similarly, the rotor equation used for estimating the counter e.m.f (^em ) in the
stationary reference frame (a  b) is given as:
 " #  !
^ema Lm  T1r xr kar Lm ias
¼ þ ð3:41Þ
^emb Lr xr  T1r kbr Tr ibs

The measured stator voltages and currents are used as inputs for two independent
models. The adjustment value of DFIG speed x ^ r is estimated from the error
between the reference and adjustable models outputs using a suitable adaptation
mechanism as:
 
KI
^ r ¼ KP þ
x ew ð3:42Þ
s
 
where ew ¼ ^ema emb  ^emb ema , KP and KI are PI parameters of speed estimator.
The stability of the modified MRAS is proved in Appendix II using
Lampooner’s stability theorem [134].

3.6 Configuration of Complete System

Figure 3.5 shows the schematic control structure of the GSC control. This scheme
begins with transforming the grid voltages Vsabc to the stationary reference frame
(Vas and Vbs ) to obtain the stator voltage vector angle he . The actual Dc-link voltage
across the capacitor Vdc can be controlled by controlling the d-axis grid current igd

in the SVO control. The Vdc is compared with its reference value Vdc to obtain d-

axis GSC reference current component igd through the PI voltage controller. The q-
axis GSC reference current value igq is set to zero to ensure zero reactive power
flow between the grid and the GSC.
The current-loop controllers are used to generate d- and q-axes control voltages
Vgd and Vgq , based on the error signals between the references and actual d- and q-

axes currents. The final d- and q-axes GSC voltage reference components Vgd and

Vgq are obtained by adding the compensation voltage terms due to cross decoupling
 
as shown in Fig. 3.5. The reference GSC voltage components Vgd and Vgq are then

transformed to three-phase GSC voltage references Vcabc using the stator voltage
vector angle he . The control inputs to the PWM block are the grid voltage refer-

ences Vcabc and predefined triangular carrier waves. In the PWM scheme, the
inverter output voltage is defined by the intersections of the grid voltage commands
and carrier waves, which are synchronized such that the carrier frequency is an
integer multiple of the frequency of grid voltage commands. This manner of syn-
chronization eliminates subharmonic generation.
32 3 A Modified MRAS Observer for Sensorless Control …

Fig. 3.5 Vector-control scheme based on SVO control for GSC of the wind driven DFIG system

Figure 3.6 shows the RSC control structure using the SFO control. The control
is implemented through a nested-loop structure consisting of an inner current loop
and an outer speed and reactive power loops. The speed reference x is generated
according to the maximum power extraction principle, while the reactive power
reference Q is based on a wind plant reactive power demand. The Q is transferred
to d-axis rotor current reference component idr through PI reactive power controller
and the x is transferred to q-axis rotor current reference component iqr through PI
speed controller.
The current-loop controllers are used to generate d- and q-axes rotor voltage
reference components Vdr and Vqr , based on the error signals between the actual and
their references d- and q-axes rotor current components. The final d- and q-axes

rotor voltage reference components Vdr and Vqr are obtained by adding the com-
pensation voltage terms due to cross decoupling as shown in Fig. 3.6. The reference

rotor voltage components Vdr and Vqr are then transformed to three-phase rotor

voltage references Vabcr using the slip angle hslip . The control inputs to the PWM

block are the rotor phase voltage references Vabcr and predefined triangular carrier
waves. In the PWM scheme, the inverter output voltage is defined by the inter-
sections of the rotor voltage commands and carrier waves, which are synchronized
such that the carrier frequency is an integer multiple of the frequency of rotor
voltage commands. This manner of synchronization eliminates subharmonic
generation.
3.7 Results and Discussions 33

Fig. 3.6 Vector-control scheme based on SFO control for RSC of the wind driven DFIG system

3.7 Results and Discussions

Digital simulations using MATLAB/Simulink software package are carried out in


order to investigate the performance of the proposed control schemes shown in
Figs. 3.5 and 3.6. The nominal parameters and data specifications of the DFIG
under study are listed in Appendix I
The simulation results of the proposed schemes are compared with the results
measured experimentally in literature [8].
Firstly, to study the performance of the proposed GSC control scheme shown in
Fig. 3.5, several tests are carried out in both transient and steady-state conditions
including bidirectional power flow with lagging, leading, and unity displacement
factors. The Dc-link voltage is regulated at 550 V, and the output voltage of the
converter is connected to grid (250 V supply).
Figure 3.7 shows the calculated instantaneous values of GSC phase voltage and
current with q-axis GSC current reference igq set to 0 for the converter operating in
rectifying mode which corresponds to subsynchronous speed operations of the
generator. The calculated instantaneous values of GSC phase voltage and current
for the inverting operation mode which corresponds to supersynchronous speed
operation is shown in Fig. 3.8 with q-axis GSC current reference igq also set to 0.
These figures show that, the phase displacement between the phase voltage and the
current is 0° during subsynchronous speed operation and 180˚ during supersyn-
chronous speed operation.
34 3 A Modified MRAS Observer for Sensorless Control …

Fig. 3.7 Grid voltage and


voltage current
current of GSC during
subsynchronous speed 300 6
operation of the DFIG system
200 4

Grid voltage (v)


100 2

Grid current (A)


0 0

-100 -2

-200 -4

-300 -6
0 0.01 0.02 0.03 0.04 0.05 0.06
Time (s)

Fig. 3.8 Grid voltage and


current voltage
current of GSC during
supersynchronous speed 300 3
operation of the DFIG system
200 2
Grid voltage (v)

Grid current (A)


100 1

0 0

-100 -1

-200 -2

-300 -3
0 0.02 0.04 0.06
Time (s)

Fig. 3.9 Calculated and


q-axis grid current
component (A)

measured [8] values of q-axis


GSC current

Time (s)

The transient performance of the proposed scheme for step change in reactive
component of grid current reference igq with power flowing from the supply to the
Dc-link is investigated.
Figure 3.9 shows the calculated and measured values of q-axis GSC current
component igq while waveforms of GSC phase voltage and current are shown in
Fig. 3.10. These figures illustrate that, the actual values of igq match their references
with fast and good response for significant step changes of igq (−4 to 4 A) at
t = 30 ms. Also, the change in phase angle between phase voltage and current from
3.7 Results and Discussions 35

Fig. 3.10 Supply phase voltage current


voltage and line current when

Supply voltage (v)


q-axis of GSC current

Line current (A)


200 10
reference stepped 5
form −4 to 4 A
0 0
-5
-200 -10

0 0.02 0.04 0.06 0.08


Time (s)

leading to lagging tacks place in one cycle. For comparison purpose, the calculated
igq response conforms to the measured one as disregards the settling time. The
above results demonstrate the capability of the GSC to supply reactive power to, or
receive reactive power from, the grid.
The transient performance of the proposed RSC control scheme for step change
in q-axis rotor current reference component iqr for supersynchronous speed oper-
ation is investigated.
Figure 3.11 shows the calculated and measured values of d-q axis rotor current
components idr , iqr , rotor phase current, and Dc-link voltage for large step change in
iqr to a 12 A (110% from its rated value) at t = 25 ms and removal at t = 250 ms
during supersynchronous operation with d-axis rotor current reference component
idr maintained at zero value. This figure shows that, the actual values of q-axis rotor
current component iqr change and matches their reference values and d-axis rotor
current component idr is maintained at its zero reference value. This confirms that
complete decoupling between active and reactive power control for the investigated
proposed speed sensorless control system is achieved. In addition to, this transient
response is considered the worst case for the GSC voltage control where dc link
power is stepped from zero to its rated value and vice versa. As shown the maxi-
mum error in dc link voltage is 20 V (4% of nominal) and fast recovery.
Consideration of iron loss and main flux saturation in machine model through the d-
q axis rotor current components idr , iqr amplitude and phase of the rotor phase
current values agree approximately with those measured one. Also, the dip and
overshoot of the Dc-link voltage are following the step change and removal of the
q-axis rotor current reference iqr . The Dc-link voltage dip and overshoot are
determined by the gains of PI voltage controller of the voltage loop.
Figure 3.12 shows the calculated and measured values of d-q axis rotor current
components idr , iqr and rotor phase current for step change in d-axis rotor current
reference component idr to 7 A (which corresponds to its rated value) at t = 50 ms
during subsynchronous speed operation with iqr maintained constant at 2 A. The
corresponding calculated values of stator phase current and voltage are shown in
Fig. 3.14. These figures show that the value of idr tracks the change in its reference
value idr and iqr is maintained at its reference value (2 A). This indicates that the
effectiveness of the proposed control scheme for independent control of active and
36 3 A Modified MRAS Observer for Sensorless Control …

d-q axis rotor current (A)


q- axis rotor current

d- axis rotor current


DC--Link voltage (V)
Rotor phase current (A)

Time (s)

Fig. 3.11 Calculated and measured [8] d-q axis rotor current component, Dc-link voltage and
rotor phase current when q-axis rotor current reference component is stepped from 0 to 12 A

reactive power. The ripples in the measured d-q axis rotor current components due
to the machine space harmonics which is disregarded in the machine model.
Consideration of core loss and main flux saturation in DFIG model provide good
agreement between the simulated values and experimental results. Also, as is evi-
dent from the phase displacement between stator phase current and voltage, the
magnetizing current is initially supplied from the stator until t = 50 ms, where the
phase shift changes to 180˚ showing that the machine is generating with all the
magnetizing current supplied from the rotor (Fig. 3.13).
The transient performance of the proposed sensorless control scheme for step
change in wind speed from 5 to 9 m/s and vice versa, with the d-axis rotor exci-
tation current component idr ¼ 0 is investigated.
3.7 Results and Discussions 37

d- axis rotor current

d-q axis rotor current (A)


q- axis rotor current
Rotor phase current (A)

Time (s)

Fig. 3.12 Calculated and measured [8] d-q axis rotor current and rotor phase current when d-axis
rotor current component reference is stepped from 0 to 7 A

Fig. 3.13 Stator voltage and voltage current


stator current when d-axis 300 12
rotor current reference
Stator voltage(V)

Stator current (A)


200 8
component is stepped
from 0 to 7 A 100 4
0 0
-100 -4
-200 -8
-300 -12
0 0.05 0.1 0.15
Time (s)

Figure 3.14 shows the calculated and measured values of d-q axis rotor current
components idr and iqr and the estimated rotor speed of the DFIG. From this figure,
it can be seen that the estimated generator speed values are agreed satisfactory with
those measured experimentally and the calculated d-q rotor current component
values are found to be close to those measured experimentally. This in addition to,
the idr is maintained at its reference value while iqr changes according to the change
of generator speed. This proves that, a modified MRAS observer for speed esti-
mation is appropriate for the sensorless vector control of the wind driven DFIG.
However, the measured d-q axis rotor current components have more ripples than
38 3 A Modified MRAS Observer for Sensorless Control …

Time (s)

Fig. 3.14 Measured [8] and calculated d-q axis rotor current and generator speed for step increase
(top) and step decrease (bottom) in wind speed

1.8
1.6
Generator speed (rpm)

Estimated
1.4 Measured
Optimal
1.2
1
0.8
0.6
0.4
4 5 6 7 8 9 10
Wind speed (m/s)

Fig. 3.15 Steady-state optimum speed tracking


3.7 Results and Discussions 39

the calculated one due to actual PWM inverter. Consideration of iron loss and main
flux saturation improves the theoretical prediction results.
Finally, comparisons between the estimated and experimental values of
steady-state speed using proposed sensorless scheme with the machine working on
the optimal power curve is shown in Fig. 3.15. From this figure, it can be noted that
the estimated values of the DFIG speed and measured one have slight deviation in
low values of wind speed. The discrepancy between the estimated and experimental
speed values, especially at high values of wind speed, is attributed due to the
mechanical loss which is disregarded in the machine model.
Chapter 4
Grid Synchronization Enhancement
of a Wind Driven DFIG Using Adaptive
Sliding Mode Control

4.1 Introduction

Besides power decoupled control, soft and fast synchronization is an important


issue because it enables the DFIG to be connected to the grid with minimum impact
on the WECS and the grid [135].
Grid synchronization enhancement of a wind driven Doubly Fed induction
generator (DFIG) using adaptive sliding mode control (SMC) is described and
evaluated in this chapter. The proposed scheme directly controls the stator terminal
voltage of the DFIG to track the grid voltage without current control loop; hence,
the structure of controller is simplified. For robustness of the control scheme,
parametric uncertainty and external disturbances are included into the formed
design procedure. A mathematical model of the DFIG as influenced by core loss is
considered to improve the theoretical prediction. Digital simulations are carried out
to demonstrate the effectiveness and robustness of the proposed scheme using
MATLAB/Simulink software package. Moreover, to validate the correctness and
accuracy of the proposed schemes the calculated performances are compared with
those results measured experimentally in the literature [135].

4.2 Description of the DFIG System

The schematic diagram of the DFIG-based WECS is shown in Fig. 4.1. The DFIG
is a wound rotor asynchronous machine mechanically coupled to a wind turbine.
The stator winding is connected to the grid through a three-phase contactor. The
rotor winding is connected to the grid through a bidirectional converter made up of
two back-to-back three-phase full-bridge inverters (referred as rotor converter and
grid converter). The rotor converter controls the voltage applied to the rotor

© Springer International Publishing AG 2018 41


A. Abdelbaset et al., Wind Driven Doubly Fed Induction Generator,
Power Systems, https://doi.org/10.1007/978-3-319-70108-0_4
42 4 Grid Synchronization Enhancement of a Wind Driven DFIG …

Wind Speed

3-phase
contactor
G .B . DFIG
GRID

RSC GSC
Dc-link
Filter

RSC GSC
control control

Fig. 4.1 DFIG-based WECS

winding of the DFIG. The grid converter controls the power flow between the
Dc-link and the grid to keep the voltage of capacitor in Dc-link constant.

4.3 Dynamic Model of a DFIG Taking Iron Losses


into Consideration

The dynamic model of the DFIG in d–q synchronous rotating reference frame,
taking the iron loss only into account, is shown in Fig. 4.2.
The differential equations of the above model were described in the previous
chapter (Sect. 3.3.2).

