You are on page 1of 203

ISBN 978-3-7376-0716-2

Vol 9
ISBN 978-3-7376-0716-2

Energy Management and Power System Operation Vol 9


9 783737 607162

9 783737 607162

Photovoltaic Systems in Power System Dynamic Studies


Gustav Lammert

Modelling, Control and Stability Analysis of


This thesis investigates the impact of: i) the low voltage ride-through Modelling, Control and
and dynamic voltage support capability; ii) the active current recov-
ery rate; iii) the local voltage control; and iv) the plant-level voltage Stability Analysis of
Photovoltaic Systems in
control of large-scale photovoltaic systems on short-term volt-
age stability and fault-induced delayed voltage recovery as well
as transient and frequency stability. The power system dynamic
performance is analysed using state-of-the-art methods, such as
Power System Dynamic Studies
phasor mode time-domain simulations and the calculation of the
critical clearing time that determines the stability margin. Moreo-
ver, the recently developed Kullback-Leibler divergence measure is
applied to assess the quality of the voltage recovery. Drawbacks of
this metric are outlined and a novel metric, the so-called voltage re-
covery index, is defined that quantifies the delayed voltage recovery
more systematically. The studies are performed with a generic pho-
tovoltaic system model and typical model parameters are used that

Gustav Lammert
were determined in collaboration with a manufacturer. The stability
analysis is performed in DIgSILENT PowerFactory using: i) a one-load
infinite-bus system; and ii) an IEEE multi-machine voltage stability
test system, namely the Nordic test system. The results show that
with the adequate control of photovoltaic systems, power system
dynamic performance can be significantly improved.
Energy Management and Power System Operation
Vol. 9

Edited by
Prof. Dr.-Ing. Martin Braun
University of Kassel
Gustav Lammert

Modelling, Control and Stability Analysis


of Photovoltaic Systems in Power System Dynamic
Dynamic Studies

kassel
university
press
This work has been accepted by the Faculty of Electrical Engineering / Computer Sciences of
the University of Kassel as a thesis for acquiring the academic degree of Doktor der
Ingenieurwissenschaften (Dr.-Ing.).

Supervisor: Prof. Dr.-Ing. Martin Braun (University of Kassel, Germany)


Co-Supervisor: Prof. Dr. Thierry Van Cutsem, (University of Liège, Belgium)

Prof. Dr.-Ing. Mike Meinhardt (University of Kassel, Germany)


Prof. Dr.-Ing. habil. Christian Rehtanz (TU Dortmund, Germany)

Defense day: 30 April 2019


th

The e-book is licensed under a Creative Commons Attribution-ShareAlike


4.0 International Public License.

Bibliographic information published by Deutsche Nationalbibliothek


The Deutsche Nationalbibliothek lists this publication in the Deutsche Nationalbibliografie;
detailed bibliographic data is available in the Internet at http://dnb.dnb.de.

Zugl.: Kassel, Univ., Diss. 2019


ISBN 978-3-7376-0716-2 (print)
ISBN 978-3-7376-0717-9 (e-book)
DOI: http://dx.medra.org/10.19211/KUP9783737607179
URN: https://nbn-resolving.org/urn:nbn:de:0002-407174

© 2019, kassel university press GmbH, Kassel


www.upress.uni-kassel.de

Printing Shop: Print Management Logistics Solutions, Kassel


Printed in Germany
v

Dedicated to those who have taught me in one way or


another.
vii

Abstract
This thesis investigates the impact of: i) the low voltage ride-through and dynamic
voltage support capability; ii) the active current recovery rate; iii) the local voltage
control; and iv) the plant-level voltage control of large-scale photovoltaic systems
on short-term voltage stability and fault-induced delayed voltage recovery as well
as transient and frequency stability. The power system dynamic performance is
analysed using state-of-the-art methods, such as phasor mode time-domain simula-
tions and the calculation of the critical clearing time that determines the stability
margin. Moreover, the recently developed Kullback-Leibler divergence measure is
applied to assess the quality of the voltage recovery. Drawbacks of this metric
are outlined and a novel metric, the so-called voltage recovery index, is defined that
quantifies the phenomenon of fault-induced delayed voltage recovery more systemat-
ically. The studies are performed with a generic photovoltaic system model, which
was developed by the Western Electricity Coordinating Council. Typical model
parameters are used that were determined in collaboration with a manufacturer.
The implemented model is successfully validated against the Renewable Energy
Model Validation tool that was developed by the Electric Power Research Institute.
Moreover, the implemented photovoltaic model is open-source software and can be
used by academia and industry. The stability analysis is performed in DIgSILENT
PowerFactory using: i) a one-load infinite-bus system in order to show the impact of
the photovoltaic system control modes on the fundamental concepts and principles
of short-term voltage stability; and ii) an IEEE multi-machine voltage stability test
system, namely the Nordic test system, that additionally illustrates interactions
with other power system components, such as synchronous generators. The results
show that without the low voltage ride-through capability, the multi-machine sys-
tem is short-term voltage and transient unstable. Only the low voltage ride-through
and dynamic voltage support capability help to avoid instability. The fastest active
current recovery rate achieves the best voltage recovery. However, this recovery rate
also leads to overfrequencies in the system that could cause disconnection of gener-
ation. Therefore, the rate should be tuned considering both, voltage and frequency
dynamics. In case of local constant voltage control, photovoltaic systems try to
restore their pre-fault voltage by increasing their reactive and reducing their active
currents. However, due to the rather stiff grid behaviour, restoring grid voltage
is impossible and the system collapses owing to the lack of active power produced
by the photovoltaic systems. To overcome this problem, adequate reactive power
limitation is required or the current limit logic needs to be changed. If plant control
is used, plant-level constant voltage and local coordinated reactive power/voltage
control should be applied. Finally, the results show that the voltage recovery in-
dex provides useful information about the delayed voltage recovery and helps to
compare different short-term voltage controls of photovoltaic systems.
ix

Acknowledgements
This dissertation was part of my work as a research assistant at the Department
of Energy Management and Power System Operation at the University of Kassel.
Throughout this time, I met remarkable people, whom I would like to thank in the
following paragraphs.
First and foremost, I would like to thank my supervisor Prof. Dr.-Ing. Martin
Braun, for giving me the opportunity to work under his guidance. He has al-
ways encouraged and trusted me and I am thankful that I have been part of his
research group. I would also like to express my gratitude to my co-supervisor
Prof. Dr. Thierry Van Cutsem, for his professional supervision during my external
research stay in Belgium. He helped me not only to develop valuable skills as a
researcher but also as a person. I wish to thank Prof. Dr.-Ing. Mike Meinhardt
for his continuous support over the least years. His lecture on power electronics
for renewable energy resources was outstanding at the University of Kassel. Many
thanks to Prof. Dr.-Ing. habil. Christian Rehtanz, for devoting his time to be mem-
ber of the examining committee and evaluate this thesis. I owe a special thanks to
Prof. Dr.-Ing. habil. István Erlich, who passed away unexpectedly in May 2018. He
was initially part of the examining committee and his exceptional knowledge and
the critical as well as passionate discussions gave the thesis the final shape.
A big thank you to all of my colleagues from the Department of Energy Man-
agement and Power System Operation at the University of Kassel. In particular,
I would like to thank my office colleagues Dr.-Ing. Dirk Fetzer and Dr.-Ing. Darı́o
Lafferte for the fruitful scientific and non-scientific discussions. Their help and mo-
tivation was fundamental to direct my research towards promising areas and to
avoid dead-ends. Needless to say that I also thank the other team members Kai
Fischbach, Christian Hachmann and Alexander Klingmann who supported me in
many different but always useful ways. Furthermore, I would also like to thank the
former colleagues Priyanka Chaudhari, Dr. Stefan Gehler, Dr. Jan Hegemann, Tina
Paschedag and Dr.-Ing. Robert Schmoll for the insightful discussions on technical
and non-technical topics alike.
I owe many thanks to my colleagues from the Fraunhofer Institute for Energy
Economics and Energy System Technology. Throughout the process of this disser-
tation, I have greatly benefited from the close collaboration with Dr.-Ing. Jan von
Appen, Dr. Thomas Degner, Malte Hof, Markus Kraiczy, Maria Nuschke and Luis
David Pabón Ospina. Not to forget, the former colleagues Dr.-Ing. Dominik Geibel,
Dr.-Ing. Tobias Hennig and Prof. Dr.-Ing. Thomas Stetz that I would like to thank
for their support and advise.
I express my deep gratitude to all members of the CIGRE C4/C6.35/CIRED
joint working group on Modelling and Dynamic Performance of Inverter-Based Gen-
eration in Power System Transmission and Distribution Studies. The participation
x

in this group was essential to direct my investigations towards an international


relevance. In particular, I want to thank Dr. Petros Aristidou, Sergio Martı́nez
Villanueva, Prof. Dr. Herwig Renner and Koji Yamashita. Their encouragement
and expertise improved my research significantly.
Many thanks to several other external colleagues that influenced my work in
a beneficial and enriching way. I should thank Dr. Jens C. Boemer, Dr. Anton
Ishchenko, Dr. Pouyan Pourbeik and Daniel Premm for the numerous technical
discussions and the very helpful and illuminating remarks.
I would like to extend my appreciation to my former colleagues from the Uni-
versity of Liège. I thank Gilles Chaspierre, Dr. Lampros Papangelis, Dr. Hamid
Soleimani Bidgoli and Dr. Tilman Weckesser. The common activities made my ex-
ternal research stay in Belgium really enjoyable. I need to thank you for the support
and the pleasant working environment.
I owe a big thank you to my former students Imran Maqbool, Georges Mbewou
Mboumene, Hossam Nasser and Philipp Schmoll, whose absolute dedication to their
work and patience with my never-ending demands contributed significantly to this
thesis. The supervision of their work was a period of intense learning for me, not
only in the scientific area, but also on a personal level.
I am deeply thankful to all of my dear friends around the world, who are the
proof that friendship does not care about borders or nations. I would like to thank
my lovely flatmates Azada, Gertrud, Janosch, Klara and Sarah, who cheered me
up every time I was feeling down, spending time in our gardening project or play-
ing tabletop football in our flat. I would like to extend my thanks to Federico,
Hanna and Linda for the regular conversations about psychedelic vinyl records, the
ups and downs of a researcher, or simply, the creation of a society based on free-
dom, equal rights and solidarity. Needless to say that I thank my dearest friends
Christoph, Karsten, Kristian and Paul for their humor and loyalty. Their support
and friendship are cherished. I thank Olga and Michelle for the reminder of the
most important thing in life.
The last thanks go to my family. I would like to thank my lovely parents Andrea
and Hans for their encouragement and unconditional support throughout my entire
life. I wish to thank my dear brother Friedrich for his valuable suggestions about
professional and private aspects in life. His continuous support is highly appreciated.
I also want to thank my grand parents, in particular, my grand mother Renate, for
the heart-warming conversations on the phone and the delicious food she prepared
for me at every opportunity. Without my family, this dissertation would not have
been possible.

Thank you all,


xi

Publications

Journal publications

5 G. Lammert, D. Premm, L. D. Pabón Ospina, J. C. Boemer, M. Braun, and


T. Van Cutsem, “Control of Photovoltaic Systems for Enhanced Short-Term
Voltage Stability and Recovery”, IEEE Transactions on Energy Conversion,
vol. 34, no. 1, pp. 243–254, 2019.

4 D. Fetzer, G. Lammert, S. Gehler, J. Hegemann, R. Schmoll, and M. Braun,


“Integration of Voltage Dependent Power Injections of Distributed Generators
Into the Power Flow by Using a Damped Newton Method”, International
Journal of Electrical Power & Energy Systems, vol. 99, pp. 695–705, 2018.

3 K. Yamashita, H. Renner, S. Martı́nez Villanueva, G. Lammert, P. Aristi-


dou, L. Zhu, J. Carvalho Martins, L. D. Pabón Ospina, and T. Van Cutsem,
“Industrial Recommendation of Modeling of Inverter-Based Generators for
Power System Dynamic Studies with Focus on Photovoltaic”, IEEE Power
and Energy Technology Systems Journal, vol. 5, no. 1, pp. 1–10, 2018.

2 G. Lammert, K. Yamashita, L. D. Pabón Ospina, H. Renner, S. Martı́nez Vil-


lanueva, P. Pourbeik, F.-E. Ciausiu, and M. Braun, “Modelling and Dynamic
Performance of Inverter Based Generation in Power System Studies: An Inter-
national Questionnaire Survey”, CIRED – Open Access Proceedings Journal,
vol. 2017, no. 1, pp. 1899–1902, 2017.

1 G. Lammert, K. Yamashita, L. D. Pabón Ospina, H. Renner, S. Martı́nez Vil-


lanueva, P. Pourbeik, F.-E. Ciausiu, and M. Braun, “International Industry
Practice on Modelling and Dynamic Performance of Inverter Based Generation
in Power System Studies”, CIGRE Science & Engineering, vol. 8, pp. 25–37,
2017.
xii

Conference publications

16 D. Fetzer, G. Lammert, A. Ishchenko, L. Tabit, and M. Braun, “A Flexible


MATLAB/Simulink RMS-Framework for Electrical Power Systems Designed
for Research and Education”, in IEEE Innovative Smart Grid Technologies
(ISGT) Conference Europe, Sarajevo, 2018.

15 G. Lammert, A. Klingmann, C. Hachmann, D. Lafferte, H. Becker, T. Pasche-


dag, W. Heckmann, and M. Braun,“Modelling of Active Distribution Net-
works for Power System Restoration Studies”, in 10th IFAC Symposium on
Control of Power and Energy Systems (CPES), Tokyo, 2018.