4.4 Direct Relationship Between Stator Voltage and Rotor


Voltage

During grid synchronization, the electrical equations of the DFIG in a synchronous


rotating reference frame disregards core loss are [136].

dkds
Vds ¼  xe kqs ð4:1Þ
dt
dkqs
Vqs ¼ þ xe kds ð4:2Þ
dt
4.4 Direct Relationship Between Stator Voltage and Rotor Voltage 43

ids Rs Lls Llr


(ωe −ωr )λqr Rr idr
ωe Lls iqs
idFe idm

Vds RFe Lm Vdr


ωeλqm

d-axis equivalent circuit


ωe Llsids (ωe − ωr )λdr
iqs Rs Lls Llr Rr iqr

iqFe iqm

Vqs RFe Lm Vqr

ωeλdm

q-axis equivalent circuit

Fig. 4.2 d–q equivalent circuit of the DFIG in synchronous rotating reference frame taking iron
loss into account

dkdr
Vdr ¼ Rr idr þ  ðxe  xr Þkqr ð4:3Þ
dt
dkqr
Vqr ¼ Rr iqr þ þ ðxe  xr Þkdr ð4:4Þ
dt

The stator and rotor flux linkages are given by:

kds ¼ Lm idr ð4:5Þ

kqs ¼ Lm iqr ð4:6Þ

kdr ¼ Lr idr ð4:7Þ

kqr ¼ Lr iqr ð4:8Þ

In the above equations, all the rotor parameters and variables are referred to the
stator side.
For the purpose to achieve direct voltage control, the direct relationship between
the stator voltage and the rotor voltage will be developed. Substituting Eqs. (4.7)
and (4.8) into Eqs. (4.3) and (4.4) and rearranging gives:
44 4 Grid Synchronization Enhancement of a Wind Driven DFIG …

didr Vdr Rr
¼  idr þ xslip iqr ð4:9Þ
dt Lr Lr

diqr Vqr Rr
¼  iqr  xslip idr ð4:10Þ
dt Lr Lr

where

xslip ¼ xe  xr

Substituting Eqs. (4.5) and (4.6) into Eqs. (4.1) and (4.2) and differentiating
them with respect to time yields:

dVds d2 idr diqr


¼ Lm 2  xe Lm ð4:11Þ
dt dt dt

dVqs d2 iqr didr


¼ Lm 2 þ xe Lm ð4:12Þ
dt dt dt

Substituting Eq. (4.10) into Eq. (4.11) and Eq. (4.9) into Eq. (4.12) and
arranging them in matrix form yields:
    " xe Lm Rr
# 
d Vds d2 idr xe xslip Lm Lr idr
¼ Lm 2 þ
dt Vqs dt iqr  xe LLmr Rr xe xslip Lm iqr
" #   ð4:13Þ
0  xLe Lr m Vdr
þ xL
e m
L 0 Vqr
r

4.4.1 External Disturbances and Parametric Uncertainty

There is unpredictable electromagnetic interference in operating locale, which may


cause unpredictable noises due to the sensors in the wind energy conversion system.
So the real voltage signals used by the controller are:
meas
Vdqg ¼ Vdqg þ DVdqg
meas
Vdqs ¼ Vdqs þ DVdqs
dqr ¼ idqr þ Didqr
imeas

where DVdqg and Didqr represent the unknown noises.


In addition, the machine parameters are obtained by identification experiments in
which errors are unavoidable, and furthermore, these parameters may vary with
ambient temperature, skin effect, and exciting saturation. Considering the
4.4 Direct Relationship Between Stator Voltage and Rotor Voltage 45

uncertainties of the machine parameters, it is assumed that the machine parameters


in Eq. (4.13) are bounded as follows:

Rr min \Rr ¼ Rro þ DRr \Rr max


Lr min \Lr ¼ Lro þ DLr \Lr max
Lm min \Lm ¼ Lmo þ DLm \Lm max

where Rro , Lro and Lmo denote the nominal values of machine parameters, and DRr
DLr and DLm denote their deviations.

4.5 Design of Proposed Adaptive Sliding Mode Control

To obtain high dynamic performance in a variable structure system, the system is


assumed to consist of several continuous subsystems, henceforth referred to as
structures, each of these structure can prove unacceptable from the point of view of
the quality of the control process, i.e., instability. While SMC is a robust control
scheme based on the concept of changing state of the system in order to obtain the
desired response (i.e., stability and fast response).
The control objective is to control the stator direct and quadrature voltage to
track the grid direct and quadrature voltages, respectively, such that the stator and
grid voltages will have equal magnitude, frequency, and phase. The sliding surfaces
of the proposed adaptive SMC for the stator direct and quadrature voltage control of
the DFIG are set as:

S ¼ ½ Sd Sq T ð4:14Þ

In order to maintain the enhanced transient response and minimize the


steady-state value of error, the switching surfaces can be in the integral forms [137].
8
>
> Zt
>
>
>
> Sd ¼ ed þ Kd ed ðsÞds þ edo
>
<
1
ð4:15Þ
>
> Zt
>
>
>
> Sq ¼ eq þ Kq eq ðsÞds þ eqo
>
:
1

where Kd and Kq are the positive control gains, edo and eqo are the initial errors
between the references and the actual values of d–q components of the DFIG stator
voltage, ed and eq are the respective errors between the references and the actual
values of d–q components of the DFIG stator voltage and are defined as:
46 4 Grid Synchronization Enhancement of a Wind Driven DFIG …


ed ¼ Vds  Vds ¼ Vds  Vdg ð4:16Þ

eq ¼ Vqs  Vqs ¼ Vqg  Vqs ð4:17Þ

The manifolds Sd ¼ 0 and Sq ¼ 0 represent the precise tracking of d–q compo-


nents of DFIG stator voltage. When the system state reaches the sliding manifolds,
the structure of the feedback loop is adaptively alter to slide the system state along
the sliding surface.
Differentiate Eq. (4.15) with respect to time yields:
8
> dSd ded d   
< ¼ þ Kd ed ¼ Vds  Vdg þ Kd Vds  Vdg
dt dt dt ð4:18Þ
: dSq ¼ deq þ K e ¼ d V  V  þ K V  V 
>
q q qg qs q qg qs
dt dt dt

Substituting Eq. (4.13) into Eq. (4.18) leads to:

dS
¼ F þ DU dqr ð4:19Þ
dt

where
    " x e Lm R r
# 
Fd d2 idr xe xslip Lm Lr idr
¼ Lm 2 þ x
Fq dt iqr e L
Lr
m Rr
xe xslip Lm iqr
  "  #
d Vdg Kd Vds  Vdg
þ þ  
dt Vqg Kq Vqg  Vqs

" #
0  xLe Lr m
D¼ and U dqr ¼ ½ Vdr Vqr  T
 xLe Lr m 0
The system uncertainties occur due to the deviation of the machine parameters
from their nominal values, and external disturbances may occur. Thus, Eq. (4.19)
can be modified as:

dS
¼ Fn þ Dn U dqr þ H ð4:20Þ
dt

where H ¼ ½ Hd Hq T represent the lumped uncertainty and Fn and Dn are the


nominal value of F and D.
In SMC, a Lyapunov approach is used for deriving conditions in the control law
that will drive the state orbit to the equilibrium manifold. The quadratic Lyapunov
function is selected as [138]:
4.5 Design of Proposed Adaptive Sliding Mode Control 47

1
W ¼ ST S  0 ð4:21Þ
2

The switch control law must be chosen so that the time derivative of W is
definitely negative. This can be assured if [75]
    
dW Hd Kd1 0 signðSd Þ
¼ ST  ð4:22Þ
dt Hq 0 Kq1 sign Sq
 
where Kd1 and Kq1 are the positive control gains, signðSd Þ and sign Sq are
respective switch functions for direct and quadrature voltage components and
defined as:
8
  < þ 1 Sj [ 0
sign Sj ¼ 0 Sj ¼ 0 ; j ¼ d; q ð4:23Þ
:
1 Sj \0

It is worth noting that the time derivative of Lyapunov function is still definitely
negative if the positive control gains fulfill the following condition:
 
Kd1 [ jHd j and Kq1 [ Hq 
where
8 9
>
> Lr Lmo  Lm Lro dVqg > >
>
> K q eq þ
meas
>
>
> xe Lm Lmo dt > >
>
>
> >
>
>
> þ ð  Þi þ x ð  Þi >
>
xe Lmo < R r R ro dr slip rL L ro qr =
Hd ¼ L d 2
i
Lro > > 
r qr
 Rro Didr þ slip Lro Diqr >
>
>
> x >
>
>
> e dt 2
>
>
>
> >
>
>
> L dDV L K De >
>
: ro qg

r q q
;
xe Lmo dt xe Lm
8 9
>
> Lr Lmo  Lm Lro dVdg >
>
>
> K e
d d
meas
 >
>
>
> x L L dt >
>
>
<
e m mo
>
=
xe Lmo L d 2
i
Hq ¼ r dr
Lro > > þ ðRr  Rro Þiqr þ xslip ðLr  Lro Þidr þ xe dt2 >
>
>
> >
>
>
> De >
d>
>
:  Rro Diqr  xslip Lro Didr þ
L ro dDV dg

L K
r d >
;
xe Lmo dt xe Lm

The time derivative of W on the state trajectories of Eq. (4.20) is given by:

dW 1 T dS dST  
¼ S þS ¼ ST Fn þ Dn U dqr þ H ð4:24Þ
dt 2 dt dt
48 4 Grid Synchronization Enhancement of a Wind Driven DFIG …

By equating Eqs.(4.22)–(4.24), the control law can be obtained as:


    
Fdn Kd1 0 signðSd Þ
U dqr ¼ D1
n þ ð4:25Þ
Fqn 0 Kq1 sign Sq

The switching action in (4.25) will cause chattering phenomenon, which can be
conquered by introducing boundary layer. The sign functions are replaced by sat-
uration functions sat ðSÞ in a small vicinity of the sliding surface, hence control
discontinuities and switching action in the control loop is avoided.

S
sat ðSÞ ¼
jSj þ k

where k is a small positive gain and jSj [ [ k.


However, the parameter variations and the load disturbances of the system are
difficult to be measured with exact values. Also selection of the upper bound of
lumped uncertainty has a significant effect on the control performance. If the bound
is small, the stability condition may not be satisfied and the controlled system may
become unstable. Therefore a simple adaptive algorithm is suggested to estimate the
upper bound of lumped uncertainties.

K^_ ¼ jSd j
 
^_ q1 ¼ Sq 
K
where K ^_ q1 are the time derivative of Kd1 and Kq1 , respectively.
^_ d1 and K
Therefore the control law will become:
   
Fdn ^ d1 sat ðSd Þ
K
U dqr ¼ D1 þ ^ ð4:26Þ
n Fqn Kq1 sat ðSq Þ

The sensitivity of the controlled system to variation of uncertainties and external


disturbances still exists in reaching phase, before sliding mode occurs, which means
the loss of robustness in the transient state. To guarantee the sliding mode in the
initial time instant, the initial conditions of the integrators should be chosen as:
eqo
Ido ¼ e
Kd and Iqo ¼
do
Kq
where Iqo and Ido are the initial conditions of the integrator and defined as:

Z0
Ido ¼ ed ðsÞds
1

,
4.5 Design of Proposed Adaptive Sliding Mode Control 49

Z0
Iqo ¼ eq ðsÞds
1

Hence, at t ¼ 0

Sdo ¼ edo þ Kd Ido ¼ 0 ð4:27Þ

Sqo ¼ eqo þ Kq Iqo ¼ 0 ð4:28Þ

Equations (4.27) and (4.28) mean that the system states are on the sliding
surfaces in the initial time instant without the reaching phase, and the complete
robustness can be obtained during the entire response.

4.6 Complete System of a Wind Generation


with Proposed SMC

Figure 4.3 shows the block diagram of the proposed adaptive SMC scheme. In this
scheme, the three-phase grid and stator voltages as well as rotor currents are
measured simultaneously. The measured three-phase grid voltage is transformed to
the stationary reference frame to obtain the grid voltage vector angle using the
following equation.
Z
Vbg
he ¼ xe dt ¼ tan1 ð4:29Þ
Vag

With the help of angle he , the stator and grid voltage components in syn-
chronously (d–q) rotating reference frame are obtained. The d- and q-axes com-
ponents of rotor current are obtained using the slip angle hslip ¼ he  hr , where hr is
the rotor angular position.
The d–q axis grid voltage components (Vdg ; Vqg ) are compared to the d–q axis
stator voltage components (Vds ; Vqs ), respectively. The resulting errors (ed ,eq ) are
processed by the adaptive SMC to generate the d–q axis rotor voltage reference

components (Vdr ,Vqr ). Using the slip angle hslip , the reference rotor voltage com-
ponents Vdr and Vqr are then transformed to three-phase rotor voltage references


Vabcr .
The control inputs to the PWM block are the three-phase rotor voltage references

Vabcr and predefined triangular carrier waves. In the PWM scheme, the inverter
output voltage is defined by the intersections of the rotor voltage commands and
carrier waves, which are synchronized such that the carrier frequency is an integer
multiple of the frequency of rotor voltage commands. This manner of
50 4 Grid Synchronization Enhancement of a Wind Driven DFIG …

Grid
Vα g ,Vβ g Vabcg
Vdg ,Vqg dq αβ
αβ abc
θe Voltage angle
calculation based
on eq. (4-29)

Vds ,Vqs dq Vabcs


abc

θr Position
DFIG model
taking core loss
Encoder
into account

θe θ slip
d dq iabcr
dt abc
Vds* idr , iqr
ωe , ωslip
ed θ slip
Vds Kd ∫
Adaptive Vdr* Vabcr*
SMC dq PWM
based on
* abc
Vqs eq Eq. (4-26) V
Kq ∫ qr

Vqs*

Fig. 4.3 Block diagram of a wind driven DFIG with proposed adaptive SMC

synchronization eliminates subharmonic generation. The output voltage of the


inverter is used as an input data to the rotor of DFIG model tacking core loss into
account.

4.7 Results and Discussions

Digital simulations using MATLAB/Simulink software package are carried out in


order to investigate the performance of the proposed control scheme shown in
Fig. 4.3. The nominal parameters and data specifications of the DFIG under study
are listed in Appendix C. Some of simulated results are compared with the results
measured experimentally in the literature [135].
The performance of the proposed adaptive SMC scheme for both subsyn-
chronous and supersynchronous speed is investigated. The grid voltage is applied at
t = 0.04 s.
Figure 4.4 shows the calculated and measured instantaneous values of the grid
and stator line voltages at 1600 rpm of rotor speed, which corresponds to super-
synchronous speed operations of the generator obtained from reference [135].
4.7 Results and Discussions 51

Line voltage (v)

Time (s)

Fig. 4.4 Measured [135] and calculated waveforms of the grid and stator line voltages during
supersynchronous speed operation of the DFIG system
Direct and quadrature voltage (v)

Time (s)

Fig. 4.5 Measured [135] and calculated direct and quadrature components of the grid and stator
voltages during supersynchronous speed of the DFIG system

The corresponding calculated and measured values of direct and quadrature voltage
components are shown in Fig. 4.5.
These Figures show that the values of instantaneous stator line voltage track the
values of grid voltage rapidly with the same amplitude, frequency, and phase after
0.04 s. The actual values of direct and quadrature components of stator voltage
match their references (direct and quadrature components of grid voltage) with fast
and good response. However, some oscillations appear in the measured values of
direct and quadrature stator voltage components due to actual PWM inverter.
Consideration of core loss in DFIG model provides good agreement between the
calculated values and those measured experimentally. On the other hand, there is a
significant deviation of the measured values from those calculated.
Figure 4.6 shows the calculated instantaneous values of the grid and stator line
voltages at 1400 rpm of rotor speed, which corresponds to subsynchronous speed
operations of the generator. The corresponding calculated values of direct and
quadrature voltage components are shown in Fig. 4.7.
52 4 Grid Synchronization Enhancement of a Wind Driven DFIG …

Vg Vs (1400 rpm)

500

Line voltage (v)


0

-500
0 0.05 0.1 0.15 0.2
Time (s)

Fig. 4.6 Calculated waveforms of the grid and stator line voltages during subsynchronous speed
operation of the DFIG system
Direct and quadrature voltage (v)

300

V dg
200
V qg
V ds (1400 rpm)
100
V qs (1400 rpm)

-100
0 0.05 0.1 0.15 0.2
Time (s)

Fig. 4.7 Calculated direct and quadrature components of the grid and stator voltages during
subsynchronous speed operation of the DFIG system

From Figs. 4.6 and 4.7 we can see that the voltage responses at subsynchronous
speed are very similar to the voltage responses at supersynchronous speed.
This confirms that the proposed adaptive SMC schemes are capable of syn-
chronizing the DFIG at both subsynchronous and supersynchronous speed.
The transient performances are investigated for comparing the proposed adaptive
SMC with conventional PI control schemes against external disturbances. White
Gaussian noise with a power spectral density 0.005 (unit) is added to the measured
current and voltage signals at 1600 rpm of rotor speed.
Figure 4.8 shows the calculated and measured instantaneous values of the grid
and stator line voltages using adaptive SMC. The corresponding calculated and
measured values of direct and quadrature voltage components are shown in
Fig. 4.9.
Figure 4.10 shows the calculated and measured waveforms of the grid and the
stator line voltages with conventional PI control scheme. The corresponding cal-
culated and measured values of direct and quadrature voltage components are
shown in Fig. 4.11.
4.7 Results and Discussions 53

Line voltage (v)

Time (s)

Fig. 4.8 Measured [135] and calculated waveforms of the grid and stator line voltages under
disturbances using adaptive SMC
Direct and quadrature voltage (v)

Time (s)

Fig. 4.9 Measured [135] and calculated direct and quadrature components of the grid and stator
voltages under disturbances using adaptive SMC
Line voltage (v)

Time (s)

Fig. 4.10 Measured [135] and calculated waveforms of the grid and stator line voltages under
disturbances using conventional PI control
54 4 Grid Synchronization Enhancement of a Wind Driven DFIG …

Direct and quadrature voltage (v)

Time (s)

Fig. 4.11 Measured [135] and calculated direct and quadrature components of the grid and stator
voltages under disturbances using conventional PI control

Figures 4.8, 4.9, 4.10, 4.11 show that the stator voltages of the DFIG contain
large amplitude high frequency noises when conventional PI control is used, but the
proposed synchronization scheme can suppress the noises and has better robustness.
For comparison purpose, the values of calculated results when taking iron loss into
account in the machine model agree satisfactory with those results measured
experimentally taken from reference [135]. This in addition to, the measured d–
q axis stator voltage components has more ripples than the calculated one.
The above results demonstrate the robustness of the proposed adaptive SMC
against external disturbances.
The transient performances of the proposed adaptive SMC and conventional PI
control schemes against parametric uncertainty are investigated.
Figure 4.12 shows the calculated values of the grid, the stator waveforms, and
the corresponding d–q axis voltage components with the rotor self-inductance of the
DFIG is changed to 50% of its nominal value. Also, the identical calculated-type
voltage with the rotor winding resistance of the DFIG is changed by 50% of
nominal value and is shown in Fig. 4.13.
4.7 Results and Discussions 55

Line voltage (v)

Time (s)
350
Direct and quadrature voltage (v)

300
Vqg
250
Vdg
200 Vds(with PI controller)

150 Vqs(with PI controller)


Vds( with adaptive SMC)
100
Vqs(with adaptive SMC)
50

-50
0 0.05 0.1 0.15 0.2
Time (s)

Fig. 4.12 Calculated grid, stator line voltages, and the corresponding direct and quadrature
voltages with rotor self-inductance variation

It can be seen that the amplitude of stator voltage using PI control scheme was
observed to be higher than that of reference one (grid voltage). The adaptive SMC
scheme provides fast and good matching between calculated values of direct and
quadrature components of stator voltage and their references (direct and quadrature
components of grid voltage).
56 4 Grid Synchronization Enhancement of a Wind Driven DFIG …

Vg Vs (PI controller) Vs (adaptive SMC)

Line voltage (v) 500

-500
0 0.05 0.1 0.15 0.2
Time (s)
350

300
Direct and quadrature voltage (v)

Vqg
250
Vdg
200 Vds (with PI controller)
150 Vqs (with PI controller)

100 Vds (with adaptive SMC)


Vqs (with adaptive SMC)
50

-50

-100
0 0.05 0.1 0.15 0.2
Time (s)

Fig. 4.13 Calculated grid, stator line voltages, and the corresponding direct and quadrature
voltages with rotor resistance variation

Also, it can be seen that the responses shown in Figs. 4.12 and 4.13 (using
adaptive SMC) are almost the same as the responses shown in Figs. 4.4 and 4.5,
respectively. Hence, the system response depends only on the predefined sliding
surface and remains insensitive to variations of machine parameters. This confirms
the parametric robustness of the proposed adaptive SMC scheme.
Chapter 5
Adaptive Sliding Mode Control for Grid
Synchronization of a Wind Driven DFIG
Under Unbalanced Grid Voltage

5.1 Introduction

However, wind energy conversion system (WECS) is usually located in rural areas
with weak grid connection, in which grid voltage unbalance may arise even during
normal operation. The unbalanced grid voltage may be caused by unbalanced
transmission line impedance, three-phase unbalanced load, and single-phase
high-power load. Hence, the strategies to improve the performance of DFIGs
under unbalanced network conditions have obtained a worldwide concern.
Adaptive sliding mode control (SMC) for grid synchronization of a wind driven
Doubly Fed induction generator (DFIG) under unbalanced grid voltage is described
and evaluated in this chapter. The proposed scheme consists of positive sequence
controller based on adaptive SMC and negative sequence controller based on
integral control. The positive sequence controller directly controls the positive
sequence stator voltage to track the positive sequence grid voltage, whereas the
negative sequence controller directly controls the negative sequence stator voltage
to track the negative sequence grid voltage. Thus, no extra current control loops
are required, thereby simplifying the design of the controller. Digital simulations
are carried out to demonstrate the effectiveness and robustness of the proposed
scheme using MATLAB/Simulink software package. Moreover, to validate the
correctness and accuracy of the proposed scheme, some calculated performances
are compared with those results measured experimentally in the literature [139].

5.2 DFIG Model During Grid Synchronization Under


Unbalanced Grid Voltage

Based on symmetrical component theory, all unbalanced voltage, current, and flux
vectors in the stationary reference frame can be represented as the superposition of
positive, negative, and zero sequence components. Since the neutral points of DFIGs
© Springer International Publishing AG 2018 57
A. Abdelbaset et al., Wind Driven Doubly Fed Induction Generator,
Power Systems, https://doi.org/10.1007/978-3-319-70108-0_5
58 5 Adaptive Sliding Mode Control for Grid Synchronization …

Fig. 5.1 Relationships βs


between as bs ; ar br ; dq þ , βr
and r dq reference frames q+ F q−

ωe d+

θe = ωe t αr

θ r = ωr t
αs
θe = − ωet

−ωe
d−

are not connected to grids, all the zero sequence components are inexistent. Hence, in
 can
the stationary reference frame, the unbalanced voltage, current, and flux vectors F
be decomposed into positive and negative sequence components as follows [140]:
   
F  ab ðtÞ ¼ F
 ab þ ðtÞ þ F
 ab ðtÞ ¼ F  ab þ ejðxe t þ / þ Þ þ F ab ejðxe t þ / Þ ð5:1Þ

where / þ and / are the respective phase shift for positive and negative sequence
components.
Figure 5.1 shows the spatial relationship between the stator stationary (as bs )
reference frame, rotor (ar br ) frame rotating at the angular speed of xr , dq þ and
dq reference frames rotating at the synchronous speeds xe and xe .
From Fig. 5.1, the coordinate transformation among the stationary, positive
sequence synchronous, and negative sequence synchronous reference frames is:

þ ¼ F
F  ab s ejxe t
dq
 ¼ F
F  ab s ejxe t
dq
ð5:2Þ
Fþ ¼ F
  ej2 xe t
dq dq

F ¼  þ j2 xe t
Fdq e
dq

During grid synchronization, the positive sequence model of the DFIG in the
positive sequence synchronous reference frame ðdq þ Þ is:
þ
dkds þ þ
Vdsþþ ¼  xe kqs þ ð5:3Þ
dt
þ
dkqs þ þ
Vqsþþ ¼ þ xe kds þ ð5:4Þ
dt
5.2 DFIG Model During Grid Synchronization … 59

dkdrþ þ
Vdrþþ ¼ Rr idrþ þ þ  xslip þ kqrþ þ ð5:5Þ
dt

dkqrþ þ
Vqrþþ ¼ Rr iqrþ þ þ þ xslip þ kdrþ þ ð5:6Þ
dt

where xslip þ ¼ ðxe  xr Þ.


The stator and rotor positive sequence flux linkages are given by:
þ þ
kds þ ¼ Lm idr þ ð5:7Þ
þ þ
kqs þ ¼ Lm iqr þ ð5:8Þ

kdrþ þ ¼ Lm idrþ þ ð5:9Þ

kqrþ þ ¼ Lm iqrþ þ ð5:10Þ

In the negative sequence synchronous reference frame ðdq Þ, the negative model
of the DFIG during grid synchronization is:

dk

Vds ¼ ds
þ xs k
qs ð5:11Þ
dt
dk

 x s k
qs
Vqs ¼ ds ð5:12Þ
dt
dk

Vdr ¼ Rr i
dr þ
dr
 xslip k
qr ð5:13Þ
dt
dk

¼ Rr i þ xslip k
qr
Vqr qr þ dr ð5:14Þ
dt

where xslip ¼ ðxe  xr Þ.


The stator and rotor negative sequence flux linkages are given by:

k 
ds ¼ Lm idr ð5:15Þ

k 
qs ¼ Lm iqr ð5:16Þ

k 
dr ¼ Lm idr ð5:17Þ

k 
qr ¼ Lm iqr ð5:18Þ

The q-axes of the positive and negative sequence synchronous reference frames
are aligned to the positive and negative sequence components of grid voltage
60 5 Adaptive Sliding Mode Control for Grid Synchronization …

vector, respectively, and hence, the d-axes are orthogonal to respective components
of grid voltage vector, which gives:
    
Vqgþ þ ¼ Vg þ ; Vqg ¼ Vg  and Vdg
þ 
¼ Vdg ¼0 ð5:19Þ

According to the model of the DFIG under unbalanced grid voltage, two con-
trollers are needed to regulate stator voltages to achieve the same magnitude, fre-
quency, and phase with unbalanced grid voltages.
The positive sequence controller
The negative sequence controller

5.3 Positive Sequence Controller Design

The objective of the positive sequence controller is to regulate the positive sequence
stator direct and quadrature voltages to track corresponding positive sequence grid
voltages in a positive sequence synchronous reference frame, respectively.

5.3.1 Controlled Plant Analysis

In order to choose appropriate input–output control pairs in DFIG under unbalanced


grid voltage, the relative gain array (RGA) methodology [141] is used to calculate
the degrees of relevance between the input and output variables.
The RGA is calculated as the element by element product of the system transfer
matrix and the inverse of its transposed matrix. An RGA element close to unity
indicates that the input and output variables constitute a suitable pair that forms the
formation of a control loop, while a small positive RGA element shows a low
correlation between the input and output variables.
Using Eqs. (5.3–5.10), the Laplace equations relating the positive sequence
variables are:
    
Vdsþþ Lm s xe Lm idrþ þ
¼ ð5:20Þ
Vqsþþ xe Lm Lm s iqrþ þ
    
Vdrþþ Rr þ Lr s xslip þ Lr idrþ þ
¼ ð5:21Þ
Vqrþþ xslip þ Lr Rr þ Lr s iqrþ þ
5.3 Positive Sequence Controller Design 61

The RGA of the matrix in Eq. (5.20) is:

   T !1  2 
Lm s xe Lm Lm s xe Lm 1 s x2e
: ¼ 2 ð5:22Þ
xe Lm Lm s xe Lm Lm s s þ x2e x2e s2

where : means element by element multiplication.