14 L. D. Pabón Ospina, A. Correa, M. Valov, G. Lammert, and D. Premm, “Im-


pact of Plant Level Voltage Control of Large-Scale Inverter Based Generators
on Long-Term Voltage Stability”, in 20th Power System Computation Confer-
ence (PSCC), Dublin, 2018.

13 C. Hachmann, M. Valov, G. Lammert, W. Heckmann, and M. Braun,“Unter-


stützung des Netzwiederaufbaus durch Ausregelung der dezentralen Erzeu-
gung im Verteilnetz”, in Konferenz Zukünftige Stromnetze für erneuerbare
Energien, Berlin, 2018.

12 G. Lammert, K. Yamashita, H. Renner, S. Martı́nez Villanueva, J. Carvalho


Martins, P. Aristidou, T. Van Cutsem, L. D. Pabón Ospina, M. Braun, and
J. C. Boemer, “Activities of the Joint Working Group CIGRE C4/C6.35/
CIRED: Modelling and Dynamic Performance of Inverter Based Generation
in Power System Transmission and Distribution Studies”, in 1st International
Conference on Large-Scale Grid Integration of Renewable Energy in India,
New Delhi, 2017.

11 D. Lafferte, A. Klingmann, D. Fetzer, G. Lammert, C. Hachmann, T. Pasche-


dag, and M. Braun, “Black Start and Island Operation of Distribution Grids
with Significant Penetration of Renewable Resources”, in 1st International
Conference on Large-Scale Grid Integration of Renewable Energy in India,
New Delhi, 2017.

10 D. Fetzer, G. Lammert, T. Paschedag, D. Lafferte, K. Fischbach, M. Nuhn,


C. Jähner, H. Becker, L. R. Roose, and M. Braun, “Reconnection of Pho-
tovoltaic Systems in Low-Voltage Diesel-Powered Microgrids”, in 1st Inter-
national Conference on Large-Scale Grid Integration of Renewable Energy in
India, New Delhi, 2017.

9 C. Hachmann, G. Lammert, D. Lafferte, and M. Braun, “Power System Resto-


ration and Operation of Island Grids with Frequency Dependent Active Power
Control of Distributed Generation”, in NEIS Conference, Hamburg, 2017.
xiii

8 G. Lammert, J. C. Boemer, D. Premm, O. Glitza, L. D. Pabón Ospina, D. Fet-


zer, and M. Braun, “Impact of Fault Ride-Through and Dynamic Reactive
Power Support of Photovoltaic Systems on Short-Term Voltage Stability”, in
IEEE PowerTech, Manchester, 2017.

7 L. D. Pabón Ospina, A. Correa, and G. Lammert, “Implementation and Val-


idation of the Nordic Test System in DIgSILENT PowerFactory”, in IEEE
PowerTech, Manchester, 2017.

6 G. Lammert, K. Yamashita, L. D. Pabón Ospina, H. Renner, S. Martı́nez Vil-


lanueva, P. Pourbeik, F.-E. Ciausiu, and M. Braun, “Modelling and Dynamic
Performance of Inverter Based Generation in Power System Studies: An In-
ternational Questionnaire Survey”, in 24th International Conference on Elec-
tricity Distribution (CIRED), Glasgow, 2017.

5 S. Martı́nez Villanueva, K. Yamashita, H. Renner, J. Carvalho Martins, P. Aris-


tidou, T. Van Cutsem, G. Lammert, and L. D. Pabón Ospina, “Current Status
of Joint Working Group CIGRE C4/C6.35/CIRED: Modeling and Dynamic
Performance of Inverter Based Generation in Power System Transmission and
Distribution Studies”, in XVII Encuentro Regional Iberoamericano del CIGRE
(ERIAC), Ciudad del Este, 2017.

4 I. Maqbool, G. Lammert, A. Ishchenko, and M. Braun, “Power System Model


Reduction with Grid-Connected Photovoltaic Systems Based on Hankel Norm
Approximation”, in 6th Solar Integration Workshop, Vienna, 2016.

3 D. Fetzer, G. Lammert, K. Fischbach, M. Nuhn, J. Weide, D. Lafferte,


T. Paschedag, and M. Braun, “Modelling of Small-Scale Photovoltaic Sys-
tems with Active and Reactive Power Control for Dynamic Studies”, in 6th
Solar Integration Workshop, Vienna, 2016.

2 G. Lammert, L. D. Pabón Ospina, P. Pourbeik, D. Fetzer, and M. Braun, “Im-


plementation and Validation of WECC Generic Photovoltaic System Models
in DIgSILENT PowerFactory”, in IEEE Power & Energy Society General
Meeting, Boston, 2016.

1 M. Kraiczy, G. Lammert, T. Stetz, S. Gehler, G. Arnold, M. Braun, S. Schmidt,


H. Homeyer, U. Zickler, F. Sommerwerk, and C. Elbs, “Parameterization of
Reactive Power Characteristics for Distributed Generators: Field Experience
and Recommendations”, in International ETG Congress, Bonn, 2015.
xiv

Other publications

4 CIGRE Joint Working Group C4/C6.35/CIRED, “Modelling of Inverter-Based


Generation for Power System Dynamic Studies”, Technical Brochure 727,
2018.

3 Fraunhofer Institute for Energy Economics and Energy System Technology


(IEE), “NETZ:KRAFT: Netzwiederaufbau unter Berücksichtigung zukünftiger
Kraftwerksstrukturen, Project Report, 2018.

2 Fraunhofer Institute for Energy Economics and Energy System Technology


(IEE), “DEA-STABIL: Beitrag der Windenergie und Photovoltaik im Vertei-
lungsnetz zur Stabilität des deutschen Verbundnetzes”, Project Report, 2017.

1 University of Kassel, “SMART GRID MODELS: Modelle und Netzanalysen


zukünftiger Smart Grids”, Project Report, 2017.

Supervised theses

4 P. Schmoll, “Vergleich von Simulationsmodellen einer PV-Anlage in


DIgSILENT PowerFactory und MATLAB/Simulink für dynamische Netz-
studien”, Master thesis at the University of Kassel, 2017.

3 H. Nasser, “Entwicklung generischer Simulationsmodelle für Photovoltaikan-


lagen”, Master thesis at the University of Kassel, 2017.

2 G. Mbewou Mboumene, “Power System Load Modelling for Restoration Stud-


ies with DIgSILENT PowerFactory”, Master thesis at the University of Kassel,
2016.

1 I. Maqbool, “Dynamic Equivalencing of Power Systems Using DIgSILENT


PowerFactory”, Master thesis at the Brandenburg University of Technology
Cottbus-Senftenberg, 2015.
xv

Contents

1 Introduction 1
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Literature review . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.4 Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2 Stability definitions and analysis methods 9


2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Definitions and classification . . . . . . . . . . . . . . . . . . . . . . 10
2.2.1 Short-term voltage stability . . . . . . . . . . . . . . . . . . 11
2.2.2 Transient stability . . . . . . . . . . . . . . . . . . . . . . . 12
2.2.3 Frequency stability . . . . . . . . . . . . . . . . . . . . . . . 13
2.2.4 Fault-induced delayed voltage recovery . . . . . . . . . . . . 13
2.2.5 Distinction between short-term voltage and transient stability 14
2.3 Time-domain simulation . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3.1 Time scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3.2 Power system modelling . . . . . . . . . . . . . . . . . . . . 18
2.3.3 Numerical integration methods . . . . . . . . . . . . . . . . 18
2.3.4 Time step . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.4 Critical clearing time . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.5 Kullback-Leibler divergence measure . . . . . . . . . . . . . . . . . 20
2.5.1 Voltage violation criterion . . . . . . . . . . . . . . . . . . . 21
2.5.2 Probability density function of voltage sample . . . . . . . . 22
2.5.3 Probability density function of reference voltage recovery . . 23
2.5.4 Calculation of divergence measure . . . . . . . . . . . . . . . 24
2.5.5 Limitations of Kullback-Leibler divergence measure . . . . . 25
2.6 Voltage recovery index . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.6.1 Voltage violation criterion . . . . . . . . . . . . . . . . . . . 26
2.6.2 Probability density function of voltage sample . . . . . . . . 26
2.6.3 Weighting function . . . . . . . . . . . . . . . . . . . . . . . 28
2.6.4 Calculation of index . . . . . . . . . . . . . . . . . . . . . . 29
2.7 Conclusions and future work . . . . . . . . . . . . . . . . . . . . . . 30
xvi

3 Modelling and control of photovoltaic systems 31


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
3.2 Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.2.2 Generator/converter model . . . . . . . . . . . . . . . . . . . 34
3.2.3 Electrical control model . . . . . . . . . . . . . . . . . . . . 35
3.2.4 Plant control model . . . . . . . . . . . . . . . . . . . . . . . 37
3.3 Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.3.1 Low voltage ride-through and dynamic voltage support . . . 39
3.3.2 Active current recovery rate . . . . . . . . . . . . . . . . . . 42
3.3.3 Local voltage control . . . . . . . . . . . . . . . . . . . . . . 43
3.3.4 Plant-level voltage control . . . . . . . . . . . . . . . . . . . 44
3.3.5 Combinations of control modes . . . . . . . . . . . . . . . . 45
3.4 Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.4.1 Generator/converter model . . . . . . . . . . . . . . . . . . . 46
3.4.2 Electrical control model . . . . . . . . . . . . . . . . . . . . 46
3.4.3 Plant control model . . . . . . . . . . . . . . . . . . . . . . . 46
3.5 Test system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.5.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.5.2 Operating point . . . . . . . . . . . . . . . . . . . . . . . . . 47
3.5.3 Contingency . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.6 Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.6.1 Low voltage ride-through and dynamic voltage support . . . 49
3.6.2 Active current recovery rate . . . . . . . . . . . . . . . . . . 50
3.6.3 Local voltage control . . . . . . . . . . . . . . . . . . . . . . 50
3.6.4 Plant-level voltage control . . . . . . . . . . . . . . . . . . . 50
3.7 Open source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.8 Conclusions and future work . . . . . . . . . . . . . . . . . . . . . . 52

4 Stability analysis of a one-load infinite-bus system 53


4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
4.2 Test system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
4.2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
4.2.2 Load modelling . . . . . . . . . . . . . . . . . . . . . . . . . 56
4.2.3 Modelling of photovoltaic system . . . . . . . . . . . . . . . 60
4.2.4 Operating point . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.2.5 Contingency . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.3 Case studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.3.1 Low voltage ride-through and dynamic voltage support . . . 63
4.3.2 Active current recovery rate . . . . . . . . . . . . . . . . . . 68
4.3.3 Local voltage control . . . . . . . . . . . . . . . . . . . . . . 72
4.3.4 Plant-level voltage control . . . . . . . . . . . . . . . . . . . 76
4.4 Conclusions and future work . . . . . . . . . . . . . . . . . . . . . . 80
xvii

5 Stability analysis of a multi-machine system 81


5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
5.2 Test system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
5.2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
5.2.2 Load modelling . . . . . . . . . . . . . . . . . . . . . . . . . 85
5.2.3 Substitution of synchronous generators . . . . . . . . . . . . 85
5.2.4 Modelling of photovoltaic systems . . . . . . . . . . . . . . . 86
5.2.5 Operating point . . . . . . . . . . . . . . . . . . . . . . . . . 87
5.2.6 Contingencies . . . . . . . . . . . . . . . . . . . . . . . . . . 88
5.3 Case studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
5.3.1 Without photovoltaic systems . . . . . . . . . . . . . . . . . 89
5.3.2 Low voltage ride-through and dynamic voltage support . . . 90
5.3.3 Active current recovery rate . . . . . . . . . . . . . . . . . . 96
5.3.4 Local voltage control . . . . . . . . . . . . . . . . . . . . . . 101
5.3.5 Plant-level voltage control . . . . . . . . . . . . . . . . . . . 106
5.4 Conclusions and future work . . . . . . . . . . . . . . . . . . . . . . 110

6 Conclusions and future work 111


6.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
6.2 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114

Appendices 117

A Stability definitions and analysis methods 119


A.1 Kullback-Leibler divergence measure . . . . . . . . . . . . . . . . . 119
A.1.1 Voltage violation criterion . . . . . . . . . . . . . . . . . . . 119
A.1.2 Probability density function of voltage sample . . . . . . . . 119
A.2 Voltage recovery index . . . . . . . . . . . . . . . . . . . . . . . . . 120
A.2.1 Weighting function . . . . . . . . . . . . . . . . . . . . . . . 120

B Modelling and control of photovoltaic systems 121


B.1 Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
B.1.1 Electrical control model . . . . . . . . . . . . . . . . . . . . 121
B.2 Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
B.2.1 Generator/converter model . . . . . . . . . . . . . . . . . . . 124
B.2.2 Electrical control model . . . . . . . . . . . . . . . . . . . . 125
B.2.3 Plant control model . . . . . . . . . . . . . . . . . . . . . . . 127
B.3 Test system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
B.3.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
B.4 Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
B.4.1 Low voltage ride-through and dynamic voltage support . . . 130
B.4.2 Active current recovery rate . . . . . . . . . . . . . . . . . . 132
B.4.3 Local voltage control . . . . . . . . . . . . . . . . . . . . . . 135
B.4.4 Plant-level voltage control . . . . . . . . . . . . . . . . . . . 137
xviii

C Stability analysis of a one-load infinite-bus system 141


C.1 Test system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
C.1.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
C.1.2 Load modelling . . . . . . . . . . . . . . . . . . . . . . . . . 142
C.1.3 Modelling of photovoltaic system . . . . . . . . . . . . . . . 143
C.1.4 Operating point . . . . . . . . . . . . . . . . . . . . . . . . . 143
C.2 Case studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144
C.2.1 Active current recovery rate . . . . . . . . . . . . . . . . . . 144
C.2.2 Local voltage control . . . . . . . . . . . . . . . . . . . . . . 145
C.2.3 Plant-level voltage control . . . . . . . . . . . . . . . . . . . 146

D Stability analysis of a multi-machine system 147


D.1 Test system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
D.1.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
D.1.2 Load modelling . . . . . . . . . . . . . . . . . . . . . . . . . 148
D.1.3 Modelling of photovoltaic systems . . . . . . . . . . . . . . . 151
D.1.4 Operating point . . . . . . . . . . . . . . . . . . . . . . . . . 153
D.2 Case studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
D.2.1 Low voltage ride-through and dynamic voltage support . . . 155
D.2.2 Active current recovery rate . . . . . . . . . . . . . . . . . . 158
D.2.3 Local voltage control . . . . . . . . . . . . . . . . . . . . . . 160
D.2.4 Plant-level voltage control . . . . . . . . . . . . . . . . . . . 163

Nomenclature 165

Bibliography 175
1

Introduction
1
1.1 Motivation
Electrical power systems are undergoing fundamental structural and technological
changes. The term “structural changes” means that power systems are transformed
from vertically structured systems, with unidirectional power flows from the trans-
mission to the distribution system, to horizontally structured systems, with bidi-
rectional power flows. This includes also reverse power flows from the distribution
back to the transmission system as well as power flows among voltage levels. The
term “technological changes” means that conventional power plants equipped with
synchronous generators are replaced by generation connected via power electronic
converters. This replacement leads to a transition from a highly centralised elec-
tricity production of a few large power plants to a decentralised production of many
small units based on renewable energy resources with a stochastic nature. These
small units are also referred to as distributed generation and most of them are
inverter based and mainly consist of PhotoVoltaic (PV) and wind generation. In
Germany, for instance, inverter-based generation is connected to all voltage levels,
as shown in Fig. 1.1. It should be noted that PV generation is mostly connected
to the low and medium voltage level, whereas wind generation is predominant in
the medium and high voltage level. However, large-scale PV plants connected to
the high or extra high voltage level, such as the 349 MW PV park in Australia [1],
are also planned. The term “large-scale” refers to large PV or wind parks with a
nominal power of 50 MW and higher [2]. Aside from the change of the generation,
also the composition of the load is changing. In Germany, for instance, air condi-
tioning systems [3] or heat pumps [4] are being increasingly installed. Some of them
include induction motors that are directly coupled to the grid. This, in turn, re-
quires specific power system dynamic studies, such as short-term voltage stability or
Fault-Induced Delayed Voltage Recovery (FIDVR) studies. It should be noted that
there are also air conditioning systems or heat pumps, which are coupled via power
electronics, i.e., frequency converters, to the grid. This thesis focuses on the control
of large-scale PV plants and its impact on the dynamic performance of the power
2 Chapter 1. Introduction

system, specifically induction motor loads. Large-scale PV parks are similar in many
aspects with wind power plants that use full converter wind turbines. Therefore,
the achieved results in this thesis of the impact analysis of PV systems on power
system stability, can be transferred to full converter wind turbines due to the similar
modelling and control approach. Nevertheless, the dynamic characteristics of wind
and PV generation are different from conventional power plants, which may impact
the performance of the power system following disturbances. Moreover, generation
based on renewable energy resources that are mostly coupled via inverters, can al-
ready exceed load consumption in some cases. In Germany, for instance, on May 1,
2018, at 2 p.m., all of the load was covered by renewable energy resources, as shown
in Fig. 1.2. These extreme operating conditions require detailed power system dy-
namic studies that analyse the impact of inverter-based generation, specifically PV
generation, on stability of the overall system. The outcome of these studies are grid
code requirements that describe operational aspects of inverter-based generation,
such as the standard for North America [5], the network code for Europe [6], or, in
particular, the guideline for the high voltage level in Germany [7]. The grid codes
set the minimum technical requirements in terms of electrical performance, planning
and operation of an individual unit or park. However, two recent examples from
Europe illustrate the effects of not carefully designed grid codes. The first issue was
the non-existent Fault Ride-Through (FRT) requirement [8]. Initially, grid code
requirements defined the disconnection of inverter-based generation during voltage
dips, which is referred to as passive behaviour. Wind and PV generation would trip
in the wide area, where the fault occurs. This would lead to a considerable loss of
production after the fault, which could result in a significant frequency drop. After

120
Low voltage
Medium voltage
100 High voltage
Extra high voltage
Active power [GW]

80

60

40

20

0
2012 2013 2014 2015 2016 2017 2018 2019 2020 2021 2022
Year

Figure 1.1: Installed power of inverter-based generation in Germany [9]


1.1. Motivation 3

power system dynamic studies [10] identified the problem, FRT requirements [11]
for inverter-based generation were defined that enhance the dynamic performance
of the system. The second issue was the 50.2 Hz problem [8]. Initially, grid codes
defined the disconnection of wind and PV generation for frequencies higher than
50.2 Hz. At that time, about 30 GW PV generation was installed in Germany at
all voltage levels. The simultaneous disconnection of generation in a situation of
high PV power production would exceed the total reserve power in Europe of 3 GW
and most probably lead to a blackout. This issue was solved by a system wide
retrofit program that changed the protection settings of PV units in Germany [12].
In general, these issues show how important it is to make right assumptions on the
future structure of the electrical power system. On the other hand, some of the
aforementioned issues are still under investigation, such as the FRT requirements
for PV systems [13]. Furthermore, recent incidents, such as the Southern California
event [14] or the South Australian blackout [15], both in 2016, highly motivate
the studies reported in this work. The former event was influenced by missing or
incorrect FRT requirements and too slow active current recovery rates [14]. The
driving forces for the latter event were the inability to ride through multiple faults
and the missing Dynamic Voltage Support (DVS) capability, i.e., no additional
reactive current injection in response to the voltage dip [15]. Both incidents require
further power system dynamic studies that determine the impact of different control
and protection methods of inverter-based generation on power system stability. In
this context, adequate model parameters and validated models are crucial in order
to receive a realistic system behaviour and to derive feasible conclusions [16]. All
of the aforementioned aspects are addressed in detail within this thesis.

100 Biomass
Hydro, pumped storage
90 Wind
80 Photovoltaic
Coal, nuclear, gas
Energy [GWh]

70 Consumption
60
50
40
30
20
10
0
00:00 04:00 08:00 12:00 16:00 20:00 24:00
Time [hh:mm]

Figure 1.2: Power generation and consumption in Germany on May 1, 2018 [17]
4 Chapter 1. Introduction

1.2 Literature review


A comprehensive literature review of research work related to this thesis is given in
Table 1.1. The table is divided according to the main chapters of this thesis, i.e.,
analysis methods, modelling and control of PV systems as well as the stability anal-
ysis. The presented references in Table 1.1 are selected based on several criteria, i.e.,
i) the publications are the most relevant in the research area of this thesis; ii) only
renowned and reputable publishers are considered; and iii) the literature should be
recently published. The first part in Table 1.1, i.e., analysis methods, indicates
that time-domain simulation is the state-of-the-art method to study the dynamic
performance of the system and is used in all of the listed references. The second
method in this part is the Critical Clearing Time (CCT), which is used in about
one third of the cases to determine the stability margin of the power system. On
the other hand, the Kullback-Leibler (KL) divergence measure is rarely considered
for the analysis of the delayed voltage recovery, which constitutes the research gap
in this part. The second part in Table 1.1, i.e., modelling and control of PV sys-
tems, is divided into the modelling approach and the control method of PV systems.
The majority of the studies listed in the table neither use generic nor validated PV
system models. Moreover, the control methods with respect to the active current
recovery rate or the plant-level voltage control are rarely investigated. These miss-
ing aspects form the research gap in the modelling and control part. It should be
noted that the table also includes references related to other technologies, such as
energy storage systems, combined heat and power plants, static var compensators,
or doubly fed induction generators. Nevertheless, these references are selected due
to their relevance to the other parts in this table, i.e., analysis methods and sta-
bility analysis. The third and last part in Table 1.1 presents the stability analysis
and outlines the investigated phenomena as well as the used test systems. It is
shown that in most of the cases, multi-machine systems are considered to perform
the stability analysis. With respect to the investigated phenomena, several studies
analysed the impact of PV systems on transient stability and frequency stability.
Nevertheless, only a few studies exist that show the influence of PV systems on
short-term voltage stability, which is also confirmed by a comprehensive review of
power system stability challenges for PV integration [18]. These aspects represent
the research gap in the stability analysis part. It can be concluded that in all three
parts of Table 1.1, i.e., analysis methods, modelling and control of PV systems as
well as the stability analysis, research gaps can be identified. Therefore, further
investigations are required that are addressed accordingly in this thesis, as shown
in the last row of Table 1.1.
1.2. Literature review 5

Table 1.1: Literature review

Analysis Photovoltaic system Stability analysis


methods Model Control Phenomenon Test system
Time-domain simulation

Multi-machine system
Fault-induced delayed
through and dynamic
Critical clearing time

Local voltage control


divergence measure

Frequency stability

infinite-bus system
Transient stability
Low voltage ride-
Kullback-Leibler

voltage recovery
voltage stability
voltage support

voltage control
Active current
recovery rate

Short-term
Plant-level
Reference

Validated

One-load
Generic

[9]           
[19]       
[20]          
[21]         
[22]  
[23]         
[24]    
[25]    
[26]     
[27]     
[28]  
[29]     
[30]     
[31]    
[32]    
[33]     
[34]      
[35]    
[36]       
[37]    
[38]     
[39]       
[40]      
[41]     
[42]      
[43]        
[44]       
[45]       
[46]        
[47]         
[48]    
Thesis               
6 Chapter 1. Introduction

1.3 Contributions
This section presents the scientific contributions of this thesis. According to the re-
search gaps identified in Section 1.2, the results of this work contribute significantly
to fill these gaps. The contributions are separated according to the main chapters
of this thesis, as presented in the sequel.

Stability definitions and analysis methods: The scientific contributions are


focused on the methods for the analysis of the delayed voltage recovery, i.e., the
phenomenon of FIDVR.
• The FIDVR phenomenon is analysed using the recently developed KL diver-
gence measure. This measure is applied for the first time to analyse short-term
controls of PV systems. Drawbacks of the KL divergence measure are pointed
out and improvements are suggested.
• In order to overcome the drawbacks of the KL divergence measure, a novel
enhanced metric, namely the VRI, is developed to quantify the phenomenon
of FIDVR more systematically. The VRI provides useful information about
the delayed voltage recovery and helps to compare different short-term voltage
controls of PV systems.

Modelling and control of PV systems: The scientific contributions are focused


on the modelling and control approach of a large-scale PV plant connected to the
high or extra high voltage level.
• The generic PV system model, which was developed by the Western Electricity
Coordinating Council (WECC), is implemented in DIgSILENT PowerFactory.
The implementation procedure and the required modifications are detailed.
• The relevant PV system control modes that impact the dynamic performance
of the power system, specifically short-term dynamics, are identified. The
considered control modes are the LVRT and DVS capability, the active current
recovery rate, the local voltage control and the plant-level voltage control.
• Typical model parameters of the PV system are provided. These were de-
termined in collaboration with a manufacturer and represent the average be-
haviour of presently required control modes and settings according to inter-
national guidelines and grid codes.
• The implemented model is successfully validated for the defined control modes
against the Renewable Energy Model Validation (REMV) tool, which was
developed by the Electric Power Research Institute (EPRI).
• To serve the need of academia and industry, the DIgSILENT PowerFactory
model of the WECC generic PV system is open-source software and can be
downloaded free of charge and used for any purpose.
1.3. Contributions 7

Stability analysis: A comprehensive stability analysis is performed. The results


show the improvement of the power system dynamic performance using an adequate
control of PV systems.

• A comprehensive analysis of different power system phenomena is performed.


The main focus is the investigation of short-term voltage stability and the phe-
nomenon of FIDVR. Moreover, the impact of PV system control on transient
stability and frequency stability is studied briefly. This exhaustive stability
analysis allows a detailed evaluation of various PV system control methods.

• For the impact analysis of PV system controls, different test systems are con-
sidered. First, a one-load infinite-bus system is used to present fundamental
concepts and principles of power system stability, specifically short-term volt-
age stability and FIDVR. Second, an IEEE multi-machine test system, i.e.,
the Nordic test system, is considered to study additionally the interactions
with other power system components, such as synchronous generators. The
latter test system is particularly suitable for voltage stability studies. Both
test systems enable a precise evaluation on how PV system control influences
the dynamic performance of the power system.

• A significant variety of PV system control methods that impact the power


system dynamic performance, specifically short-term voltage stability and
FIDVR, are investigated in detail. The considered control modes are the
LVRT and DVS capability, the active current recovery rate, the local voltage
control and the plant-level voltage control of large-scale PV systems. The
stability analysis shows that even the relatively slow plant control can signif-
icantly influence short-term dynamics of the power system.
8 Chapter 1. Introduction

1.4 Structure
This section presents the structure of the thesis. The remainder of this work is
organised as follows.

Chapter 2: This chapter presents the relevant stability definitions and their meth-
ods for the analysis of power system dynamic performance. First, the time-domain
simulation method is introduced. Next, the calculation of the CCT, which deter-
mines the stability margin of the system, is presented. The delayed voltage recovery
is assessed using the KL divergence measure. The major steps for the computation
of this metric as well as the drawbacks are outlined and a novel metric, the so-called
VRI, is introduced that quantifies the phenomenon of FIDVR more systematically.
Finally, the conclusions are drawn and directions for future work are outlined.