Since the RGA is only used to analyze the DFIG model, it is sufficient to
evaluate at zero frequency to give valid input–output pairs at steady state [141],
which means that s terms are zero. Since the values of the diagonal elements are
zero and the values of the off-diagonal elements are unit in Eq. (5.22). The
appropriate input–output control pairs in Eq. (5.20) should be direct rotor current to
quadrature stator voltage and quadrature rotor current to direct stator voltage.
The RGA of the matrix in Eq. (5.24) is:

    !1
Rr þ Lr s xslip þ Lr xslip þ Lr T
Rr þ Lr s
:
xslip þ Lr Rr þ Lr s xslip þ Lr
R r þ Lr s
"  2 # ð5:23Þ
1 ðRr þ Lr sÞ2 xslip þ Lr
¼  2  2
ðRr þ Lr sÞ2 þ xslip þ Lr xslip þ Lr ðRr þ Lr sÞ2

If the DFIG operates at synchronous speed values of the diagonal elements are
unit and values of the off-diagonal elements are zero in Eq. (5.23). Hence, the
appropriate input–output control pairs in Eq. (5.21) should be direct rotor voltage to
direct rotor current and quadrature rotor voltage to quadrature rotor current.
According to the RGA analysis, the appropriate input–output control relation-
ship among positive sequence variables in a positive sequence synchronous refer-
ence frame is:

Vdrþþ ) idrþ þ ) Vqsþþ
Vqrþþ ) iqrþ þ ) Vdsþþ

5.3.2 Direct Relationship Between Positive Sequence Stator


and Rotor Voltages

To simplify the structure of the controller, the direct relationship between the
positive sequence stator and rotor voltages will be developed. Substituting
Eqs. (5.9) and (5.10) into Eqs. (5.5) and (5.6), respectively, and rearranging gives:

didrþ þ Vdrþþ Rr þ
¼  idr þ þ xslip þ iqrþ þ ð5:24Þ
dt Lr Lr
62 5 Adaptive Sliding Mode Control for Grid Synchronization …

diqrþ þ Vqrþþ Rr þ
¼  i  xslip þ idrþ þ ð5:25Þ
dt Lr Lr qr þ

Substituting Eqs. (5.7) and (5.8) into Eqs. (5.3) and (5.4), respectively, and
differentiating them with respect to time yields:

dVdsþþ d2 i þ diqrþ þ
¼ Lm dr2 þ  xe Lm ð5:26Þ
dt dt dt

dVqsþþ d2 iqrþ þ didrþ þ


¼ Lm þ xe Lm ð5:27Þ
dt dt2 dt

Substituting Eqs. (5.24) and (5.25) into Eqs. (5.26) and (5.27), respectively, and
arranging them in matrix form yield:
" # " # " xe Lm Rr
#" # " #" #
þ
d Vds þ
þ
d2 idr þ xe xslip þ Lm Lr idrþ þ 0  xLe Lr m Vdrþþ
þ ¼ L þ þ þ
iqrþ þ Vqrþþ
m
dt Vqs þ 2
dt iqr þ  xe LLm Rr r
xe xslip þ Lm xe Lm
Lr 0
ð5:28Þ

5.3.3 Parametric Uncertainty

The machine parameters are obtained by identification experiments in which errors


are unavoidable, and furthermore, these parameters may vary with ambient tem-
perature, skin effect, and exciting saturation. Considering the uncertainties of the
machine parameters, it is assumed that the machine parameters in Eq. (5.28) are
bounded as follows:

Rr min \Rr ¼ Rro þ DRr \Rr max

Lr min \Lr ¼ Lro þ DLr \Lr max

Lm min \Lm ¼ Lmo þ DLm \Lm max

where Rro , Lro and Lmo denote the nominal values of machine parameters, and DRr ,
DLr and DLm denote their deviations.

5.3.4 Adaptive Sliding Mode Control Design

Sliding mode controller enforces the system state on the predefined sliding surface
in the system state space by adaptively changing the structure of the controller.
5.3 Positive Sequence Controller Design 63

Hence, the system response depends only on the sliding surface and is insensitive to
variations of system parameters and external disturbances [142].
The sliding surfaces are chosen in the integral forms as:

S þ ¼ ½ Sd þ Sq þ  T ð5:29Þ

where
8
> Rt
>
< Sd þ ¼ ed þ þ Kd þ ed þ ðsÞds þ edo þ
1
Rt ð5:30Þ
>
>
: Sq þ ¼ eq þ þ Kq þ eq þ ðsÞds þ eqo þ
1

where Kd þ and Kq þ are the positive control gains, ed þ and eq þ are the respective
errors between d-q positive sequence grid and stator voltage components and edo þ
and eqo þ the initial error.
The ed þ and eq þ are defined as:

ed þ ¼ Vdsþþ  Vdsþþ ¼ Vdsþþ  Vdg


þ
þ ð5:31Þ

eq þ ¼ Vqsþþ  Vqsþþ ¼ Vqgþ þ  Vqsþþ ð5:32Þ

The manifolds Sd þ ¼ 0 and Sq þ ¼ 0 represent the precise tracking of d-q pos-


itive sequence components of DFIG stator voltage. When the system state reaches
the sliding manifolds, the structure of the feedback loop is adaptively alter to slide
the system state along the sliding surface.
Differentiate Eq. (5.30) with respect to time, this yield:
8  
< dSd þ ¼ ded þ þ Kd þ ed þ ¼ d Vdsþþ  Vdg
þ
þ þ Kd þ Vdsþþ  Vdg
þ
þ
dt dt dt   ð5:33Þ
: dSq þ ¼ deq þ þ Kq þ eq þ ¼ d V þ  V þ þ Kq þ V þ  V þ
dt dt dt qg þ qs þ qg þ qs þ

Substituting (5.28) into (5.33) leads to:

dS þ
¼ F þ þ DU dqr þ ð5:34Þ
dt
64 5 Adaptive Sliding Mode Control for Grid Synchronization …

where
  " # " x e L m Rr
#" # " #
þ
Fd þ d2 idr þ xe xslip þ Lm Lr idrþ þ þ
d Vdg þ
¼ Lm 2 þ þ
Fq þ dt iqrþ þ x e L m Rr
Lr xe xslip þ Lm iqrþ þ dt Vqgþ þ
2  3
þ þ
K V  Vdg þ
6 d ds þ 7
þ4  5
Kq Vqgþ þ  Vqsþþ

" #
0  xLe Lr m
and U dqr þ ¼ ½ Vdrþþ Vqrþþ 
T

 xLe Lr m 0

The system uncertainties occur due to the deviation of the machine parameters
from their nominal values and external disturbances may occur. Thus, Eq. (5.34)
can be modified as:

dS þ
¼ Fn þ þ Dn U dqr þ þ H þ ð5:35Þ
dt

where H þ ¼ ½ Hd þ Hq þ T represents the lumped uncertainty and Fn þ and Dn


are the nominal value of F þ and D, respectively.
In SMC, a Lyapunov approach is used for deriving conditions in the control law
that will drive the state orbit to the equilibrium manifold. The quadratic Lyapunov
function is selected as [138]:

1
W þ ¼ STþ S þ  0 ð5:36Þ
2

The switch control law must be chosen so that the time derivative of W þ is
definitely negative. This can be assured if:
    
dW þ Hd þ Kd1 þ 0 sign ðSd þ Þ
¼ STþ  ð5:37Þ
dt Hq þ 0 Kq1 þ sign Sq þ
 
where Kd1 þ and Kq1 þ are the positive control gains, signðSd þ Þ and sign Sq þ are
respective switch functions for direct and quadrature positive voltage components
and defined as:
8
  < þ1 Sj þ [ 0
sign Sj þ ¼ 0 Sj þ ¼ 0 ; j ¼ d; q ð5:38Þ
:
1 Sj þ \0

It is worth noting that the time derivative of Lyapunov function is still definitely
negative if the positive control gains fulfill the following condition:
5.3 Positive Sequence Controller Design 65
 
Kd1 þ [ jHd þ j and Kq1 þ [ Hq þ 
where
8 9
> Lr Lmo  Lm Lro dVqgþ þ Lr d iqrþ þ
2
>
xe Lmo < Kq þ eq þ þ  =
Hd þ ¼ xe Lm Lmo dt xe dt2
Lro > : þ þ
>
;
þ ðRr  Rro Þ idr þ  xslip þ ðLr  Lro Þ iqr þ
8 ! 9
> þ >
> L L  L L ddVdg Lr d2 idrþ þ >
xe Lmo < =
r mo m ro þ
Kd þ ed þ  þ
Hq þ ¼ xe Lm Lmo dt xe dt2
Lro > > >
>
: þ ðR  R Þ i þ þ x þ ;
r ro qr þ slip þ ðLr  Lro Þ idr þ

The time derivative of W þ on the state trajectories of Eq. (5.35) is given by:

dW þ 1 T dS þ dSTþ  
¼ Sþ þ Sþ ¼ STþ Fn þ þ Dn U dqr þ þ H þ ð5:39Þ
dt 2 dt dt

By equating (5.37–5.39), the control law can be expressed as:


    
Fdn þ Kd1 þ 0 signðSd þ Þ
U dqr þ ¼ D1
n þ ð5:40Þ
Fqn þ 0 Kq1 þ sign Sq þ

As in previous chapter, to avoid control discontinuities and switching action in


the control loop, the sign functions are replaced by saturation functions satðSÞ in a
small vicinity of the sliding surface.
However, the parameter variations and the load disturbances of the system are
difficult to be measured with exact values. Also selection of the upper bound of
lumped uncertainty has a significant effect on the control performance. If the bound
is small, the stability condition may not be satisfied and the controlled system may
become unstable. There for a simple adaptive algorithm is suggested to estimate the
upper bound of lumped uncertainties.

^ ¼ j Sd þ j
K
d1 þ

  
^ ¼  Sq þ 
K
q1 þ

 
^ and K
where K ^ are the time derivative of Kd1 þ and Kq1 þ , respectively.
d1 þ q1 þ
66 5 Adaptive Sliding Mode Control for Grid Synchronization …

Therefore, the control law will become:


    
Fdn þ ^ d1 þ
K 0 satðSd þ Þ
U dqr þ ¼ D1
n þ ^ q1 þ ð5:41Þ
Fqn þ 0 K sat Sq þ

The sensitivity of the controlled system to variation of uncertainties and external


disturbances still exists in reaching phase, before sliding mode occurs, which means
the loss of robustness in the transient state. To guarantee the sliding mode in the
initial time instant, the initial conditions of the integrators should be chosen as:
edo þ
Ido þ ¼ ;
Kd þ
eqo þ
Iqo ¼
Kq þ

where Iqo þ and Ido þ are the initial conditions of the integrator defined as:

Z0
Ido þ ¼ ed þ ðsÞds;
1
Z0
Iqo þ ¼ eq þ ðsÞds
1

Hence at t ¼ 0

Sdo þ ¼ edo þ þ Kd þ Ido þ ¼ 0 ð5:42Þ

Sqo þ ¼ eqo þ þ Kq þ Iqo þ ¼ 0 ð5:43Þ

Equations (5.42) and (5.43) mean that the system states are on the sliding sur-
faces in the initial time instant without the reaching phase, and the complete
robustness can be obtained during the entire response.

5.4 Negative Sequence Controller Design

The objective of the negative sequence controller is to regulate the negative


sequence stator direct and quadrature voltages to track corresponding negative
sequence grid voltages in a negative sequence synchronous reference frame,
respectively.
Substituting Eqs. (5.15) and (5.16) into Eqs. (5.11) and (5.12), respectively, and
arranging them in matrix form yield:
5.4 Negative Sequence Controller Design 67

  " # 

Vds Lm ddt xe Lm i
 ¼ dr
ð5:44Þ
Vqs xe Lm Lm dtd i
qr

Also, substituting Eqs. (5.17) and (5.18) into Eqs. (5.13) and (5.14), respec-
tively, and arranging them in matrix form yield:
  " # 

Vdr Rr þ Lr ddt Lr xslip i
 ¼ dr
ð5:45Þ
Vqr Lr xslip Rr þ Lr dtd i
qr

According to the RGA analysis, the appropriate input–output control relation-


ship among negative sequence variables in a negative sequence synchronous ref-
erence frame is:
 
Vqr ) i 
dr ) Vqs

Vdr ) i
qr ) V 
ds

Also, the diagonal terms of matrices in Eqs. (5.44) and (5.45) can be regarded as
disturbances and neglected, these yield:
 
   
Vds 0 xe Lm i
 ¼ dr
ð5:46Þ
Vqs xe Lm 0 i
qr
 
   
Vdr 0 Lr xslip i
 ¼ dr
ð5:47Þ
Vqr Lr xslip 0 i
qr

Substituting (5.46) into (5.47) and rearranging gives:


 
  

Vds Lm xe Vdr
 ¼  ð5:48Þ
Vqs Lr xslip Vqr

Because there are no transient terms in Eq. (5.48), integral controllers are
enough to achieve the first order dynamic responses in a closed-loop system, as
shown in (5.49) and (5.50):
Z 
  
Vdr ¼ Kd Vdg  Vds dt ð5:49Þ

Z 
  
Vqr ¼ Kq Vqg  Vqs dt ð5:50Þ

where Kd and Kq are the constants of the integral controllers.
68 5 Adaptive Sliding Mode Control for Grid Synchronization …

5.5 Configuration of the Complete Control Scheme

Figure 5.2 shows the block diagram of the proposed grid synchronization control
scheme of the DFIG under unbalanced voltage.
In this scheme, the three-phase grid, stator voltages and rotor currents are
measured simultaneously as well as the rotor phase angle. The measured
three-phase grid and stator voltages are transformed to the stationary reference
frame. With the help of angle hr , the rotor current components in stationary ref-
erence frame are obtained. The delay angle cancellation method proposed in [143]
is used to obtain positive and negative sequence components as:
2 3 2 32 3
F a þ ðt Þ 1 0 0 1 Fa ðtÞ
6 Fb þ ðtÞ 7 1 6 0 1 1 0 7 6 Fb ðtÞ 7
6 7 6 76 7
4 Fa ðtÞ 5 ¼ 2 4 1 0 0 1 5 4 Fa ðt  ðp=2xs ÞÞ 5
ð5:51Þ
Fb ðtÞ 0 1 1 0 Fb ðt  ðp=2xs ÞÞ

The positive sequence grid voltage vector angle hs þ is calculated using positive
sequence components of the grid voltage in the stationary reference frame.