Chapter 3: This chapter is dedicated to the modelling and control approach of PV


systems. First, the implementation of the WECC generic PV system model is pre-
sented. Next, relevant control modes and typical model parameters are identified.
Then, the model validation along with the considered test system, is presented.
Finally, the open-source access to the implemented PV system model as well as the
conclusions and directions for future work are given.

Chapter 4: In this chapter, the stability analysis of the one-load infinite-bus system
is presented. First, the test system is introduced, which includes details about
the load and the PV system model, the operating point as well as the considered
contingency. Next, the case studies are presented, specifically the results of the
impact analysis of different PV system control modes on short-term voltage stability
and FIDVR. Finally, based on the achieved results, the conclusions are drawn and
directions for future work are outlined.

Chapter 5: This chapter presents the stability analysis of the multi-machine sys-
tem, namely the Nordic test system. First, an overview of the test system is given,
which includes details about the load model, the substitution of synchronous gen-
erators, the PV system model, the operating point as well as the considered con-
tingencies. Next, the case studies are introduced. In particular, the results of the
impact analysis of different PV system control modes on short-term voltage stabil-
ity and FIDVR as well as transient and frequency stability are illustrated. Finally,
the conclusions and directions for future work are presented.

Chapter 6: In this chapter, the general conclusions of the thesis are summarised
and the main directions for future work are outlined.
9

Stability definitions and analysis methods


2
2.1 Introduction
The content of this chapter is partly based on the following publication:
G. Lammert, D. Premm, L. D. Pabón Ospina, J. C. Boemer, M. Braun,
and T. Van Cutsem, “Control of Photovoltaic Systems for Enhanced
Short-Term Voltage Stability and Recovery”, IEEE Transactions on En-
ergy Conversion, vol. 34, no. 1, pp. 243–254, 2019.
Motivation: The majority of methods to analyse power system dynamic perfor-
mance is based on synchronous generator dominated systems. However, with the
increase of inverter-based generation, some of the state-of-the-art methods might
not be adequate anymore. Therefore, new methods are required.

Contributions: This chapter identifies suitable analysis methods in order to pre-


cisely evaluate the dynamic performance of the power system including large-scale
PV plants. A clear definition of power system stability and its categories is given.
Identifying the causes of instability, leads to appropriate analysis methods. Not only
state-of-the-art methods, such as time-domain simulations, are applied, but also the
recently developed KL divergence measure. The application of this measure as well
as the enhanced VRI that is newly defined, encompass the main contributions of
this chapter.

Structure: The remainder of this chapter is organised as follows. The relevant


power system stability definitions and categories are presented in Section 2.2. Sec-
tion 2.3 presents the time-domain simulation method that is used for the performed
studies. Section 2.4 presents the CCT, its definition and calculation. The major
steps for the computation of the KL divergence measure are presented in Section 2.5.
Section 2.6 duly discusses the novel metric, the so-called VRI, to quantify the phe-
nomenon of FIDVR more systematically. Finally, the conclusions and directions for
future work are given in Section 2.7.
10 Chapter 2. Stability definitions and analysis methods

2.2 Definitions and classification


Power system stability is understood in this thesis based on the definition given in
the well-known report [49] as follows:
Power system stability is the ability of an electric power system, for a
given initial operating condition, to regain a state of operating equilib-
rium after being subjected to a physical disturbance, with most system
variables bounded so that practically the entire system remains intact.
Furthermore, report [49] provides a systematic basis for classifying power system
stability, identifying and defining different categories, and providing a broad picture
of the phenomena. The overview of power system stability and its categories and
subcategories is given in Fig. 2.1. From these categories, three stability criteria are
of particular interest for the research presented in this thesis, namely:
• short-term voltage stability,
• transient stability, and
• frequency stability.
The main focus of this work is the impact analysis of PV systems on short-term
voltage stability, as highlighted in dark grey in Fig. 2.1. The corresponding defi-
nition and analysis methods are presented in Section 2.2.1. However, in any given

Power System
Stability

Rotor Angle Frequency Voltage


Stability Stability Stability

Small-Disturbance Transient Large-Disturbance Small-Disturbance


Angle Stability Stability Voltage Stability Voltage Stability

Short Term Short Term Long Term

Short Term Long Term

Figure 2.1: Classification of power system stability [49]


2.2. Definitions and classification 11

situation, any one form of instability may not occur in its pure form, especially in
highly stressed systems [49]. Therefore, the short-term control of PV systems does
not only influence short-term voltage stability, but also transient and frequency sta-
bility, which is also briefly investigated in this work. Hence, these stability criteria
are also highlighted in light grey in Fig. 2.1 and their definitions and analysis meth-
ods are given in Sections 2.2.2 and 2.2.3, respectively. Another important aspect
that is closely linked to short-term voltage stability is the delayed voltage recovery,
specifically the FIDVR phenomenon. This phenomenon can take place before insta-
bility is reached, as explained in Section 2.2.4. Furthermore, the distinction between
short-term voltage stability and transient stability is not always easy and both may
go hand in hand, which is presented in Section 2.2.5. It should be noted that the
presented power system stability categories and definitions are mainly based on re-
port [49] and represent synchronous generator dominated systems. However, as the
penetration level of inverter-based generation is going to increase, it is essential to
revisit some of the existing definitions. Eventually, new terms and definitions need
to be introduced, where appropriate that will adequately reflect and describe the
existing and future dynamic behaviour of evolving power systems [50].

2.2.1 Short-term voltage stability


Definition: In this thesis, voltage stability is distinguished according to its time
scale, which may vary from a few seconds to tens of minutes. In that sense, voltage
stability may be either a short-term or long-term phenomenon, as shown in Fig. 2.1.
The focus of this work is short-term voltage stability and its general definition is
based on the well-known report [49] as follows:
Short-term voltage stability involves dynamics of fast acting load com-
ponents such as induction motors, electronically controlled loads, and
HVDC converters.
It should be noted that short-term voltage stability can be also understood as tran-
sient voltage stability or induction motor stability [51]. In this work, short-term
voltage stability is understood as a phenomenon related to the stalling of induction
motor loads. For slowly cleared fault conditions, induction motors cannot reaccel-
erate. The mechanical and electromagnetic torque curves intersect, but at fault
clearing the induction motor slip exceeds the unstable equilibrium value [52]. This
is also referred to as lack of attraction towards the stable post-disturbance equilib-
rium of short-term dynamics [52]. As the dynamic behaviour of induction motor
loads determines short-term voltage stability, its adequate modelling is crucial. In
this thesis, the term “voltage collapse” refers to the process by which the sequence
of events accompanying voltage instability leads to a blackout or abnormally low
voltages in a significant part of the power system [49]. It should be noted that
voltage stability can be also distinguished according to the considered type of dis-
turbance [49], i.e., large-disturbance voltage stability and small-disturbance voltage
stability, as also shown in Fig. 2.1.
12 Chapter 2. Stability definitions and analysis methods

Analysis methods: The study period of interest to analyse short-term voltage


stability is in the order of several seconds [49]. As short-term voltage stability is a
non-linear stability problem, the analysis is based on solving Differential Algebraic
Equations (DAEs) of the system that capture these non-linearities. The most com-
mon method is the application of software tools to perform numerical time-domain
simulations. Analytical methods, such as transient P–V curves [53], or data-driven,
model-free approaches, such as Lyapunov exponents [54], could also be applied. An-
other metric that can be used to determine the short-term voltage stability margin
is the calculation of the CCT of the fault. In this work, time-domain simulations
and the calculation of the CCT are utilised to analyse short-term voltage stability,
as detailed in Sections 2.3 and 2.4, respectively. Note that the study period of
short-term voltage stability is similar to the study period of transient stability, and
sometimes both phenomena go hand in hand. More details about the distinction
between both stability criteria are given in Section 2.2.5.

2.2.2 Transient stability


Definition: Transient stability is a special case of rotor angle stability, as shown
in Fig. 2.1. It is determined by the dynamic behaviour of synchronous machines,
specifically synchronous generators. In this thesis, transient stability is understood
based on the definition given in the well-known report [49] as follows:
Transient stability is the ability of the power system to maintain syn-
chronism when subjected to a severe disturbance, such as a short circuit
on a transmission line.
Transient stability depends on both the initial operating state of the system and the
severity of the disturbance [49]. In contrast to short-term voltage stability, faults
near generation are important.

Analysis methods: The time frame of interest in transient stability studies is


usually 3 s to 5 s following the disturbance. It may extend to 20 s for very large
systems with dominant inter-area swings [49]. Hence, the study period of interest
to analyse transient stability is similar to the study period of short-term voltage
stability. Transient stability involves large excursions of generator rotor angles
and is influenced by the non-linear power–angle relationship. Therefore, analysis
requires solution of appropriate system DAEs. Thus, it is similar to the analysis of
short-term voltage stability. Numerical time-domain analysis is often used to study
transient stability. Another well-known method for transient stability analysis is
based on Lyapunov’s function [55, 56]. Furthermore, the calculation of the CCT
of the fault is another approach that can be used to analyse transient stability. In
this work, time-domain simulations and the calculation of the CCT are utilised to
analyse transient stability, as detailed in Sections 2.3 and 2.4, respectively. It should
be noted that the distinction between short-term voltage stability and transient
stability is detailed in Section 2.2.5.
2.2. Definitions and classification 13

2.2.3 Frequency stability


Definition: Frequency stability may be either a short-term or a long-term phe-
nomenon, as shown in Fig. 2.1. In this thesis, frequency stability is understood
based on the definition given in the well-known report [49] as follows:
Frequency stability refers to the ability of a power system to maintain
steady frequency following a severe system upset resulting in a significant
imbalance between generation and load.
Severe system upsets generally result in large excursions of frequency, power flows,
voltage, and other system variables. In particular, short-term frequency stability
with regard to transmission network faults is of interest in this thesis. The time
scale of short-term frequency stability is in the order of several seconds. Note that
the counterpart of frequency stability in the short-term time scale is rotor angle
stability, since in this time scale there is no common frequency [52].

Analysis methods: Frequency excursions may vary from a few seconds to several
minutes [49]. A single-bus equivalent with the representation of aggregated load and
generation can be used for simplified analyses. For detailed frequency studies, a full
network representation is advisable. A common method to analyse frequency sta-
bility is to perform numerical time-domain simulations. In this work, time-domain
analysis of the full network representation is used to study frequency stability, as
detailed in Section 2.3.

2.2.4 Fault-induced delayed voltage recovery


Definition: In this thesis, FIDVR is understood based on the definition given in
report [57] as follows:
Fault-induced delayed voltage recovery describes a depressed voltage for
5 s to 30 s following a network fault caused by the presence of large
amounts of induction motor loads at distribution level.
The phenomenon of FIDVR relates to the dynamic behaviour of induction motor
loads, which tend to decelerate following a large disturbance, resulting in low volt-
ages in a significant portion of the power system. FIDVR can take place before
short-term voltage instability is reached.

Analysis methods: As the phenomenon of FIDVR is closely related to short-


term voltage stability, similar analysis methods can be used. Based on time-domain
simulations, the voltage samples of the recovery period are used to apply, e.g., slope-
based methods [58, 59], integral error-based methods [60, 61], or entropy-based
methods [26]. In this thesis, the entropy-based KL divergence measure [26], and an
enhanced novel metric, the so-called VRI, are used to analyse the phenomenon of
FIDVR, as presented in Sections 2.5 and 2.6, respectively.
14 Chapter 2. Stability definitions and analysis methods

2.2.5 Distinction between short-term voltage and transient


stability
As mentioned in Sections 2.2.1 and 2.2.2, the time scale of interest for short-term
voltage stability analysis is also the time scale of interest for transient stability
analysis. Analysis methods for both phenomena require the solution of appropriate
system DAEs that include the non-linearities. On the one hand, there are cases
where one form of instability predominates. For instance, report [62] points out the
two extreme situations, as depicted in Fig. 2.2. The first condition shows a remote
synchronous generator connected via transmission lines to a large system, and faults
near the generator lead to pure transient instability, as shown in Fig. 2.2 (a). The
second situation shows a load connected via transmission lines to a large system,
and faults near the induction motor lead to pure short-term voltage instability, as
illustrated in Fig. 2.2 (b). On the other hand, the distinction between short-term
voltage instability and transient instability is not always clear, and aspects of both
phenomena may exist [51]. By way of example, if short-term voltage instability is
present at one, or a few, of the network buses, the depressed voltage may cause a
voltage reduction at the neighbouring buses, leading to further short-term voltage
instabilities at these buses. If this voltage sag increases throughout the network,
synchronous generators are also affected. If these affected generators are weakly
connected to the system, they may lose synchronism, or, in other words, become
transient unstable [63]. This is a case where short-term voltage instability causes
transient instability. Finally, it should be noted that the loss of synchronism of
machines, as rotor angles between two groups of machines approach 180◦ , causes
a rapid drop in voltages at intermediate points in the network close to the elec-
trical center [64]. The voltages near the electrical center rapidly oscillate between
high and low values as a result of repeated pole slips between the two groups of
machines [49]. Voltage oscillations are the result, loss of rotor angle stability is the
cause. In contrast, the type of sustained fall of voltage that is related to voltage
instability involves loads and may occur where rotor angle stability is not an is-
sue [49]. Sustained fall of voltages is the result, stalling of induction motors is the
cause.

Large Transmission Synchronous Large Transmission Induction


system line 1 generator system line 1 motor
Transmission Transmission
SG IM
line 2 line 2

(a) Bus 1 Bus 2 (b) Bus 1 Bus 2

Figure 2.2: Simple stability examples showing extreme situations


(a) Pure transient instability
(b) Pure short-term voltage instability
2.3. Time-domain simulation 15

2.3 Time-domain simulation


2.3.1 Time scale
In order to study the electrical power system via time-domain simulation, the time
scale has to be clearly defined [65]. Depending on the considered time scale, the
state and algebraic variables can be defined accordingly. In this work, the time scale
of interest concerns short-term voltage stability and FIDVR analysis, as outlined
in Section 2.2. As transient and frequency stability are interlinked with short-term
voltage dynamics, these phenomena are also considered. Different power system
phenomena are shown in Fig. 2.3. The grey stripe indicates the time frame object
of this work, i.e., from 0.01 s to 20 s or, which is the same, from 0.05 Hz to 100 Hz.
Moreover, the phenomena discussed in detail in this thesis are pointed out with a
solid arrow, whereas the phenomena that a briefly investigated are pointed out with
a dashed arrow. According to the time frame object of this work, the type of power
system dynamic simulation has to be selected. In general, dynamic simulations of
power systems can be classified into two categories, i.e., Electro-Magnetic Transient
(EMT) and Root Mean Square (RMS) simulation. The difference between these
two types of simulations is outlined in the sequel.

EMT simulation RMS simulation

Transient overvoltages

Sub-synchronous resonance
Power system phenomena

Short-term voltage stability

Transient stability
Relevant
phenomena
Fault-induced delayed voltage recovery

Long-term voltage stability

Frequency stability

10−3 10−2 10−1 100 101 102 103


Time scale [s]

Figure 2.3: Time scales of relevant power system phenomena


16 Chapter 2. Stability definitions and analysis methods

RMS simulation: RMS simulations are also called phasor mode or simply pha-
sor simulations. RMS models are mainly used to study power system stability
for large interconnected systems. This includes electromechanical oscillations, i.e.,
small-signal stability, transient stability, voltage stability and frequency stability, as
shown in Fig. 2.1. The power system is modelled by a set of non-linear DAEs [64].
Phasor simulation methods are used if the magnitude and phase of the voltages
and currents are of interest. It is not necessary to solve all the differential equa-
tions resulting from the interaction of resistances, inductances and capacitances.
Hence, the network is simulated with fixed complex impedances instead of differen-
tial equations [66]. RMS simulation tools consider phenomena with a bandwidth of
typically 0.1 Hz to 3 Hz, since the network model’s fidelity diminishes rapidly for
phenomena with frequencies significantly outside of this range. However, the con-
trol loops modelled on individual power plants may cover phenomena up to 10 Hz.
Another commonly used terminology for RMS type models is to call them positive-
sequence stability models. In essence, many of these models assume a perfectly
balanced network and consider only the positive-sequence components of all phe-
nomena. Converters are included in RMS programs using their averaged models.
They provide the behaviour of the converter ignoring fast switching transients and
any control with very small time constants compared to the time steps and the
phenomena that are considered. For instance, in stability analysis there is usually
no need to model the inverter in detail since its transients are much faster than the
dynamics being studied. Therefore, only the fundamental frequency outputs, i.e.,
50 Hz or 60 Hz, of the converter are modelled, which are mainly reflected in the
electrical control model, as shown in Section 3.2.3. Due to this averaged represen-
tation, the modelling of different technologies of inverter-based generation can be
unified using sub-models for each basic component according to their characteristics.
TM
Typical tools for RMS simulations are DIgSILENT PowerFactory R
, GE PSLF ,
PowerWorld Simulator, RAMSES, or Siemens PTI PSS E. It should be noted that
R

the use of different RMS software tools might lead to different results performing
the same stability studies [67, 68]. The investigations presented in this work fall
under the category of positive-sequence RMS simulations and the simulation tool
DIgSILENT PowerFactory [69] is used.
2.3. Time-domain simulation 17

EMT simulation: If the phenomena to be studied are significantly outside of


the range of fidelity of RMS models, e.g., studying electro-magnetic transients or
sub-synchronous torsional interactions, EMT simulation software should be used,
together with detailed equipment specific models. This distinction means that EMT
analysis is capable of representing electro-magnetic transients, the frequency depen-
dence of network components, e.g., change in transmission line resistance and re-
actance with frequency, harmonics, unbalanced networks, power electronic devices,
including the switching transients as well as detailed controls and protection sys-
tems. EMT analysis programs use, for instance, numerical integration substitution
as a method that transforms the basic electrical elements, i.e., resistances, induc-
tances and capacitances, into an equivalent parallel circuit composed of a resistance
and a current source [70]. Users need to be careful on the selection of the type
of models between RMS and EMT. RMS models usually do not include the inner
current control loops of converters, the detailed Phase-Locked Loop (PLL) model
and various other details of the converter. Nevertheless, RMS simulations are much
faster to execute than EMT. Therefore, RMS simulations are dominant for spe-
cific types of power system dynamic studies. EMT models may also be required
under the following special system conditions, where RMS time-domain simulation
programs may not be accurate or suitable anymore, for instance:

• studies of weak system conditions with a very low short circuit ratio,

• studies of detailed inverter and collector system design,

• studies of detailed equipment and system interactions, or

• studies of unbalanced faults.

It should be noted that many RMS models are positive-sequence models. Typical
TM TM
programs for EMT simulations are PSCAD /EMTDC , EMTP-RV, OPAL-RT
R R
HYPERSIM , RTDS or DIgSILENT PowerFactory .

Combined RMS and EMT simulation: Another approach is the combination


of RMS and EMT based models, which is also known as co-simulation or hybrid
simulation [71, 72]. In general, a hybrid or co-simulation is a combination of at
least two different tools for performing a single multiphysics or multimodel simula-
tion [71]. The main advantage of this approach is to combine the accuracy of EMT
with the computational efficiency of RMS simulations. The term hybrid simulation
should not be confused with hybrid dynamical systems [65]. Electrical power sys-
tem models are hybrid dynamical systems because they undergo discrete transitions
during the simulation.
18 Chapter 2. Stability definitions and analysis methods

2.3.2 Power system modelling


The electrical power system is a dynamic system that may be described by a set of
non-linear DAEs [65] in the form of:

ẋ = f (x, y, t) (2.1)
0 = g(x, y, t) (2.2)

where x refers to the state vector, y to the vector of inputs and the time is denoted
as t. The derivative of the state vector is denoted by ẋ. With the initial values of
the DAE system:

x(t0 ) = x0 (2.3)
y(t0 ) = y0 (2.4)

that are usually obtained by performing a power flow computation with the initial
time t0 . It should be noted that the form of algebraic equations g can be twofold,
i.e., a current-injection or power-injection model [65]. The most common model
used in proprietary software packages is the current-injection one [73]. In order to
solve Eqs. (2.1) and (2.2), different numerical integration methods are available.

2.3.3 Numerical integration methods


There is a huge variety of numerical integration methods. In general, these methods
can be separated into explicit and implicit methods, as explained in the sequel.

Explicit methods: Explicit numerical integration methods are well suited for non-
stiff ordinary differential equations. Many different explicit methods exist and for
power system applications, Runge-Kutta’s and the modified Euler’s method have
been widely used [64, 74].

Implicit methods: As time-domain analysis in electrical power systems concerns


the integration of non-linear and generally stiff DAE systems, the most adequate
are implicit or semi-implicit methods. Moreover, another advantage is that implicit
methods can be absolute stable [65]. Examples for implicit methods in power system
analysis are the backward Euler’s or the Trapezoidal method [65]. The time-domain
simulations performed in this work use the Trapezoidal method.

2.3.4 Time step


The time step Δt can be fix or variable throughout the simulation and should be
tuned according to the time constants of the system. The integration time step
for the time-domain analysis in this thesis is 0.01 s, which is typical to study the
phenomena mentioned in Section 2.2.
2.4. Critical clearing time 19

2.4 Critical clearing time


Definition: In this thesis, the CCT is understood based on the definition given in
the well-known report [49] as follows:

The critical clearing time is the maximum permissible duration of the


fault for which the subsequent system response remains stable.

This definition refers to the stability of the overall power system. However, previous
definitions of the CCT were based on transient stability, where the CCT is defined
as the maximum time between the initiation and the isolation of a fault, such that
the power system is transiently stable [75]. In general, the overall goal of the CCT
is to determine the stability margin of the system. Several calculation methods,
which depend on the stability criterion, exist, as shown in Table 2.1. In this work,
the CCT refers to both, short-term voltage stability and transient stability, and is
calculated through time-domain simulations using a trial-and-error strategy, which
is explained in the sequel.

Calculation: To calculate the CCT through time-domain simulations, a trial-and-


error strategy [65] is used. The steps involved in calculating the CCT are outlined
as follows:
(=0)
1. Initial guess for tCCT .

2. Run time-domain simulation.


(+1)
3. If the system is short-term voltage and transient stable, tCCT is increased,
(+1) () (+1) (+1) ()
i.e., tCCT = tCCT + Δt; otherwise tCCT is decreased, i.e., tCCT = tCCT − Δt.

4. Go back to Step 2 until system condition changes, i.e., stable → unstable then
() (+1)
tCCT := tCCT ; or unstable → stable then tCCT := tCCT .

As shown in the trial-and-error strategy, the calculation of the CCT considers both,
short-term voltage and transient stability. If only one of these is unstable, the fault
duration exceeds the CCT. The time step Δt is 0.01 s, as presented in Section 2.3.4.

Table 2.1: Critical clearing time calculation methods

Stability criterion Calculation method Application


Short-term voltage stability Trial and error [20]
Short-term voltage stability Analytical [76]
Short-term voltage stability Lyaponov exponents [54]
Transient stability Trial and error [65]
Transient stability Lyapunov’s function [55]
20 Chapter 2. Stability definitions and analysis methods

2.5 Kullback-Leibler divergence measure


This measure has the objective to analyse the FIDVR phenomenon caused by the
dynamic behaviour of induction motor loads, which tend to decelerate and stall,
following faults in the system. Available methods in literature to quantify FIDVR
can be classified into two categories, namely:

1. Slope-based methods: Metrics based on the slope or the derivative of the


voltage evolution [58, 59] are not the most appropriate in case of oscillations
or sudden discontinuous changes in the voltage.

2. Integral error-based methods: Metrics based on the integral [60, 61] cannot
distinguish between the waveform of the following two voltages. One with a
lower initial voltage drop that takes more time to recover. The other with a
higher initial voltage drop but showing fast recovery over a small period of
time.

To deal with the aforementioned limitations, an entropy-based metric, namely the


KL divergence measure [26], was introduced for the analysis of FIDVR. The word
entropy is referred to information theory, as the average rate at which information
is produced by a stochastic source of data. The overall goal of the KL divergence
measure is to consider the recovery rate of the voltage and its steady-state value that
are two important factors for the characterisation of the FIDVR phenomenon. The
initially developed KL divergence measure [77] has been applied in various fields,
such as statistics [78] and information theory [79]. In the field of power systems,
the KL divergence measure has been used to investigate the:

• dynamic performance of induction motor loads [26], or

• dynamic voltage control areas for placing static var and static synchronous
compensators [27].

In both studies, inverter-based generation was not considered. The objective is


to apply this recently developed metric to analyse the short-term control of PV
systems. The major steps involved in calculating the KL divergence measure for
a specific set of voltage samples are outlined in Sections 2.5.1 to 2.5.4. However,
in some cases the application of the KL divergence measure is limited. These
limitations are outlined in Section 2.5.5.
2.5. Kullback-Leibler divergence measure 21

2.5.1 Voltage violation criterion


To evaluate the quality of the voltage recovery, a criterion is defined that determines
the lowest acceptable voltage. This voltage violation criterion specifies the critical
value of the KL divergence measure. It should be noted that this performance
criterion involves both, temporal and magnitude information of the voltage time
series. In this work, the WECC voltage violation criterion [80] is considered, which
has been used in several studies [26, 27, 81] to evaluate FIDVR. The WECC criterion
states that following an N − 1 contingency, the voltage drop at all load and non-
load buses should not exceed 25 % and 30 % of the nominal voltage, respectively.
Furthermore, the voltage drop at all load buses should not exceed 20 % of the
nominal voltage for more than 20 cycles. An example of the WECC voltage violation
criterion, along with two sample voltage recoveries, is illustrated in Fig. 2.4. The
parameters V1 , V2 , V3 and t1 , t2 denote the voltage and time limits, respectively,
and the values are given in Table A.1 of Appendix A. The critical value of the KL
divergence measure is calculated using the WECC voltage violation criterion, which
forms the boundary for the envelope VWECC according to:


⎨V (t) ≥ V1 for tcl ≤ t < t1
VWECC = V (t) ≥ V2 for t1 ≤ t < t2 and V2 > V1 (2.5)


V (t) ≥ V3 for t2 ≤ t < tf and V3 > V2

The considered voltage time series data is from fault clearing instant tcl to the final
observation time tf with the time limits tcl < t1 < t2 < tf .