Grid

DFIG model
taking
Vabcg Vabcs core loss
into account iabcr
Vβ g +
abc tan −1
abc
Vα g + d dt ωe
αβ αβ abc iαβ r
θr
Vαβ g iαβ r ω Vαβ s θs + Position αβ
e Encoder
Vdg+ + θs +
Vαβ g + αβ+ +
V ed +
Positive and
Negative iαβ r +
ds +
Kd + ∫ θr
Sequence +
idqr +
θslip + PWM
separation Vαβ s + Adaptive
SMC Vdr+ +
Eq.(5-47)
dq+ dq+
Vqg+ + Vqr+ +
+
Eq. (5-
37) αβ+ *
Vabcr
Vαβ g − Vαβ s − V qs + eq +
Vds− − Kq+ ∫ Vαβ r +
θr abc
αβ− V −
dg − − αβ
− V dr −
V Kd − ∫ dq− Vαβ r −
θs −
qg −
αβ− ∑
dq− Vqs− − Vαβ r
− θ slip −
V
Vβ g − Kq− ∫ qr −
θr
tan −1
Vα g − θs −
θs −

Fig. 5.2 Block diagram of proposed synchronization scheme of the DFIG under unbalanced grid
voltages
5.5 Configuration of the Complete Control Scheme 69

With the help of hs þ , the d- and q-axes positive sequence components of the
grid and stator voltages and rotor current in positive synchronous reference frame
dq þ are obtained. The d-q axis grid voltage positive sequence components in dq þ
þ þ þ þ
ðVdg þ ; Vqg þ Þ are compared with those of stator voltage ðVds þ ; Vqs þ Þ, respec-
tively. The resulting errors (ed þ , eq þ ) are processed by the adaptive SMC to
þ
generate the d-q axis positive sequence components  of rotor voltage  in dq
þ þ
(Vdr þ ,Vqr þ ). Using the positive sequence slip angle hslip þ ¼ hs þ  hr , the rotor
voltage components Vdrþþ and Vqrþþ are then transformed to rotor voltage positive
 
sequence components in stationary reference frame Var þ ; Vbr þ .
Similarly, the negative sequence grid voltage vector angle hs is calculated using
negative sequence components of the grid voltage in the stationary reference frame.
Using hs , the d- and q-axes negative sequence components of the grid and stator
voltages in negative synchronous reference frame dq are obtained. The d-q axis
grid voltage negative sequence components in dq ðVdg  
; Vqg Þ are compared with
 
those of stator voltage ðVds ; Vqs Þ, respectively. The resulting errors are processed
by integral controllers to generate the d-q axis negative sequence components of
rotor voltage in dq ðVdr  
; Vqr Þ. Using the negative sequence slip angle
   
hslip ¼ hs  hr , the rotor voltage components Vdr and Vqr are then trans-
formed
 to rotor
 voltage negative sequence components in stationary reference frame
Var ; Vbr .

The reference rotor voltage components in stationary reference frame Var ; Vbr 
   
are obtained by adding positive Var þ ; Vbr þ and negative Var ; Vbr sequence
components of rotor voltage. Using the rotor phase angle hr , the reference rotor
 
voltage components Var and Vbr are then transformed to three-phase rotor voltage

references Vabcr . The control inputs to the PWM block are the rotor phase voltage

references Vabcr and predefined triangular carrier waves. In the PWM scheme, the
inverter output voltage is defined by the intersections of the rotor voltage com-
mands and carrier waves, which are synchronized such that the carrier frequency is
an integer multiple of the frequency of rotor voltage commands. This manner of
synchronization eliminates subharmonic generation. The output voltage of the
inverter is used as an input data to the rotor of DFIG model tacking core loss into
account as presented in the previous chapter.

5.6 Results and Discussions

Digital simulations using MATLAB/Simulink software package are carried out in


order to investigate the performance of the proposed control scheme shown in
Fig. 5.3. The nominal parameters and data specifications of the DFIG under study
are listed in Appendix C. Some of simulated results are compared with those
measured experimentally in literature [139].
70 5 Adaptive Sliding Mode Control for Grid Synchronization …

Fig. 5.3 Calculated and

Pos-sequence voltage (v)


measured [139] d-q positive
sequence grid and stator
voltages during
supersynchronous speed

Time (s)

The validity of the proposed grid synchronization control scheme of the DFIG
under unbalanced voltage for supersynchronous and subsynchronous speed is
investigated.
Firstly, magnitudes of A and B phase voltages are kept at the rated value while
magnitude of C phase voltage is reduced to provide a three-phase unbalanced grid
voltage. The ratio of negative sequence voltage to positive sequence voltage is 10%.
The grid voltage is applied at t = 0.02 s, and the rotor speed is 1600 rpm
(supersynchronous).
Figure 5.3 shows the calculated and measured values (obtained from reference
[139]) of q-axis positive sequence grid voltage and d-q axis positive sequence stator
voltage. This figure shows that, positive sequence stator voltages follow positive
sequence grid voltages with the adaptive SMC in a positive sequence synchronous
reference frame.
Figure 5.4 shows the calculated and measured values of q-axis negative
sequence grid voltage and d-q axis negative sequence stator voltage. In this figure,
negative sequence stator voltages follow negative sequence grid voltages with the
negative sequence controller in a negative sequence synchronous reference frame.
The transient responses are not good enough because many s terms are neglected in
the design of the negative sequence controller.
Figure 5.5 shows the calculated and measured values of grid and stator line
voltages between phase A and phase B and between phase B and phase C. From
this figure, it can be seen that, the values of instantaneous stator line voltage track

Fig. 5.4 Calculated and


measured [139] d-q negative
Neg-sequence voltage (v)

sequence grid and stator


voltages during
supersynchronous speed

Time (s)
5.6 Results and Discussions 71

Fig. 5.5 Calculated and


measured [139] waveforms of
the grid and stator line

Line voltage a-b (v)


voltages during
supersynchronous speed

Line voltage b-c (v)

Time (s)

the values of grid voltage rapidly with the same amplitude, frequency, and phase
after few periods.
Figure 5.6 shows the calculated and measured values of grid and stator voltage
vectors at stationary reference frame during steady state. This figure shows that the
stator voltage vector track the grid voltage vector. Some oscillations appear in the
measured values of stator voltage vector due to actual PWM inverter.
In all above results, Consideration of iron loss in machine model improves the
theoretical prediction results.
Secondly, magnitudes of A, B, and C phase voltages are kept at rated values
while initial phase angles of A, B, and C phase voltages are set at 0°, −120°, and
−223°, respectively. The grid voltage is applied at t = 0.02 s, and the rotor speed is
1400 rpm (subsynchronous).
Figure 5.7 shows the calculated values of q-axis positive sequence grid voltage
and d-q axis positive sequence stator voltage. This figure shows that, positive
sequence stator voltages follow positive sequence grid voltages with the adaptive
SMC in a positive sequence synchronous reference frame.
Figure 5.8 shows the calculated values of q-axis negative sequence grid voltage
and d-q axis negative sequence stator voltage component. This figure shows that,
negative sequence stator voltages follow negative sequence grid voltages with the
negative sequence controller in a negative sequence synchronous reference frame.
The transient responses are not good enough because many s terms are neglected to
simplify the control algorithm.
72 5 Adaptive Sliding Mode Control for Grid Synchronization …

–axis voltage (v)

Stationary α - axis voltage (v)

Fig. 5.6 Calculated and measured [139] grid and stator voltage vectors at steady state during
supersynchronous speed
Pos-sequence voltage (v)

400

300
Vqg+
200
Vds+
100 Vqs+

0 0.05 0.1 0.15 0.2 0.25 0.3

Time (s)

Fig. 5.7 Calculated d-q positive sequence grid and stator voltages during subsynchronous speed
Neg-sequence voltage (v)

40

20
Vqg- Vds- Vqs-

-20
0 0.05 0.1 0.15 0.2 0.25 0.3
Time (s)

Fig. 5.8 Calculated d-q negative sequence grid and stator voltages during subsynchronous speed
5.6 Results and Discussions 73

Figure 5.9 shows the calculated values of grid and stator line voltages between
phase A and phase B and between phase B and phase C. In this figure, the values of
instantaneous stator line voltage track the values of grid voltage rapidly with the
same amplitude, frequency, and phase after few periods. Also, good tracking
between the stator voltage vector and the grid voltage vector is shown in Fig. 5.10.
These results (Figs. 5.4, 5.5, 5.6, 5.7, 5.8, 5.9, 5.10, and 5.11) confirm that the
improved control scheme can effectively control stator voltages of the DFIG to
accurately follow unbalanced grid voltages at supersynchronous and subsyn-
chronous speed.
For comparison the performances of the conventional cascaded proportional
integral (PI) control scheme are investigated. In this scheme, the grid and stator
voltages are not decomposed into positive and negative sequence components and
the controller is designed in the synchronous reference frame. Magnitudes of A and
B phase voltages are kept at the rated value while magnitude of C phase voltage is
reduced to provide a three-phase unbalanced grid voltage. The grid voltage is
applied at t = 0.02 s, and the rotor speed is 1600 rpm (supersynchronous).
Figure 5.11 shows the calculated and measured values (obtained from reference
[139]) of q-axis grid voltage and d-q axis stator voltage in synchronous reference
frame with conventional PI control scheme. The corresponding calculated and
measured values of grid and stator line voltages between phase A and phase B and
between phase B and phase C are shown in Fig. 5.12.
Figure 5.11 shows that dc components are formed by the vectors of positive
sequence voltage components, while 100 Hz ripples are formed by the vectors of
negative sequence voltage components. Figure 5.12 shows that, small amplitude of
stator line voltage between phase A and phase B whereas large amplitude between
phase B and phase C. In addition to, the values of calculated results agree satis-
factory with those results measured experimentally taken from reference [139].

Vabg Vabs
Line voltage a-b (v)

500

-500
0 0.05 0.1 0.15 0.2 0.25 0.3

Vbcs Vbcg
Line voltage b-c (v)

500

-500
0 0.05 0.1 0.15 0.2 0.25 0.3

Time (s)

Fig. 5.9 Calculated waveforms of the grid and stator line voltage during supersynchronous speed
74 5 Adaptive Sliding Mode Control for Grid Synchronization …

Fig. 5.10 Calculated grid 400


and stator voltage vectors at
steady during subsynchronous
300
speed

–axis voltage (v)


200

100
Grid voltage
0 Stator voltage

-100

-200

-300

-400
-400 -200 0 200 400
Stationary α - axis voltage (v)
Direct and quadrature voltage (v)

Fig. 5.11 Calculated and


measured [139] d-q grid and
stator voltages with PI control
scheme during unbalanced
grid voltages

Time (s)

Figure 5.13 shows the calculated and measured values of grid and stator voltage
vectors at stationary reference frame during steady state. From these figures, it is
obvious that the stator voltage vector shift from the grid voltage vector.
Figures 5.11, 5.12, and 5.13 show that a conventional PI control scheme cannot
control stator voltage to accurately follow unbalanced grid voltages.
The robustness of the improved control scheme against grid frequency is
investigated. Magnitudes of A and B phase voltages are kept at the rated value
while magnitude of C phase voltage is reduced to provide a three-phase unbalanced
grid voltage. The grid voltage is applied at t = 0.02 s, and the grid frequency is
48 Hz.
Figure 5.14 shows the calculated values of q-axis positive sequence grid voltage
and d-q axis positive sequence stator voltage. While the calculated values of q-axis
5.6 Results and Discussions 75

Fig. 5.12 Calculated and


measured [139] waveforms of

Line voltage a-b (v)


the grid and stator line
voltages with PI control
scheme during unbalanced
grid voltage

Line voltage b-c (v)

Time (s)

Fig. 5.13 Calculated and


measured [139] values of grid
and stator voltage vector at
steady state under unbalanced
–axis voltage (v)

grid voltage with PI control


scheme

Stationary α - axis voltage (s)

negative sequence grid voltage and d-q axis negative sequence stator voltage are
shown in Fig. 5.15.
Figure 5.16 shows the calculated values of grid and stator line voltages between
phase A and phase B and between phase B and phase C. The identical calculated
type voltages when the grid frequency is 52 Hz are shown in Figs. 5.17, 5.18 and
5.19, respectively.
It can be seen that, the responses shown in Fig. 5.14, 5.15, 5.16, 5.17, 5.18, and
5.19 are almost the same as the responses shown in Figs. 5.3, 5.4, and 5.5. Since the
76 5 Adaptive Sliding Mode Control for Grid Synchronization …

Pos-sequence voltage (v)


400

300
Vqg+
200
Vds+
100 Vqs +

0 0.05 0.1 0.15 0.2 0.25 0.3


Time (s)

Fig. 5.14 Calculated d-q positive sequence grid and stator voltages during when the grid
frequency is 48 Hz
Neg-sequence voltage (v)

60

40

20

-20 Vqg- Vds- Vqs-


-40
0 0.05 0.1 0.15 0.2 0.25 0.3
Time (s)

Fig. 5.15 Calculated d-q negative sequence grid and stator voltages when the grid frequency is
48 Hz

Vabg Vabs
Line voltage a-b (v)

500

-500
0 0.05 0.1 0.15 0.2 0.25 0.3

Vbcg Vbcs
Line voltage b-c (v)

500

-500
0 0.05 0.1 0.15 0.2 0.25 0.3
Time (s)

Fig. 5.16 Calculated waveforms of the grid and stator line voltages when the grid frequency is
48 Hz
5.6 Results and Discussions 77

48 Hz.

Pos-sequence voltage (v)


400

300
Vqg+
200
Vds+
100
Vqs+
0

0 0.05 0.1 0.15 0.2 0.25 0.3


Time (s)

Fig. 5.17 Calculated d-q positive sequence grid and stator voltages during when the grid
frequency is 52 Hz
Neg-sequence voltage (v)

40

20

0
Vqg- Vds- Vqs-
-20
0 0.05 0.1 0.15 0.2 0.25 0.3

Time (s)

Fig. 5.18 Calculated d-q negative sequence grid and stator voltages when the grid frequency is
52 Hz

Vabg Vabs
Line voltage a-b (v)

500

-500
0 0.05 0.1 0.15 0.2 0.25 0.3

Vbcg Vbcs
Line voltage b-c (v)

500

-500
0 0.05 0.1 0.15 0.2 0.25 0.3
Time (s)

Fig. 5.19 Calculated waveforms of the grid and stator line voltages when the grid frequency is
52 Hz
78 5 Adaptive Sliding Mode Control for Grid Synchronization …

synchronous speed xe used in the control algorithm is calculated from grid voltage
rather than fixed, the performance of a control scheme is not influenced by grid
frequency error.
The performances of the proposed synchronization scheme against parametric
uncertainty are investigated. Magnitudes of A and B phase voltages are kept at the
rated value while magnitude of C phase voltage is reduced to provide a three-phase
unbalanced grid voltage. The grid voltage is applied at t = 0.02 s.
Figure 5.20. shows the calculated values of d-q axis positive and negative
sequence components of grid and stator voltages with the rotor winding resistance
of the DFIG is changed to 50 and 150% of its nominal value. While Fig. 5.21
shows the corresponding calculated values of grid and stator line voltages between
phase A and phase B and between phase B and phase C.
From Fig. 5.20, it can be seen that positive sequence stator voltages follow
positive sequence grid voltages with adaptive SMC and negative sequence stator
voltages follow negative sequence grid voltages with negative sequence controller.
Also, the transient amplitudes of negative sequence stator voltages increase when
the rotor winding resistance decreases. However, amplitudes of these transient
oscillations are very small with respect to the amplitude of positive sequence
components. Hence, the influence in stator voltage is hard to be observed as shown
in Fig. 5.21. The identical calculated type voltages with the rotor self-inductance of
the DFIG is changed to 50 and 150% of its nominal value are shown in Figs. 5.22
and 5.23, respectively.