Vmax
Vpre

V3
WECC criterion
V2
V1
Voltage

Vgood
Vbad
Vmin
tcl t1 t2 tf
Time

Figure 2.4: WECC voltage violation criterion and sample voltage recoveries
22 Chapter 2. Stability definitions and analysis methods

2.5.2 Probability density function of voltage sample


If the voltage recovers quickly to its steady-state value, the corresponding Probabil-
ity Density Function (PDF) has a high peak near that steady-state value, as shown
in Fig. 2.5. To obtain the PDF, the voltage axis from Vmin to Vmax is partitioned
into L sub-intervals, where the index i refers to the i-th partition. The samples are
observed from the fault clearing instant tcl to the final observation time tf . The cor-
responding parameter values are given in Table A.2 of Appendix A. The number of
samples in a certain sub-interval PiKL is divided by the total number of samples A:

1  KL
L
P KL = P (2.6)
A i=1 i

which leads to the PDF P KL in the form of a normalised distribution. The details
of the construction of the PDF P KL are shown in Algorithm 2.1.
1: procedure PDF for KL measure
2: initialise PiKL = 0 for all i
3: for each voltage interval i
4: for each discrete time t in i
5: PiKL (V (t)) := PiKL (V (t)) + 1
6: endfor
7: endfor
8: divide all PiKL by the total number of discrete times A
9: end procedure

Algorithm 2.1: Pseudocode for construction of probability density function

Vmax

V3

V2 Pgood
KL
of Vgood
V1
Voltage

Pbad
KL
of Vbad

Vmin
0 0.05 0.1 0.15 0.2 0.25 0.3
Probability density function

Figure 2.5: Probability density functions of sample voltage recoveries


2.5. Kullback-Leibler divergence measure 23

2.5.3 Probability density function of reference voltage


recovery
In this step, the PDF of the reference voltage recovery is constructed. This ideal
density function Pref
KL
corresponds to a voltage waveform, which recovers fast to
a defined voltage value, e.g., the pre-disturbance value Vpre . Hence, Pref KL
will be
a Dirac-delta function with all its mass concentrated at Vpre . However, the Dirac-
delta function does not qualify the definition of a PDF P KL . Therefore, the following
approximation is used:
1 −γ(Vi −Vpre )2
Pref,i
KL
= e for i = 1, 2, ..., L (2.7)
W
where the parameter γ controls the concentration of the density near Vpre . The
normalising factor W makes the summation of the reference distribution equal to
one with:
L
2
W = e−γ(Vi −Vpre ) (2.8)
i=1

Assuming a value for the parameter γ of 450 [27], the resulting PDF for the reference
voltage recovery Pref
KL
is shown in Fig. 2.6. The objective is to compare the distance
between the PDF of a given voltage recovery P KL and the PDF of the reference
voltage recovery Pref
KL
using the KL divergence measure K. In the same figure, the
PDF of the WECC voltage violation criterion is depicted, which consists of three
peaks at the voltage limits V1 , V2 and V3 accordingly.

Vmax
Pref
KL
Vpre

V3

V2
V1
Voltage

PWECC
KL

Vmin
0.0 0.1 0.2 0.3 0.4 0.5 0.6
Probability density function

Figure 2.6: Probability density functions of reference voltage recovery and WECC
voltage violation criterion
24 Chapter 2. Stability definitions and analysis methods

2.5.4 Calculation of divergence measure


In the last step, the KL divergence measure K can be calculated using the relative
entropy between two PDFs, P KL and Pref KL
, denoted by K(P KL ||Pref
KL
), and given by:
 
 L
PiKL
K(P KL ||Pref
KL
)= PiKL ln (2.9)
i=1
Pref,i
KL

Combining Eqs. (2.7) and (2.9) yields:


L 
L
K = ln W + PiKL ln PiKL + γ PiKL (Vi − Vpre )2 (2.10)
i=1 i=1

Analysing Eq. (2.10), it can be concluded that if the voltage recovery is delayed,
the KL divergence measure is high. This is due to the fact that PiKL is higher
for smaller i. The smaller Vi , the larger the weighting factor (Vi − Vpre )2 . This
would lead to a higher KL divergence measure K for a delayed voltage recovery. In
general, the value of the KL divergence measure is bounded below by zero, which
corresponds to the PDF of the voltage recovery P KL matching exactly the PDF
of the reference voltage recovery PrefKL
. Therefore, it can be concluded, the smaller
the KL divergence measure K, the better the voltage recovery. According to the
WECC criterion with the envelope VWECC given in Eq. (2.5) and the definitions of
Δt1 = t1 − tcl , Δt2 = t2 − t1 , Δt3 = tf − t2 , and Δt4 = tf − tcl , the critical value of
the KL divergence measure, denoted as KWECC , is calculated according to:
1
KWECC = (Δt1 log Δt1 + Δt2 log Δt2 + Δt3 log Δt3 )
Δt4
γ
+ (Δt1 (V1 − Vpre )2 + Δt2 (V2 − Vpre )2 + Δt3 (V3 − Vpre )2 ) (2.11)
Δt4
+ log W − log Δt4

If the KL divergence measure K is greater than KWECC , the WECC criterion is vio-
lated. It should be noted that the values of the concentration parameter γ and the
sub-intervals L significantly influence the resulting KL divergence measure. Consid-
ering the evolution of the sample voltage recoveries shown in Fig. 2.4 and the PDF
for the reference voltage recovery defined in Eq. (2.7), the KL divergence measures
are calculated accordingly, as shown in Table 2.2. Note that the aforementioned
procedure for the calculation of the KL divergence measure, along with the assumed
parameters, is considered for the stability analysis in Chapters 4 and 5, respectively.

Table 2.2: Kullback-Leibler divergence measures

Kgood [–] KWECC [–] Kbad [–]


0.99 17.96 66.63
2.5. Kullback-Leibler divergence measure 25

2.5.5 Limitations of Kullback-Leibler divergence measure


By way of example, consider the evolution of two voltages depicted in Fig. 2.7.
One shows an ideal, instantaneous voltage recovery, denoted as Videal . In this case,
the voltage reaches immediately its pre-disturbance value after the fault clearing
instant tcl . On the other hand, the second evolution, denoted as Vosc , shows a delayed
voltage recovery with an oscillatory dynamic behaviour during the post-fault period.
It is clear that the voltage recovery of Vosc is worse. Therefore, the KL divergence
measure for the ideal voltage recovery Kideal is expected to be smaller than the
measure for the delayed voltage recovery Kosc . However, the calculated divergence
measures yield Kideal = 1.94 and Kosc = 0.84. As Kideal is higher than Kosc , the
former suggests a lower quality of recovery, which is contradictory. This is due to
the fact that the distance of the PDF of the delayed voltage recovery Posc KL
to the PDF
of the reference voltage recovery Pref
KL
is shorter, i.e., K(Posc
KL
||Pref
KL
) < K(Pideal
KL
||Pref
KL
).
The aforementioned drawback using the KL divergence measure for the analysis of
the FIDVR phenomenon and short-term control of PV systems gives the motivation
to define an improved metric, namely the VRI. The steps involved in calculating
the VRI are outlined in the following sections.

Vmax

V3

V2
V1
Voltage

WECC criterion

Videal
Vosc
Vmin
tcl t1 t2 tf
Time

Figure 2.7: Ideal and oscillatory voltage recovery


26 Chapter 2. Stability definitions and analysis methods

2.6 Voltage recovery index


Due to the limitations of the KL divergence measure mentioned in Section 2.5.5, a
new metric is defined, i.e., the VRI, in order to systematically quantify the severity
of FIDVR. Instead of using an entropy-based metric, the VRI resorts to a weighting
function that rewards (penalises) a good (bad) voltage recovery. The steps involved
in calculating the VRI for a specific voltage waveform are outlined in Sections 2.6.1
to 2.6.4.

2.6.1 Voltage violation criterion


Similar to Section 2.5.1, a voltage violation criterion is defined that determines the
lowest acceptable voltage. The WECC voltage violation criterion [80] is considered
and an example of the envelope VWECC is depicted in Fig. 2.4. The parameters
V1 , V2 , V3 and t1 , t2 denote the voltage and time limits, respectively, and the specific
parameter values are the same as for the calculation of the KL divergence measure,
and given in Table A.1 of Appendix A.

2.6.2 Probability density function of voltage sample


To define the PDF for the VRI, the voltage axis from Vmin to Vmax is partitioned into
L sub-intervals, where the index i refers to the i-th partition. The voltage samples
are observed from the fault clearing instant tcl to the final observation time tf . The
corresponding parameter values are the same as for the KL divergence measure, and
given in Table A.2 of Appendix A. The construction of the PDF itself is different
compared to the PDF construction for the KL divergence measure in Section 2.5.2.
Considering the WECC voltage violation criterion with the envelope VWECC and a
given voltage time series V (t), the rewarding PDF P VRI+ and the penalising PDF
P VRI− can be constructed, as shown in Algorithm 2.2. The number of samples in
1: procedure PDF for VRI
2: initialise PiVRI+ = PiVRI− = 0 for all i
3: for each voltage interval i
4: for each discrete time t in i
5: if V (t) > VWECC (t) then
6: PiVRI+ (V (t)) := PiVRI+ (V (t)) + 1
7: else
8: PiVRI− (V (t)) := PiVRI− (V (t)) + 1
9: endif
10: endfor
11: endfor
12: divide all PiVRI+ and PiVRI− by the total number of discrete times A
13: end procedure

Algorithm 2.2: Pseudocode for construction of probability density function


2.6. Voltage recovery index 27

a particular sub-interval PiVRI is divided by the total number of samples A to obtain


a normalised distribution. Hence, the overall PDF P VRI can be defined as:

1   VRI+
M L
P VRI = (P + PjiVRI− ) (2.12)
A j=1 i=1 ji

where M denotes the number of voltage constraints and the index j refers to the
j-th voltage constraint. It should be emphasised that compared to the PDF for the
KL divergence measure P KL given in Eq. (2.6), the PDF for the VRI P VRI defined
in Eq. (2.12) does not only give information about the distribution of the voltage,
but it also shows its relation to the constraints, i.e., if and when voltage limits are
violated. The specific PDF with regard to the WECC criterion with M = 3, for
instance, consists of six PDFs in total, two according to the voltage violations, i.e.,
V1 , V2 , V3 , times three according to the time intervals, i.e., [tcl , t1 [; [t1 , t2 [; [t2 , tf [,
as given in:
 L 
1  VRI+  VRI+  VRI+
L L
P VRI+ = P1i + P2i + P3i (2.13)
A i=1 i=1 i=1

 
1 
L 
L 
L
P VRI−
= P1i
VRI−
+ P2i
VRI−
+ P3i
VRI−
(2.14)
A i=1 i=1 i=1

For instance, Fig. 2.4 shows a good and a bad voltage recovery and the corresponding
PDFs are constructed in Fig. 2.8.

Vmax

V3

V2 Pgood
VRI
of Vgood
V1
Voltage

VRI+
P1i
Pbad
VRI
of Vbad VRI−
P1i
VRI+
P2i
VRI−
P2i
VRI+
P3i
VRI−
P3i
Vmin
0 0.05 0.1 0.15 0.2 0.25 0.3
Probability density function

Figure 2.8: Probability density functions of sample voltage recoveries


28 Chapter 2. Stability definitions and analysis methods

2.6.3 Weighting function


Instead of using an entropy-based metric as in Section 2.5.4, the VRI applies weight-
ing functions in order to reward (+) or penalise (–) the voltage recovery. The
weighting functions ηj+ and ηj− are defined according to:
+ 2
ηj+ = e−λj (Vi −Vpre ) for j = 1, 2, ..., M and i = 1, 2, ..., L (2.15)

− 2
ηj− = −e−λj (Vi −0) for j = 1, 2, ..., M and i = 1, 2, ..., L (2.16)
where the parameters and λ+
j λ−
control the width of the rewarding and penal-
j
ising weighting function, respectively. The parameter j refers to the j-th voltage
constraint. It should be noted that these distribution functions are very appropri-
ate because voltage oscillations but also overvoltages can be taken into account.

Furthermore, it should be emphasised that the values of λ+ j and λj significantly
influence the resulting VRI. As a general rule, the coefficients should be tuned ac-
cording to the voltage violation criterion. In case of the WECC voltage violation
criterion VWECC , the coefficients are tuned according to:

|η| < σ for Vi = VWECC (2.17)

Assuming a value for the parameter σ of 0.01, and considering the WECC criterion
with M = 3, the so-defined weighting functions are depicted in Fig. 2.9. The corre-
− −
sponding values for the coefficients λ+
1 , ..., λ3 and λ1 , ..., λ3 are listed in Table A.3
+

of Appendix A.