300
Pos-sequence voltage (v)

250
Vqg+
200 Vds+ (1.5 R )
ro
150 Vqs+ (1.5 R )
ro
100 Vds+ (0.5 R )
ro
50 Vqs+ (0.5 R ro
)

0
-50
0 0.05 0.1 0.15 0.2 0.25 0.3
Neg-sequence voltage (v)

Time(s)

Fig. 5.20 Calculated d-q positive and negative sequence grid and stator voltages with rotor
resistance variation
5.6 Results and Discussions 79

Vabg Vabs (1.5 Rro) Vabs (0.5 Rro)

Line voltage a-b (v) 500

-500
0 0.05 0.1 0.15 0.2 0.25 0.3

Vbcs (1.5 Rro) Vbcg Vbcs (0.5 Rro)


Line voltage b-c (v)

500

-500
0 0.05 0.1 0.15 0.2 0.25 0.3
Time (s)

Fig. 5.21 Calculated waveforms of the grid and stator line voltages with rotor resistance variation
Pos-sequence voltage (v)

300
250 Vqg+
200 Vds+ (1.5 L )
ro
150 Vqs+ (1.5 L )
ro
100 Vds+ (0.5 L )
ro
50 Vqs+ (0.5 L )
ro
0
-50
0 0.05 0.1 0.15 0.2 0.25 0.3
Neg-sequence voltage (v)

Time(s)

Fig. 5.22 Calculated d-q positive and negative sequence grid and stator voltages with rotor
self-inductance variation
80 5 Adaptive Sliding Mode Control for Grid Synchronization …

Vabg Vabs (1.5 Lro) Vabs (0.5 Lro)

Line voltage a-b (v) 500

-500
0 0.05 0.1 0.15 0.2 0.25 0.3

Vbcs (1.5 Lro) Vbcg Vbcs (0.5 Lro)


Line voltage b-c (v)

500

-500
0 0.05 0.1 0.15 0.2 0.25 0.3
Time (s)

Fig. 5.23 Calculated waveforms of grid and stator line voltages with rotor self-inductance
variation

Figures 5.20, 5.21, 5.22, and 5.23 confirm the parametric robustness of the
proposed synchronization scheme.
Chapter 6
Conclusions and Suggestions for Future
Work

6.1 Conclusions

In view of the analysis and investigations presented, one can draw the following
main conclusions:
1. The phase angle between grid voltage and current waveforms is changed from
leading to lagging within one cycle for step change in reactive component of
grid current reference igq . This confirms that, the capability of the GSC control
to regulate the grid received reactive power for controlling the system power
factor.
2. The actual values of d-q-axis rotor current components change and match their
reference values without the impact of any values on the other. This confirms
that the effectiveness of the proposed RSC control scheme for independent
control of active and reactive power of the wind driven DFIG.
3. When the Dc-link power is stepped from zero to its rated value and vice versa,
the GSC control has the capability to maintain the Dc-link voltage constant
around its reference value. The calculated values of Dc-link voltage close to
those measured experimentally with dip and overshoot following step change
and removal of the q-axis rotor current reference.
4. The estimated values of speed of the wind driven DFIG during subsynchronous
and supersynchronous speed operation, agree satisfactorily with those mea-
sured experimentally in the literature, this confirm the effectiveness of a
modified MRAS observer.
5. Satisfactory agreement between the calculated and measured values, in the
previously published literature by others, of d-q-axis rotor current and the
estimated rotor speed confirms that, consideration of core loss and main flux
saturation in machine modeling improve the theoretical prediction and reduces
the error between calculated and measured values.
6. The estimated values of the DFIG speed based on a modified MRAS are
approximately in agreement with those measured experimentally, in the
© Springer International Publishing AG 2018 81
A. Abdelbaset et al., Wind Driven Doubly Fed Induction Generator,
Power Systems, https://doi.org/10.1007/978-3-319-70108-0_6
82 6 Conclusions and Suggestions for Future Work

previously published literature by others, at low wind speed with the machine
operating on the optimal power curve. These values have slight deviation from
their measured one due to the mechanical loss is disregarded in the machine
model.
7. The proposed adaptive SMC schemes are an effective algorithm to synchronize
the DFIG-based wind energy system to the power grid, under balanced and
unbalanced grid voltage, rapidly at both subsynchronous and supersynchronous
speed.
8. Satisfactory agreement between the calculated and measured values of stator
and grid voltage confirms that, consideration of core loss in machine modeling
improves the theoretical prediction and reduces the error between calculated
and measured values.
9. Comparison of the performances between conventional PI control scheme and
proposed control scheme shows that proposed control scheme has better
robustness.
10. The proposed adaptive SMC algorithm showed fast and smooth synchroniza-
tion even if parametric errors are presented, this indicate the robustness of the
proposed scheme.

6.2 Suggestions for Future Work

1. In this book, the effect of core losses and main flux saturation was included in
the DFIG model only, but disregarded in the controller. Therefore, compensa-
tion of core losses and main flux saturation in the vector control of the wind
driven DFIG calls for future studies.
2. The artificial intelligence (AI) techniques such as fuzzy logic, neural network,
and genetic can be used for future work to control and synchronize the DFIG to
the grid.
3. For further complete real-time analysis, the developed system can be validated
through laboratory validation.
Appendix A: Parameters of DFIG Wind
Turbine System and Laboratory
Measurement of Magnetizing Inductance

The non-linear relationship between the air-gap voltage and the magnetizing current
is measured from no load test of the induction machine. Then, the relationship
between the magnetic flux and the magnetizing current (i.e. magnetizing curve) has
been obtained. The data of the magnetizing curve was fitted by a suitable poly-
nomial which is expressed as:

km ðIm Þ ¼ 0:00016 Im4  0:0024 Im3  0:0014 Im2 þ 0:17 Im  0:0034 ðA:1Þ

The magnetizing inductance Lm is calculated from the above polynomial as:

Em ðIm Þ
Lm ¼ ðA:2Þ
x e Im

Figure A.1 shows the relationship between the magnetizing voltage and the
magnetizing current (Table A.1).

© Springer International Publishing AG 2018 83


A. Abdelbaset et al., Wind Driven Doubly Fed Induction Generator,
Power Systems, https://doi.org/10.1007/978-3-319-70108-0
84 Appendix A: Parameters of DFIG Wind Turbine System ...

Table A.1 Parameter and Data of DFIG wind driven system [8]
Rated power of DFIG 7.5 kW
Stator voltage 415 V
Rotor voltage 440 V
Rated stator current 19 A
Rated rotor current 11 A
Stator frequency 50 Hz
Stator resistance Rs 1.06 X
Rotor resistance Rr 0.8 X
Stator self inductance Ls 0.2065 H
Rotor self inductance Lr 0.2141 H
Dc–link capacitor c 2.4 mF
Resistance of grid side inductor Rf 0.1 X
Inductance of grid side inductor Lf 0.012 H
Dc–link voltage reference Vdc 550 V
Core loss resistance RFe 1735 X
Eddy current coefficient Ke 0.095 W ∙ s2
Inertia JT 7.5 Kgm2
Pole pairs np 3
Friction coefficient B 0.06 Nms/rad
Hysteresis coefficient Kh 1.25 W ∙ s

0.8
Magnetizing voltage (v)

0.6

0.4

0.2

0
0 1 2 3 4 5 6
Magnetizing current (A)

Figure A.1 Magnetizing curve of the induction machine used in simulation


Appendix B: Adaptation Mechanisms
and Stability of MRAS

It is important to ensure that the system will be stable and the estimated quantity
will converge to the actual value for the adaptation mechanism of MRAS algo-
rithms. In general xr is a variable; thus the models are linear time-varying systems.
For the purpose of deriving an adaptation mechanism, however, it is valid to
initially treat xr as a constant parameter of the models.
By differentiating both sides of Eq. (3.41), we get:
  " #   
^e  T1r xr ^ema L2m ias
p ma ¼ þ ðB:1Þ
^emb xr  T1r ^emb Lr Tr ibs

Subtracting (B.1) for the adjustable model from the corresponding equation for
the reference model, we obtain the following state error equation:

pe ¼ Ae  W ðB:2Þ

where

e ¼ ½ ea eb  T ;
" #
 T1r xr 1
A¼ ¼ I þ xr J;
xr  T1r Tr
 
^emb
^ rÞ
W ¼ ðx r  x ^ r Þ  ^em ;
¼ ðxr  x
ema
   
1 0 0 1
I¼ and J¼ :
0 1 1 0

^ r is produced by adaptation mechanism, Eq. (3.40) describes a nonlinear


Since x
feedback system Hyperstability requires that the linear time-invariant forward path

© Springer International Publishing AG 2018 85


A. Abdelbaset et al., Wind Driven Doubly Fed Induction Generator,
Power Systems, https://doi.org/10.1007/978-3-319-70108-0
86 Appendix B: Adaptation Mechanisms and Stability of MRAS

transfer matrix be strictly positive and real, and that the nonlinear feedback
including the adaptation mechanism satisfies Popov’s criterion for hyperstability
[134].

Zt1
e  Wdt   c20 for all t1  0 ðB:3Þ
0

where c20 is a real positive constant, letting:

Zt1    
KI
ðe  ^em Þ  xr  KP þ ð^em  eÞ dt   c20 ðB:4Þ
s
0

Using the following well known inequality:

Zt1
1
ðpF ðtÞÞF ðtÞ dt   F ð0Þ2 ðB:5Þ
2
0

It can be shown that inequality (B.5) is satisfied.


Appendix C: Parameters of DFIG
as Listed in Reference [135]

See Table C.1.

Table C.1 Parameter of DFIG [135]


Stator voltage 380 V
Stator frequency 50 Hz
Synchronous speed 1500 rpm
Stator resistance Rs 2.6596 X
Stator leakage inductance Lls 0.0186 H
Rotor voltage 120 V
Rotor resistance Rr 5.8985 X
rotor leakage inductance Llr 0.0186 H
Mutual inductance Lm 0.2987 H

© Springer International Publishing AG 2018 87


A. Abdelbaset et al., Wind Driven Doubly Fed Induction Generator,
Power Systems, https://doi.org/10.1007/978-3-319-70108-0
References

1. A.D. Hansen, L.H. Hansen, Market penetration of different wind turbine concepts over the
years. 2007 European Wind Energy Conference and Exhibition, 2007
2. H. Geng, G. Yang, Output power control for variable-speed variable-pitch wind generation
systems. IEEE Trans. Energy Convers. 25, 494–503 (2010)
3. K.W.E. Cheng, J.K. Lin, Y.J. Bao, X. D. Xue, Review of the wind energy generating system,
in 8th International Conference on Advances in Power System Control, Operation and
Management (APSCOM), pp. 1–7, 2009
4. L.M. Fernández, F. Jurado, J.R. Saenz, Aggregated dynamic model for wind farms with
doubly fed induction generator wind turbines. Renew. Energy 33, 129–140 (2008)
5. E. Muljadi, C.P. Butterfield, Pitch-controlled variable-speed wind turbine generation. IEEE
Trans. Ind. Appl. 37, 240–246 (2001)
6. H.M. Jabr, N.C. Kar, Fuzzy gain tuner for vector control of doubly-fed wind driven induction
generator, in Canadian Conference on Electrical and Computer Engineering, pp. 2266–
2269, 2006.
7. R. Spee, S. Bhowmik, J.H.R. Enslin, Novel control strategies for variable-speed doubly fed
wind power generation systems. Renew. Energy 6, 907–915 (1995)
8. R. Pena, J.C. Clare, G.M. Asher, Doubly fed induction generator using back-to-back PWM
converters and its application to variable-speed wind-energy generation. IEE Proc. Electr.
Power Appl. 143, 231–241 (1996)
9. A. Tapia, G. Tapia, J.X. Ostolaza, J.R. Saenz, Modeling and control of a wind turbine driven
doubly fed induction generator. IEEE Trans. Energy Convers. 18, 194–204 (2003)
10. M. Tazil, V. Kumar, R.C. Bansal, S. Kong, Z.Y. Dong, W. Freitas, H.D. Mathur,
Three-phase doubly fed induction generators: an overview. Electr. Power Appl. IET 4, 75–89
(2010)
11. M.G. Simoes, F.A. Farret, Renewable Energy Systems: Design and Analysis with Induction
Generators (CRC Press, 2004).
12. T. Sun, Power quality of grid-connected wind turbines with DFIG and their interaction with
the grid. Ph.D. dissertation, Aalborg University, Denmark, 2004.
13. I. Munteanu, A.I. Bratcu, N.-A. Cutululis, E. Ceanga, Optimal Control of Wind Energy
Systems: Towards a Global Approach (Springer, 2008).
14. M. Yin, G. Li, M. Zhou, C. Zhao, Modeling of the wind turbine with a permanent magnet
synchronous generator for integration, IEEE Power Engineering Society General Meeting,
pp. 1–6, 2007
15. J.G. Slootweg, S.W.H. De Haan, H. Polinder, W.L. Kling, General model for representing
variable speed wind turbines in power system dynamics simulations. IEEE Trans. Energy
Convers. 18, 144–151 (2003)
16. A. Perdana, Dynamic models of wind turbines. Ph.D. dissertation, Chalmers University of
Technology, Sweden, 2008