Vmax
Vpre

V3

V2
V1
Voltage

η1+
η1−
η2+
η2−
η3+
η3−
Vmin
-1.0 -0.5 0.0 0.5 1.0
Weighting function

Figure 2.9: Weighting functions to reward or penalise voltage recovery


2.6. Voltage recovery index 29

2.6.4 Calculation of index


Finally, the VRI is calculated using the PDFs, according to Algorithm 2.2 and
Eq. (2.12) as well as the weighting functions defined in Eqs. (2.15) and (2.16). The
correlation between PDF and weighting function and the resulting VRI, denoted as
V, follows:
1   + VRI+
M L
− VRI−
V= (η P + ηji Pji ) (2.18)
A j=1 i=1 ji ji
− VRI−
where the terms ηji Pji
+ VRI+
and ηji Pji describe the rewarding and penalising part,
respectively. Analysing Eq. (2.18), it can be concluded that if the voltage recovery
is delayed, V will be low. This is due to the fact that the penalising PDF PjiVRI− is
high owing to the high concentration of voltage samples below the voltage violation
criterion. Multiplying this high term with the negative weighting function leads
to a low value of the VRI. The lower bound of the VRI corresponds to no voltage
recovery, i.e., V = 0 pu, leading to V = −1. The upper bound of the VRI is
reached for a perfect immediate recovery, i.e., V = Vpre , leading to V = +1. These
clear bounds of the VRI give another advantage of this metric compared to the
KL divergence measure. Moreover, the VRI is “neutral” if the voltage recovers
exactly to the specified voltage violation criterion. For the WECC criterion, for
instance, with a recovery V = VWECC , the VRI results in V = 0. Therefore, it
can be concluded that the higher the value of V, the better the voltage recovery.
Considering the evolution of the sample voltage recoveries shown in Fig. 2.4 and the
weighting functions depicted in Fig. 2.9, the obtained VRIs are given in Table 2.3.
The VRI Vbad for the bad voltage recovery Vbad is slightly negative with −0.06,
whereas the good voltage recovery Vgood leads to a high value with 0.79. It should
be emphasised that the aforementioned procedure for the calculation of the VRI,
along with the assumed parameters, is considered for the analysis of the power
system dynamic performance of the one-load infinite-bus and the multi-machine
system in Chapters 4 and 5, respectively.

Table 2.3: Voltage recovery indices

Vgood [–] VWECC [–] Vbad [–]


0.79 0.00 −0.06
30 Chapter 2. Stability definitions and analysis methods

2.7 Conclusions and future work


Conclusions: This chapter presents relevant stability terms and definitions that
are used throughout this thesis. Among the definitions of short-term voltage stabil-
ity and the phenomenon of FIDVR, which are the main focus of this thesis, also the
definitions of related phenomena, such as transient stability and frequency stability,
are given. Moreover, details about the distinction between short-term voltage and
transient stability are provided. According to the relevant stability criteria, state-
of-the-art approaches as well as recently developed analysis methods are identified.
The most suitable methods, i.e., RMS time-domain simulation, the calculation of
the CCT, and the KL divergence measure, are selected and described in detail.
The CCT determines the stability margin of the system and refers to both, short-
term voltage stability and transient stability. The KL divergence measure is an
entropy-based metric to quantify the phenomenon of FIDVR. Drawbacks of the KL
divergence measure are outlined and a novel enhanced metric, namely the VRI, is
introduced to quantify the phenomenon of FIDVR more systematically. The VRI
applies weighting functions in order to reward or penalise the voltage recovery. The
functions are very appropriate because voltage oscillations but also overvoltages
can be taken into account. Moreover, the VRI has clear bounds, which is another
advantage of this metric compared to the KL divergence measure.

Future work: For very high penetration levels of inverter-based generation and
weak system conditions, RMS time-domain simulations might not be accurate or
suitable anymore. Instead, combined RMS and EMT, or pure EMT analysis is
required. In addition, the KL divergence measure could be improved if the PDF
of the reference voltage recovery is changed, specifically the concentration param-
eter of the density near the pre-fault voltage. Moreover, for the calculation of the
VRI, different weighting functions for overvoltages, i.e., voltages higher than the
normal operating voltage, could be introduced. Furthermore, the application of
other promising metrics, such as Lyapunov exponents or transient P–V curves, are
of interest for the analysis of short-term voltage stability.
31

Modelling and control of photovoltaic systems


3
3.1 Introduction
The content of this chapter is partly based on the following publication:
G. Lammert, L. D. Pabón Ospina, P. Pourbeik, D. Fetzer, and M. Braun,
“Implementation and Validation of WECC Generic Photovoltaic System
Models in DIgSILENT PowerFactory”, in IEEE Power & Energy Society
General Meeting, Boston, 2016.
Motivation: For traditional technologies, such as synchronous or induction ma-
chines, there are well-established dynamic models for power system stability studies.
However, for inverter-based generation, i.e., PV systems and wind turbines, generic
and validated models are still under development and typical parameters are rarely
available.

Contributions: This chapter presents the WECC generic PV system model and
its implementation in DIgSILENT PowerFactory. Typical model parameters of the
PV system are provided that were determined in collaboration with a manufacturer.
The implemented PV system model is validated for several control modes against
another tool that represents the WECC specifications. In order to serve the need
of academia and industry, the implemented model is open-source software and can
be downloaded free of charge and used for any purpose.

Structure: The remainder of this chapter is organised as follows. The implementa-


tion of the WECC generic PV system model, along with the necessary modifications,
is presented in Section 3.2. The relevant control modes that impact short-term volt-
age stability and FIDVR, are identified in Section 3.3. In Section 3.4, typical model
parameters are provided. The test system that is used for the model validation, is
presented in Section 3.5. The detailed validation is outlined in Section 3.6. Sec-
tion 3.7 presents the open-source access to the implemented model. Finally, the
conclusions and directions for future work are given in Section 3.8.
32 Chapter 3. Modelling and control of photovoltaic systems

3.2 Model
A generic PV system model is used to study the impact on power system stabil-
ity. The model follows the WECC specifications [82, 83] and is implemented in
DIgSILENT PowerFactory [69]. The model implementation is detailed in the fol-
lowing sections and additional information is given in reports [31, 84, 85].

3.2.1 Overview
The overall structure of the positive-sequence PV system model is depicted in
Fig. 3.1. The model includes 16 state variables and consists of three parts, namely:

1. Renewable Energy Generator/Converter (REGC A) model: This model is


used to represent the generator/converter interface with the grid. It processes
the active and reactive current commands and outputs the active and reactive
current injections into the grid model. The model enables non-linear current
control capability.

2. Renewable Energy Electrical Control (REEC B) model: This model is used to


represent the electrical control of the inverter. It acts on active and reactive
power references and translates it into active and reactive current commands.
Furthermore, the model limits the current based on the current limit logic.

3. Renewable Energy Plant Control (REPC A) model: This model is optional


and can be used to represent the plant controller. It processes voltage and
reactive power to emulate reactive power control, and frequency and active
power to emulate active power control at plant level.

It should be noted that the REGC A, REEC B, and REPC A models do not in-
clude any protection systems. However, voltage protection is modelled separately
as shown in the sequel. In this work, the phenomena of interest are short-term

9W 9W
REPC_A REEC_B REGC_A
9UHJ
9UHI Plant Level 4H[W ,TFPG’ ,TFPG ,T
4UHI Q Control
V/Q Control Current
4EUDQFK Generator Network
3UHI Limit
Model Solution
3EUDQFK Plant Level ,SFPG’ Logic ,SFPG ,S
3UHI P Control
)UHTBUHI P Control
)UHT

3T)ODJ

Figure 3.1: Block diagram of WECC generic photovoltaic system model [82]
3.2. Model 33

voltage stability and FIDVR, and subsequently transient and frequency stability, as
presented in Section 2.2 and shown in Fig. 2.3. In order to determine the impact of
PV system control on these phenomena, the relevant controls have to be identified.
Therefore, different power system controls, along with the three parts of the WECC
generic PV system model, i.e., the REGC A, the REEC B, and the REPC A model,
are depicted in Fig. 3.2. The fastest control is included in the REGC A model,
followed by the electrical control in the REEC B and the rather slow plant control
in the REPC A model. These relevant controls of the WECC generic PV system
model are highlighted with a solid arrow in Fig. 3.2. Moreover, other controls,
such as the synchronous generator control, the protection system, or the prime
mover control, are important to consider because they influence the aforementioned
power system phenomena. These controls are not investigated in detail in this
work, and hence highlighted with a dashed arrow in Fig. 3.2. The details about
the implementation and the necessary modifications of the REGC A, the REEC B,
and the REPC A model as well as the relevant control modes to study the impact
on power system stability, are outlined in the following sections.

EMT simulation RMS simulation

FACTS control

Synchronous generator control


Power system controls

Generator/converter control (REGC A)

Protection systems

Electrical control (REEC B)


Relevant
controls
Prime mover control

Plant control (REPC A)

10−3 10−2 10−1 100 101 102 103


Time scale [s]

Figure 3.2: Time scales of relevant power system controls


34 Chapter 3. Modelling and control of photovoltaic systems

3.2.2 Generator/converter model


The REGC A model involves three state variables. The block diagram is depicted
in Fig. 3.3 and structured according to its functions into the REGC A 0 and the
REGC A 1 part, as explained in the sequel. It should be noted that the REGC A 1
part is of particular relevance for the investigations carried out in this work. A
complete description of the symbols can be found in the Nomenclature.

REGC A 0: The REGC A 0 part represents the high voltage region and allows
reactive current management during high voltage events. In particular, the high
voltage reactive current management provides an additional reactive current de-
pending on the terminal voltage. Note that the additional reactive current injection
is not activated until the terminal voltage passes a certain over voltage threshold.
This provision of reactive current influences indirectly the terminal voltage of the
inverter. The REGC A 0 part also includes a ramp rate limiter that limits the
rate-of-change of reactive current output.

REGC A 1: The REGC A 1 part represents the low voltage region and enables
active current control during low voltage events. On the one hand, the low voltage
active current management emulates the response of the PLL controls during volt-
age dips. On the other hand, the low voltage power logic allows for a controlled
response of active current during and immediately following voltage dips. Note that
both controls involve a computation using a non-linear map, as shown in Fig. 3.3.
Moreover, the REGC A 1 part includes a ramp rate limitation of the active current
and a measurement delay of the terminal voltage.

REGC_A_0 ,TUPD[

,TFPG 1 ,T
1 + V7JT High Voltage
9W Reactive Current
,TUPLQ
Management
Network Interface

REGC_A_1
/93/ & UUSZU

,SFPG 1 ,S
1 + V7JS
J
Low Voltage
/93/ Active Current
/YSOVZ
0
Management
/YSO
1
1 + V7IOWU
1
]HUR[ EUNSW 9

Figure 3.3: Block diagram of generator/converter model [82]


3.2. Model 35

3.2.3 Electrical control model


The REEC B model incorporates six state variables. The block diagram is shown in
Fig. 3.4 and structured according to its functions into the REEC B 0, REEC B 1,
REEC B 2, REEC B 3 and REEC B 4 part, as described in the sequel. It should
be noted that the REEC B 1 and REEC B 2 part are of particular interest for the
following studies. A complete description of the symbols can be found in the Nomen-
clature.

REEC B 0: The REEC B 0 part exhibits the current limit logic and provides the
option to set the priority either to active or reactive current. To avoid an algebraic
loop, the current limitation is implemented as follows:

Imax if PqFlag = 1 (P priority)


Ipmax = (3.1)
Imax − 2
min(Iqcmd  2 , Imax 2 ) if PqFlag = 0 (Q priority)
Imax if PqFlag = 0 (Q priority)
Iqmax = (3.2)
Imax 2 − min(Ipcmd  2 , Imax 2 ) if PqFlag = 1 (P priority)
Iqmin = −Iqmax (3.3)

where Ipcmd  and Iqcmd  are the output signals of the REEC B 3 and REEC B 2
part, respectively. The maximum apparent current is denoted as Imax , and the
maximum active and reactive current limits are referred to as Ipmax and Iqmax ,
respectively. The value of Iqmin specifies the minimum reactive current limit. The
minimum active current limit, denoted as Ipmin, is zero.
GEG, GEG ,TK
REEC_B_1 1 9WBILOW -
9W  .TY
1 + V7UY
+
(9W < 9GL
if (9W 9GLS)) or (9W
(9W > 9
9XS)
XS)) ,TO
9UHI
9ROWDJHBGLS = 1 1
else
9ROWDJHBGLS = 0 ,TLQM

9PD[
3I)ODJ ,TPD[
4PD[ 9)ODJ
1 9PD[ 4)ODJ + ,TPD[
3H × 1 +
1 + V7S  .TS + .TL 1 +
V  .YS + .YL 1 + ,TFPG
0 - V  ,TFPG
SIDUHI WDQ 4PLQ 0 -
Freeze state i f 9PLQ 0
4JHQ 9PLQ
9PLQ 9ROWDJHBGLS = 1 Freeze state i f
,TPLQ
4H[W 9WBILOW ,TPLQ
,TPLQ 9ROWDJHBGLS = 1
0.01
D
1
÷ ,PD[ Current
1 + V7LT
V7LT
REEC_B_4 REEC_B_2 N Freeze state if
9ROWDJHBGLS = 1 Limit
Logic
3T)ODJ
) J
REEC_B_3
3PD[ & G3PD[ ,SPD[
1 N ,SFPG
3UHI ÷ ,SFPG
1 + V7SRUG
Freeze state i f
D ,SPLQ =0
3PLQ & G3PLQ 9ROWDJHBGLS = 1

9WBILOW
REEC_B_0
0.01

Figure 3.4: Block diagram of electrical control model [82]


36 Chapter 3. Modelling and control of photovoltaic systems

REEC B 1: The REEC B 1 part represents the DVS capability of the PV system
to control the terminal voltage in case of a voltage dip. The voltage is controlled
by the injection of an additional reactive current, i.e., the injection of an additional
capacitive reactive current leads to a voltage increase, whereas the injection of an
additional inductive reactive current leads to a voltage decrease considering the
consumer-oriented sign convention. Moreover, the REEC B 1 part includes the
voltage dip detection function that is activated (1) or deactivated (0) according to:

1 if Vt < Vdip or Vt > Vup (fault condition)


Voltage dip = (3.4)
0 otherwise (normal operating condition)

with the upper and lower voltage limits Vup and Vdip, respectively. This function
determines the state of the PV system. The fault condition is activated if one of the
voltage limits is exceeded. The PV system returns to the normal operating state if
the voltage is within the limits.