© Springer International Publishing AG 2018 89


A. Abdelbaset et al., Wind Driven Doubly Fed Induction Generator,
Power Systems, https://doi.org/10.1007/978-3-319-70108-0
90 References

17. N.W. Miller, W.W. Price, J.J. Sanchez-Gasca, Dynamic modeling of GE 1.5 and 3.6 wind
turbine-generators. GE-Power systems energy consulting, 2003.
18. M. Arifujjaman, Modeling, simulation and control of grid connected Permanent Magnet
Generator (PMG)-based small wind energy conversion system, in IEEE Electric Power and
Energy Conference (EPEC), pp. 1–6, 2010.
19. P.M. Anderson, A. Bose, Stability simulation of wind turbine systems. IEEE Trans. Energy
Convers. 3791–3795, 1983.
20. S.M. Muyeen, M.H. Ali, R. Takahashi, T. Murata, J. Tamura, Y. Tomaki, A. Sakahara, E.
Sasano, Comparative study on transient stability analysis of wind turbine generator system
using different drive train models. Renew. Power Gener. IET 1, 131–141 (2007)
21. S.A. Papathanassiou, M.P. Papadopoulos, Mechanical stresses in fixed-speed wind turbines
due to network disturbances. IEEE Trans. Energy Convers. 16, 361–367 (2001)
22. R. MelÃcio, V.M.F. Mendes, J.P.S. Catalao, Harmonic assessment of variable-speed wind
turbines considering a converter control malfunction. Renew. Power Gener. IET 4, 139–152
(2010)
23. H. Li, Z. Chen, Transient stability analysis of wind turbines with induction generators
considering blades and shaft flexibility. 33rd Annual Conference of the IEEE Industrial
Electronics Society, IECON 2007, pp. 1604–1609, 2007
24. G. Ramtharan, N. Jenkins, O. Anaya Lara, E. Bossanyi, Influence of rotor structural
dynamics representations on the electrical transient performance of FSIG and DFIG wind
turbines. Wind Energy 10, 293–301, (2007)
25. B. Boukhezzar, H. Siguerdidjane, Nonlinear control of a variable-speed wind turbine using a
two-mass model. IEEE Trans. Energy Convers. 26, 149–162 (2011)
26. W. Qiao, W. Zhou, J.M. Aller, R.G. Harley, Wind speed estimation based sensorless output
maximization control for a wind turbine driving a DFIG. IEEE Trans. Energy Convers. 23,
1156–1169 (2008)
27. L. Mihet-Popa, F. Blaabjerg, I. Boldea, Wind turbine generator modeling and simulation
where rotational speed is the controlled variable. IEEE Trans. Energy Convers. 40, 3–10
(2004)
28. F. Mei, B.C. Pal, Modal analysis of a grid connected doubly-fed induction generator, in The
3rd IET International Conference on Power Electronics, Machines and Drives, pp. 611–615,
2006
29. T. Petru, Modeling of Wind Turbines for Power System Studies (Chalmers University of
Technology, 2003)
30. G. Quinonez-Varela, A. Cruden, Modelling and validation of a squirrel cage induction
generator wind turbine during connection to the local grid. Gener. Transm. Distrib. IET 2,
301–309 (2008)
31. C. Wang, G. Weiss, Integral input-to-state stability of the drive-train of a wind turbine, in
46th IEEE Conference on Decision and Control, pp. 6100–6105, 2007
32. Y. Lei, A. Mullane, G. Lightbody, R. Yacamini, Modeling of the wind turbine with a doubly
fed induction generator for grid integration studies. IEEE Trans. Energy Convers. 21, 257–
264 (2006)
33. J.B. Ekanayake, L. Holdsworth, X. Wu, N. Jenkins, Dynamic modeling of doubly fed
induction generator wind turbines. IEEE Trans. Energy Convers. 18, 803–809 (2003)
34. L.J. Ontiveros, P.E. Mercado, G.O. Suvire, A new model of the double-feed induction
generator wind turbine, in 2010 IEEE/PES Transmission and Distribution Conference and
Exposition: Latin America (T&D-LA), pp. 263–269, 2010
35. A. Petersson, Analysis, modeling and control of doubly-fed induction generators for wind
turbines. Ph.D. dissertation, Chalmers University of Technology, Sweden, 2005.
36. K. Protsenko, D. Xu, Modeling and control of brushless doubly-fed induction generators in
wind energy applications. IEEE Trans. Energy Convers. 23, 1191–1197 (2008)
References 91

37. Y. Zhang, Z. Li, J. Hu, W. Xu, J. Zhu, A cascaded brushless doubly fed induction generator
for wind energy applications based on direct power control, in International Conference on
Electrical Machines and Systems (ICEMS), pp. 1–6, August 2011.
38. I. Erlich, J. Kretschmann, J. Fortmann, S. Mueller-Engelhardt, H. Wrede, Modeling of wind
turbines based on doubly-fed induction generators for power system stability studies. IEEE
Trans Energy Convers. 22, 909–919 (2007)
39. I. Erlich, F. Shewarega, Modeling of wind turbines equipped with doubly-fed induction
machines for power system stability studies, in IEEE PES Power Systems Conference and
Exposition, 2006. PSCE’06, pp. 978–985, 2006
40. A. Luna, F.D.A. Lima, D. Santos, P. Rodríguez, E.H. Watanabe, S. Arnaltes, Simplified
modeling of a DFIG for transient studies in wind power applications. IEEE Trans. Energy
Convers. 58, 9–20 (2011)
41. K. Elkington, M. Ghandhari, Comparison of reduced order doubly fed induction generator
models for nonlinear analysis, in IEEE, Electrical Power & Energy Conference (EPEC),
pp. 1–6, 2009
42. P. Sørensen, A.D. Hansen, T. Lund, H. Bindner, Reduced models of doubly fed induction
generator system for wind turbine simulations. Wind Energy 9, 299–311 (2006)
43. A.S. Neto, S.L.A. Ferreira, J.P. Arruda, F.A.S. Neves, P.A.C. Rosas, M.C. Cavalcanti,
Reduced order model for grid connected wind turbines with doubly fed induction generators.
IEEE, International Symposium on Industrial Electronics, pp. 2655–2660, 2007
44. A. Abbaszadeh, S. Lesan, V. Mortezapour, Transient response of doubly fed induction
generator under voltage sag using an accurate model. IEEE PES/IAS Conference on
Sustainable Alternative Energy (SAE), pp. 1–6, 2009
45. P. Ledesma, J. Usaola, Effect of neglecting stator transients in doubly fed induction
generators models. IEEE Trans Energy Convers. 19, 459–461 (2004)
46. T. Thiringer, J. Luomi, Comparison of reduced-order dynamic models of induction
machines. IEEE Trans Energy Convers. 16, 119–126 (2001)
47. G.G. Richards, O.T. Tan, Simplified models for induction machine transients under balanced
and unbalanced conditions. IEEE Trans. Energy Convers. 15–21 (1981)
48. J.-H. Choi, H.-C. Kim, J.-S. Kwak, Indirect current control scheme in PWM voltage-sourced
converter, in Proceedings of the Power Conversion Conference-Nagaoka, vol. 1, pp. 277–
282, 1997
49. V. Blasko, V. Kaura, A new mathematical model and control of a three-phase AC-DC
voltage source converter. IEEE Trans. Energy Convers. 12, 116–123 (1997)
50. B.-H. Kwon, J.-H. Youm, J.-W. Lim, A line-voltage-sensorless synchronous rectifier. IEEE
Trans. Power Electron. 14, 966–972 (1999)
51. Y. Guo, X. Wang, H.C. Lee, B.-T. Ooi, Pole-placement control of voltage-regulated PWM
rectifiers through real-time multiprocessing. IEEE Trans. Ind. Electron. 41, 224–230 (1994)
52. J.R. Rodríguez, J.W. Dixon, J.R. Espinoza, J. Pontt, Pablo, PWM regenerative rectifiers:
state of the art. IEEE Trans. on Energy Convers. 52, 5–22 (2005)
53. N. Horiuchi, T. Kawahito, Torque and power limitations of variable speed wind turbines
using pitch control and generator power control. Power Engineering Society Summer
Meeting, 2001, vol. 1, pp. 638–643, 2001
54. M. Mansour, M.N. Mansouri, M.F. Mimouni, Study of performance of a variable-speed wind
turbine with pitch control based on a permanent magnet synchronous generator, in 2011 8th
International Multi-Conference on Systems, Signals and Devices (SSD), pp. 1–6, March
2011.
55. M. Faridi, R. Ansari, S.A. Mousavi, M. Dodman, Pitch control of wind turbine blades in
noisy and unstable wind conditions, in 2010 9th International Conference on Environment
and Electrical Engineering (EEEIC), pp. 22–25, May 2010
56. J. Zhang, M. Cheng, Z. Chen, X. Fu, Pitch angle control for variable speed wind turbines, in
Third International Conference on Electric Utility Deregulation and Restructuring and
Power Technologies, 2008. DRPT 2008, pp. 2691–2696, 2008
92 References

57. T. Senjyu, R. Sakamoto, N. Urasaki, T. Funabashi, H. Fujita, H. Sekine, Output power


leveling of wind turbine generator for all operating regions by pitch angle control. IEEE
Trans Energy Convers. 21, 467–475 (2006)
58. E.B. Muhando, T. Senjyu, A. Yona, H. Kinjo, T. Funabashi, Disturbance rejection by dual
pitch control and self-tuning regulator for wind turbine generator parametric uncertainty
compensation. Control Theory Appl. IET 1, 1431–1440 (2007)
59. N.A. Schinas, N.A. Vovos, G.B. Giannakopoulos, An autonomous system supplied only by
a pitch-controlled variable-speed wind turbine. IEEE Trans. Energy Convers. 22, 325–331
(2007)
60. I. Hamzaoui, F. Bouchafaa, A. Hadjammar, A. Talha, Improvement of the performances
MPPT system of wind generation, in 2011 Saudi International Electronics, Communications
and Photonics Conference (SIECPC), pp. 1–6, April 2011
61. S. Li, T.A. Haskew, E. Muljadi, Integrative characteristic evaluation of DFIG maximum
power extraction using lookup table approach. 2010 IEEE Power and Energy Society
General Meeting, pp. 1–8, July 2010.
62. H. Li, Z. Chen, J. K. Pedersen, Optimal power control strategy of maximizing wind energy
tracking and conversion for VSCF doubly fed induction generator system, in CES/IEEE 5th
International Power Electronics and Motion Control Conference, vol. 3, pp. 1–6, 2006
63. M. Hilal, M. Maaroufi, M. Ouassaid, Doubly fed induction generator wind turbine control for
a maximum power extraction, in 2011 International Conference on Multimedia Computing
and Systems (ICMCS), pp. 1–7, April 2011
64. Y. Zhao, X.D. Zou, Y.N. Xu, Y. Kang, J. Chen, Maximal power point tracking under
speed-mode control for wind energy generation system with doubly fed introduction
generator, in CES/IEEE 5th International Power Electronics and Motion Control
Conference, vol. 1, pp. 1–5, 2006
65. X. Zhen, Z. Xing, Y. Shuying, L. Qin, Z. Wenfeng, Study on control strategy of maximum
power capture for DFIG in wind turbine system, in 2010 2nd IEEE International Symposium
on Power Electronics for Distributed Generation Systems (PEDG), pp. 110–115, June 2010.
66. Y. Xiao, P. Jia, VSCF wind turbine control strategy for maximum power generation, in 2010
8th World Congress on Intelligent Control and Automation (WCICA), pp. 4781–4786, July
2010
67. R. Datta, V.T. Ranganathan, A method of tracking the peak power points for a variable speed
wind energy conversion system. IEEE Trans. Energy Convers. 18, 163–168 (2003)
68. B. Shen, B. Mwinyiwiwa, Y. Zhang, B.-T. Ooi, Sensorless maximum power point tracking
of wind by DFIG using rotor position phase lock loop (PLL). IEEE Trans. Power Electron.
24, 942–951 (2009)
69. I.K. Buehring, L.L. Freris, Control policies for wind-energy conversion systems, in IEE
Proceedings C (Generation, Transmission and Distribution), vol. 128, pp. 253–261, 1981
70. Q. Wang, L. Chang, An intelligent maximum power extraction algorithm for inverter-based
variable speed wind turbine systems. IEEE Trans. Energy Convers. 19, 1242–1249 (2004)
71. M.G. Simoes, B.K. Bose, R.J. Spiegel, Fuzzy logic based intelligent control of a variable
speed cage machine wind generation system. IEEE Trans. Energy Convers. 12, 87–95 (1997)
72. C. Shao, X. Chen, Z. Liang, Application research of maximum wind-energy tracing
controller based adaptive control strategy in WECS, in CES/IEEE 5th International Power
Electronics and Motion Control Conference, vol. 1, pp. 1–5, 2006
73. E. Koutroulis, K. Kalaitzakis, Design of a maximum power tracking system for
wind-energy-conversion applications. IEEE Trans. Energy Convers. 53, 486–494 (2006)
74. S. Chondrogiannis, M. Barnes, Stability of doubly-fed induction generator under stator
voltage orientated vector control. IET Renew. Power Gener. 2, 170–180 (2008)
75. C. Batlle, A. Doria-Cerezo, R. Ortega, A stator voltage oriented pi controller for the
doubly-fed induction machine, in American Control Conference, ACC’07, pp. 5438–5443,
2007
References 93

76. S. Li, R. Challoo, M.J. Nemmers, Comparative study of DFIG power control using
stator-voltage and stator-flux oriented frames. Presented at Power & Energy Society General
Meeting, 2009. PES’09. IEEE, 2009.
77. I. Takahashi, T. Noguchi, A new quick-response and high-efficiency control strategy of an
induction motor. IEEE Trans. Energy Convers. 820–827, 1986
78. I. Takahashi, Y. Ohmori, High-performance direct torque control of an induction motor.
IEEE Trans. Energy Convers. 25, 257–264 (1989)
79. T. Noguchi, H. Tomiki, S. Kondo, I. Takahashi, Direct power control of PWM converter
without power-source voltage sensors. IEEE Trans. Energy Convers. 34, 473–479 (1998)
80. J. Arbi, M.-B. Ghorbal, I. Slama-Belkhodja, L. Charaabi, Direct virtual torque control for
doubly fed induction generator grid connection. IEEE Trans. Energy Convers. 56, 4163–
4173 (2009)
81. D. Casadei, F. Profumo, G. Serra, A. Tani, FOC and DTC: two viable schemes for induction
motors torque control. IEEE Trans. Energy Convers. 17, 779–787 (2002)
82. T.G. Habetler, F. Profumo, M. Pastorelli, L.M. Tolbert, Direct torque control of induction
machines using space vector modulation. IEEE Trans. Energy Convers. 28, 1045–1053
(1992)
83. S. Arnalte, J.C. Burgos, J.L. Rodriguez-Amenedo, Direct torque control of a doubly-fed
induction generator for variable speed wind turbines. Electric Power Compon. Syst. 30, 199–
216 (2002)
84. K.C. Wong, S.L. Ho, K.W.E. Cheng, Direct torque control of a doubly-fed induction
generator with space vector modulation. Electric Power Compon. Syst. 36, 1337–1350
(2008)
85. Z. Liu, O.A. Mohammed, S. Liu, A novel direct torque control of doubly-fed induction
generator used for variable speed wind power generation. IEEE Power Engineering Society
General Meeting, pp. 1–6, 2007
86. A. Alesina, M. Venturini, Analysis and design of optimum-amplitude nine-switch direct
AC-AC converters. IEEE Trans. Energy Convers. 4, 101–112 (1989)
87. E. Tremblay, S. Atayde, A. Chandra, Comparative study of control strategies for the doubly
fed induction generator in wind energy conversion systems: a DSP-based implementation
approach. IEEE Trans. Energy Convers. 2, 288–299 (2011)
88. M. Malinowski, M.P. Kazmierkowski, A.M. Trzynadlowski, A comparative study of control
techniques for PWM rectifiers in AC adjustable speed drives. IEEE Trans. Energy Convers.
18, 1390–1396 (2003)
89. D. Santos-Martin, J.L. Rodriguez-Amenedo, S. Arnalte, Direct power control applied to
doubly fed induction generator under unbalanced grid voltage conditions. IEEE Trans.
Energy Convers. 23, 2328–2336 (2008)
90. G. Abad, M.A. Rodriguez, G. Iwanski, J. Poza, Direct power control of
doubly-fed-induction-generator-based wind turbines under unbalanced grid voltage. IEEE
Trans. Energy Convers. 25, 442–452 (2010)
91. L. Xu, P. Cartwright, Direct active and reactive power control of DFIG for wind energy
generation. IEEE Trans. Energy Convers. 21, 750–758 (2006)
92. M. Soliman, O.P. Malik, D.T. Westwick, Multiple model predictive control for wind turbines
with doubly fed induction generators. IEEE Trans. Energy Convers. 2, 215–225 (2011)
93. D. Zhi, L. Xu, B.W. Williams, Model-based predictive direct power control of doubly fed
induction generators. IEEE Trans. Energy Convers. 25, 341–351 (2010)
94. L. Xu, D. Zhi, B.W. Williams, Predictive current control of doubly fed induction generators.
IEEE Trans. Energy Convers. 56, 4143–4153 (2009)
95. J. Hu, J.G. Zhu, D.G. Dorrell, A comparative study of direct power control of AC/DC
converters for renewable energy generation, in 37th Annual Conference on IEEE Industrial
Electronics Society, pp. 3578–3583, 2011
94 References

96. B.A. Jihene, A. Khedher, M.F. Mimouni, Speed-sensorless DFIG wind drive based on DTC
using sliding mode rotor flux observer. Int. J. Renew. Energy Res. (IJRER) 2, 736–745
(2012)
97. B. Hopfensperger, D.J. Atkinson, R.A. Lakin, Stator-flux-oriented control of a doubly-fed
induction machine with and without position encoder. IEE Proc. Electr. Power Appl. 147,
241–250 (2000)
98. S. Yang, V. Ajjarapu, A speed-adaptive reduced-order observer for sensorless vector control
of doubly fed induction generator-based variable-speed wind turbines. IEEE Trans. Energy
Convers. 25, 891–900 (2010)
99. G.D. Marques, V. Pires, S. Sousa, D.M. Sousa, A DFIG sensorless rotor-position detector
based on a hysteresis controller. IEEE Trans. Energy Convers. 26, 9–17 (2011)
100. R. Pena, R. Cardenas, J. Proboste, G. Asher, J. Clare, Sensorless control of doubly-fed
induction generators using a rotor-current-based MRAS observer. IEEE Trans. Energy
Convers. 55, 330–339 (2008)
101. R. Ghosn, C. Asmar, M. Pietrzak-David, B. De Fornel, A MRAS Luenberger sensorless
speed control of doubly fed induction machine, in Proceedings of European Power
Electronic Conference, 2003
102. M.S. Carmeli, F. Castelli-Dezza, M. Iacchetti, R. Perini, A MRAS observer applied to
sensorless doubly fed induction machine drives, in IEEE International Symposium on
Industrial Electronics (ISIE), pp. 3077–3082, 2010
103. R. Teichmann, S. Bernet, A comparison of three-level converters versus two-level converters
for low-voltage drives, traction, and utility applications. IEEE Trans. Ind. Appl. 41, 855–865
(2005)
104. M. Ikonen, O. Laakkonen, M. Kettunen, Two-level and three-level converter comparison in
wind power application. Department of Electrical Engineering, Lappeenranta University of
Technology, FI-53851, Lappeenranta, Finland 2005, pp. 1–11
105. J.-S. Lai, F.Z. Peng, Multilevel converters-a new breed of power converters. IEEE Trans
Energy Convers. 32, 509–517 (1996)
106. A. Calle, J. Rocabert, S. Busquets-Monge, J. Bordonau, S. Alepuz, J. Peracaula, Three-level
three-phase neutral-point-clamped back-to-back converter applied to a wind emulator, in
13th European Conference on Power Electronics and Applications, pp. 1–10, 2009
107. M. Malinowski, S. Stynski, W. Kolomyjski, M.P. Kazmierkowski, Control of three-level
PWM converter applied to variable-speed-type turbines. IEEE Trans Energy Convers. 56,
69–77 (2009)
108. G. Abad, M.Á Rodríguez, J. Poza, Three-level NPC converter-based predictive direct power
control of the doubly fed induction machine at low constant switching frequency. IEEE
Trans. Energy Convers. 55, 4417–4429 (2008)
109. E.J. Bueno, S. Cobreces, F.J. Rodriguez, A. Hernandez, F. Espinosa, Design of a
back-to-back NPC converter interface for wind turbines with squirrel-cage induction
generator. IEEE Trans. Energy Convers. 23, 932–945 (2008)
110. R. Melício, V.M.F. Mendes, J.P.S. Catalão, Two-level and multilevel converters for wind
energy systems: a comparative study, in 13th Power Electronics and Motion Control
Conference, pp. 1682–1687, 2008
111. F. Blaabjerg, M. Liserre, K. Ma, Power electronics converters for wind turbine systems.
IEEE Trans. Energy Convers. 48, 708–719 (2012)
112. O.S. Senturk, L. Helle, S. Munk-Nielsen, P. Rodriguez, R. Teodorescu, Medium voltage
three-level converters for the grid connection of a multi-MW wind turbine, in 13th European
Conference on Power Electronics and Applications, pp. 1–8, 2009
113. E. Pouresmaeil, D. Montesinos-Miracle, O. Gomis-Bellmunt, Control scheme of three-level
H-bridge converter for interfacing between renewable energy resources and AC grid, in
Proceedings of the 2011-14th European Conference on Power Electronics and Applications,
pp. 1–9, 2011
References 95

114. P. Samuel, N. Chandrashekhar, R. Gupta, Wind energy conversion based on seven-level


cascaded H-bridge inverter using LabVIEW FPGA, in 2010 International Conference on
Power, Control and Embedded Systems (ICPCES), pp. 1–6, 2010
115. P.W. Wheeler, J. Rodriguez, J.C. Clare, L. Empringham, A. Weinstein, Matrix converters: a
technology review. IEEE Trans. Energy Convers. 49, 276–288 (2002)
116. R. Cárdenas, R. Peña, P. Wheeler, J. Clare, G. Asher, Control of the reactive power supplied
by a WECS based on an induction generator fed by a matrix converter. IEEE Trans. Energy
Convers. 56, 429–438 (2009)
117. R. Cárdenas, R. Peña, G. Tobar, J. Clare, P. Wheeler, G. Asher, Stability analysis of a wind
energy conversion system based on a doubly fed induction generator fed by a matrix
converter. IEEE Trans. Energy Convers. 56, pp. 4194–4206 (2009)
118. S.M. Barakati, M. Kazerani, J.D. Aplevich, Maximum power tracking control for a wind
turbine system including a matrix converter. IEEE Trans. Energy Convers. 24, 705–713
(2009)
119. M. Pahlevaninezhad, A. Safaee, S. Eren, A. Bakhshai, P. Jain, Adaptive nonlinear maximum
power point tracker for a WECS based on permanent magnet synchronous generator fed by a
matrix converter. IEEE Energy Conversion Congress and Exposition, pp. 2578–2583, 2009
120. A. Yarahmadi, D.A. Khaburi, H. Behnia, Direct Virtual Torque Control of DFIG grid
connection using Indirect Matrix Converter. 3rd Power Electronics and Drive Systems
Technology (PEDSTC), pp. 115–120, 2012
121. J. Park, K. Lee, D. Kim, Control method of a doubly-fed induction generator with automatic
grid synchronization, in IEEE 32nd Annual Conference on Industrial Electronics, pp. 4254–
4259, 2006
122. L. Morel, H. Godfroid, A. Mirzaian, J.M. Kauffmann, Double-fed induction machine:
converter optimisation and field oriented control without position sensor. IEE Proc Electr.
Power Appl. 145, 360–368 (1998)
123. K.C. Wong, S.L. Ho, K.W.E. Cheng, Direct voltage control for grid synchronization of
doubly-fed induction generators. Electr. Power Compon. Syst. 36, 960–976 (2008)
124. S.A. Gomez, J.L.R. Amenedo, Grid synchronisation of doubly fed induction generators
using direct torque control, in IEEE 2002 28th Annual Conference of the Industrial
Electronics Society, vol. 4, pp. 3338–3343, 2002
125. G. Yuan, J. Chai, Y. Li, Vector control and synchronization of doubly fed induction wind
generator system, in The 4th International Power Electronics and Motion Control
Conference, IPEMC 2004, vol. 2, pp. 886–890, 2004
126. L. Yang, Z. Xu, J. Ostergaard, Z.Y. Dong, K.P. Wong, Advanced control strategy of DFIG
wind turbines for power system fault ride through. IEEE Trans. Energy Convers. 27, 713–
722 (2012)
127. S. Muller, M. Deicke, R.W. De Doncker, Doubly fed induction generator systems for wind
turbines. Ind. Appl. Mag. IEEE 8, 26–33 (2002)
128. S. Engelhardt, I. Erlich, C. Feltes, J. Kretschmann, F. Shewarega, Reactive power capability
of wind turbines based on doubly fed induction generators. IEEE Trans. Energy Convers. 26,
364–372 (2011)
129. M. Sokola, E. Levi, A novel induction machine model and its application in the development
of an advanced vector control scheme. Int. J. Electr. Eng. Educ. 37, 233–248 (2000)
130. M. Abdel-Salam, S. Abou-Shadi, Y.S.M, Analysis of induction motors fed from constant
current source taking core-loss and saturation into consideration. Electr. Mach. Power Syst.
27, pp. 581–599 (1999)
131. E. Levi, M. Wang, Online identification of the mutual inductance for vector controlled
induction motor drives. IEEE Trans. on Energy Convers. 18, 299–305 (2003)
132. S. Li, T.A. Haskew, K.A. Williams, R.P. Swatloski, Control of DFIG wind turbine with
direct-current vector control configuration. IEEE Trans. Energy Convers. 3, 1–11 (2012)
133. G. Yang, T.H. Chin, Adaptive-speed identification scheme for a vector-controlled speed
sensorless inverter-induction motor drive. IEEE Trans. Energy Convers. 29, 820–825 (1993)
96 References

134. C. Schauder, Adaptive speed identification for vector control of induction motors without
rotational transducers. IEEE Trans. Energy Convers. 28, 1054–1061 (1992)
135. S.Z. Chen, N.C. Cheung, K.C. Wong, J. Wu, Grid synchronization of doubly-fed induction
generator using integral variable structure control. IEEE Trans. Energy Convers. 24, 875–
883 (2009)
136. Y. Zhang, Z. Li, W. Xu, J. Hu, J.G. Zhu, Grid synchronization of DFIG using model
predictive direct power control, in International Conference on Electrical Machines and
Systems (ICEMS) pp. 1–6, 2011
137. J. Hu, H. Nian, B. Hu, Y. He, Z.Q. Zhu, Direct active and reactive power regulation of DFIG
using sliding-mode control approach. IEEE Trans. Energy Convers. 25, 1028–1039 (2010)
138. M.I. Martinez, A. Susperregui, G. Tapia, H. Camblong, Sliding-mode control for a
DFIG-based wind turbine under unbalanced voltage. Presented at Proceedings of the 18th
IFAC World Congress, 2011
139. S.Z. Chen, N.C. Cheung, Y. Zhang, M. Zhang, X.M. Tang, Improved grid synchronization
control of doubly fed induction generator under unbalanced grid voltage. IEEE Trans.
Energy Convers. 26, 799–810 (2011)
140. H. Nian, Y. Song, P. Zhou, Y. He, Improved direct power control of a wind turbine driven
doubly fed induction generator during transient grid voltage unbalance. IEEE Trans. Energy
Convers. 26, 976–986 (2011)
141. P.A. Pérez, P. Albertos, D.A. Sala, Multivariable Control Systems: an Engineering
Approach (Springer, London, 2004)
142. S.Z. Chen, N.C. Cheung, K.C. Wong, J. Wu, Integral sliding-mode direct torque control of
doubly-fed induction generators under unbalanced grid voltage. IEEE Trans. Energy
Convers. 25, 356–368 (2010)
143. H. Nian, Y. Song, P. Zhou, Y. He, Improved direct power control of a wind turbine driven
doubly fed induction generator during transient grid voltage unbalance. IEEE Trans. Energy
Convers. 26, 976–986 (2011)

You might also like