REEC B 2: The REEC B 2 part enables different local voltage control strategies,
i.e., local constant reactive power control, local constant voltage control, and local
coordinated reactive power/voltage control. These different local voltage control
modes apply Proportional Integral (PI) controllers that can be selected using the
flags QFlag and VFlag. It should be noted that the implementation of the anti-
windup limiters of the PI controllers is crucial for the dynamic performance of
the PV system. Therefore, it is explained in Appendix B.1.1. Moreover, some
of the control blocks are equipped with a freeze function that allows a distinction
between normal operation and fault condition, as given in Eq. (3.4). The detailed
implementation of the freeze function is also described in Appendix B.1.1.

REEC B 3: The REEC B 3 part provides active power control and generates an
active power reference signal, which is passed through a first order low-pass filter
with anti-windup and ramp rate limits. The generated active power signal is trans-
ferred into an active current signal using the terminal voltage. The details about the
anti-windup limits as well as the freeze function can be found in Appendix B.1.1.

REEC B 4: The REEC B 4 part allows constant power factor control, where the
reactive power is proportional to the active power. This control is modified by using
the tangent of the power factor angle reference pfaref according to:
Qext
tan(pfaref ) = (3.5)
Pref
with the initial active power reference Pref and reactive power reference Qext com-
puted from the power flow solution.
3.2. Model 37

3.2.4 Plant control model


The REPC A model includes seven state variables. The block diagram is pre-
sented in Fig. 3.5 and structured according to its functions into the REPC A 0, the
REPC A 1 and the REPC A 2 part, as explained in the sequel. It should be noted
that the REPC A 1 part is of particular relevance for the investigations in this work.
A complete description of the symbols can be found in the Nomenclature.

REPC A 0: The REPC A 0 part allows line drop compensation at a user-defined


bus. The remote bus is defined as the bus farther from the terminal voltage and
the electrical distance is represented by the impedance, i.e., the resistance and
reactance, between the monitored bus and the remote bus to be controlled.

REPC A 1: The REPC A 1 part controls either voltage or reactive power at plant
level using the flag RefFlag. The closed loop voltage or reactive power regulation
uses a PI controller with anti-windup limits and freeze function and the implemen-
tation is detailed in Appendix B.1.1. Note that the output signal of the REPC A 1
part is either a reference reactive power set point or the equivalent reactive power
needed to regulate the voltage at a specified bus.

REPC A 2: The REPC A 2 part represents the active power control at plant level
and provides a governor response derived from frequency deviation at a user-defined
bus. The frequency droop control is capable of being activated in both directions,
i.e., over and under frequency conditions. For under frequency conditions, power has
to be kept in reserve to allow for a quick active power injection. The active power
regulation uses a PI controller with anti-windup limits and its implementation is
presented in detail in Appendix B.1.1.
9FPS)ODJ 9UHI
,EUDQFK 1 1 - +
|9UHJ–(5F
|9UHJ – (5F + j;F)·,EUDQFK|
9UHJ 1 + V7YIOWU
V7Y
7YIOWU

0 4PD[BSODQW
[
GEG HPD[
+ 1
4EUDQFK
+ .S.L
.S .L 1 + V7IW
.F  5HI)ODJ 4H[W
V 1 + V7IY
REPC_A_0 0
HPLQ
Freeze state
Q
4PLQBSODQW if 9UHJ < 9IU]
1 -

1 + V7T
V7TIOWU
7 IOWU
+
REPC_A_1 4UHI

3ODQWBSUHI
[
3PD[BSODQW )UT)ODJ
IHPD[ 0
+
3EUDQFK 1 - .SJ + .LJ 1 3UHI

1 + V7S 0 V 1 + V7ODJ 1
+ IHPLQ
IGEG, IGEG 'GQ
- + 3PLQBSODQW
Q
)UHT  
+ +
'XS
)UHTBUHI REPC_A_2
0

Figure 3.5: Block diagram of plant control model [82]


38 Chapter 3. Modelling and control of photovoltaic systems

3.3 Control
The WECC generic PV system model can be configured in over 30 unique con-
trol modes [86]. Due to the numerous ways the model can be configured, selecting
the appropriate model structure is vital. Each unique model configuration corre-
sponds to a particular control mode. In this section, the relevant modes of operation
of PV systems that impact power system stability, specifically short-term voltage
stability and FIDVR, are identified. These control modes, i.e., LVRT and DVS ca-
pability, active current recovery rate, local voltage and plant-level voltage control,
describe state-of-the-art requirements and settings according to international stan-
dards and grid codes. The overall goal is to consider control modes, in particular
voltage control modes that influence short-term voltage dynamics. The relevant
control modes are shown in Table 3.1. For each control mode, different parameter
variations and/or flag combinations of the WECC generic PV system model are
incorporated. The relevant parameters and flag combinations, along with the as-
sociated models, are presented in the same table. For all control modes, reactive
current priority with PqFlag = 0 is considered according to international guidelines
and grid codes [7]. Moreover, it is demonstrated that the identified short-term con-
trol modes of PV systems do not only impact voltage stability, but also frequency
stability and transient stability. This is shown in detail for several case studies
considering a multi-machine system, as presented in Chapter 5. It should be noted
that only the positive-sequence control of PV systems is of interest in this thesis.
The relevant control modes of the PV system to study the impact on power system
stability are outlined in the sequel.

Table 3.1: Control modes and relevant parameters

Control mode Section Parameter Model Figure


Low voltage ride-through and Lvplsw REGC A 1 3.3
3.3.1
dynamic voltage support Kqv REEC B 1 3.4
Active current recovery rate 3.3.2 rrpwr REGC A 1 3.3
QFlag
Local voltage control 3.3.3 REEC B 2 3.4
VFlag
QFlag
REEC B 2 3.4
Plant-level voltage control 3.3.4 VFlag
REPC A 1 3.5
RefFlag
3.3. Control 39

3.3.1 Low voltage ride-through and dynamic voltage


support
The following three control modes are considered for the response of the PV system
to faults in the grid, specifically voltage dips. The modes are described in the sequel.

No low voltage ride-through capability (No LVRT): This mode uses the
under voltage protection system of the PV plant, as shown in Fig. 3.6. The focus
in this work is on voltage dips. Thus, only the under voltage function of the pro-
tection system is considered. Due to the voltage sag in the grid, the PV system
immediately disconnects at the Point of Common Coupling (PCC). This setting
represents old German grid code requirements [87, 88], which is still used in some
countries. The considered protection relay is depicted in Fig. 3.7. The measure-
ment block calculates the RMS values out of the momentary values and transfers
them to the under voltage protection function, where the voltage limits and time
delays are defined. The tripping signal is sent to the OR function that opens the
circuit-breaker and disconnects the PV plant. Note that in practice the PV plant
disconnects slightly delayed due to the measurement and inherent delay of the pro-
tection system. However, this time delay is neglected due to its marginal effect on
the dynamic performance of the power system.
External Step-up Photovoltaic
grid transformer system

PV

PCC Voltage
transformer

Under voltage
V < protection function

Figure 3.6: Photovoltaic system with under voltage protection

Va
Vb Trip
(1) V <
Vc
va
Voltage vb Trip Trip
Measurement (2) V < OR
transformer vc circuit-breaker

Trip
(n) V <

Figure 3.7: Under voltage protection system


40 Chapter 3. Modelling and control of photovoltaic systems

Low voltage ride-through and blocking mode (LVRT & block): This con-
trol mode considers two aspects, i.e., the LVRT and the blocking capability of the
PV system. The LVRT capability, or in general FRT capability, allows the PV
system to remain connected to the grid during voltage dips. The considered LVRT
characteristic is presented in Fig. 3.8 and based on the German grid code for the
high voltage level [7]. If the voltage during the fault-on period lies within the no trip
area, the PV system has to stay connected to the grid and ride through the fault.
On the other hand, the PV system may disconnect if the voltage is below the LVRT
characteristic. In practice, the LVRT capability can be enabled using a step-wise
under voltage protection system with n functions, as shown in Fig. 3.7. The second
aspect of this control mode is related to the behaviour of active and reactive current
during the fault-on period. With regard to this, blocking means that the PV sys-
tem does not inject any active or reactive current during the fault-on period. The
blocking mode is also called zero power or momentary cessation mode and is applied
in many inverters in the United States of America [14] and in Germany [89]. The
blocking mode is activated using the static gain in the REEC B 1 part of Fig. 3.4
with Kqv = 0 pu, and the low voltage power logic in the REGC A 1 part of Fig. 3.3
with Lvplsw = 1. The low voltage power logic utilises a non-linear map that reduces
the active current as a function of the voltage. It should be noted that an interme-
diate control option between the blocking mode and the DVS capability, which is
described in the next paragraph, is possible. Instead of reducing active and reactive
power to zero during the fault-on period, this control mode would try to keep them
at their pre-fault values. However, this control method is out of scope.

1.0

0.8 No trip area


Voltage [pu]

0.6
May trip area
0.4

0.2

0.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure 3.8: Low voltage ride-through characteristic [7]


3.3. Control 41

Low voltage ride-through and dynamic voltage support (LVRT & DVS):
This control mode includes two aspects. The first aspect is similar to the former
mode, i.e., the LVRT capability. This capability allows the PV system to remain
connected to the grid during voltage dips according to the LVRT characteristic
shown in Fig. 3.8. The LVRT curve is based on the German grid code for the high
voltage level [7]. The second aspect of this control mode enables voltage support
during the disturbance, which is the DVS capability, also known as dynamic reactive
power support or dynamic voltage control. The DVS capability defines an additional
reactive current injection Iqinj determined in proportion to the voltage deviation
ΔV . The additional reactive current injection is based on the German grid code
for the high voltage level [7] and can be calculated as follows:

Iqinj = Kqv · ΔV (3.6)

with the static gain Kqv . The voltage deviation ΔV is calculated from the filtered
terminal voltage Vt filt and the reference voltage Vref0 according to:

ΔV = Vref0 − Vt filt (3.7)

where Vref0 is determined from the pre-fault condition. The DVS capability is
implemented in the REEC B 1 part of Fig. 3.4 and activated only if a voltage dip
in the system is detected, i.e., Voltage dip = 1. In this control mode, the static
gain is set to Kqv = 4 pu, as shown in Fig. 3.9. This value gives a higher sensitivity
for the additional reactive current in terms of voltage deviations, compared to the
default value of 2 pu.

1.0
Additional reactive current [pu]

0.5
Inductive

0.0

Capacitive Kqv = 4 pu
−0.5

−1.0
−0.5 −0.4 −0.3 −0.2 −0.1 0.0 0.1 0.2 0.3 0.4 0.5
Voltage deviation [pu]

Figure 3.9: Additional reactive current injection for dynamic voltage support [7]
42 Chapter 3. Modelling and control of photovoltaic systems

3.3.2 Active current recovery rate


The following four control modes relate to the active current recovery rate during
the post-fault period that determines the time until the active power of the PV
system is restored after a disturbance. The considered recovery rates are depicted
in Fig. 3.10. By way of example, a fault occurs at t = 0.0 s and is cleared at
t = 0.1 s. The details to the different rates are explained in the sequel.

Very slow recovery (10 %/s): This mode represents a recovery rate of rrpwr =
10 %/s to restore the active current of the PV plant. The value of the active current
recovery rate is widely used in the United States of America [14].

Slow recovery (20 %/s): This control mode describes an active current recovery
rate of rrpwr = 20 %/s. The value is the lower limit defined by the German grid
code for the high voltage level [7].

Medium recovery (100 %/s): This mode defines a recovery rate of rrpwr =
100 %/s to restore the active current of the PV system. The value is the lower limit
defined by the grid code in Great Britain for faults with a duration of more than
140 ms [90].

Fast recovery (1000 %/s): This control mode specifies a recovery rate of rrpwr =
1000 %/s to restore the active current. The value is the lower limit defined by the
grid code in Great Britain for faults with a duration of less than 140 ms [90].

0.0 rrpwr = 10 %/s


rrpwr = 20 %/s
rrpwr = 100 %/s
−0.2 rrpwr = 1000 %/s
Active current [pu]

−0.4

−0.6

−0.8

−1.0

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0
Time [s]

Figure 3.10: Active current recovery rates


3.3. Control 43

3.3.3 Local voltage control


The following three control modes refer to different quasi-stationary voltage control
strategies during normal operation, i.e., Voltage dip = 0. The objective is to in-
vestigate the influence of fast and slow voltage controls on power system stability.
The LVRT and DVS capability of PV systems belongs to fast voltage control, as
described in Section 3.3.1, whereas the local voltage control is slower. The different
local voltage control strategies can be selected through specific flag combinations,
as shown in Table 3.2. The local voltage control modes are explained in detail in
the sequel.

Local reactive power control (Local Q ctr): This mode allows local constant
reactive power control using the feed-forward loop in the REEC B 2 part of Fig. 3.4
with QFlag = 0. The reactive power is kept at the value given by the initial power
flow computation.

Local voltage control (Local V ctr): This mode enables local constant voltage
control of the PV system using the second PI controller in the REEC B 2 part of
Fig. 3.4 with QFlag = 1 and VFlag = 0. The voltage is kept at its initial power
flow value.

Local coordinated reactive power/voltage control (Local Q/V ctr): This


mode allows local coordinated reactive power/voltage control using the two PI con-
trollers in series in the REEC B 2 part of Fig. 3.4 with QFlag = 1 and VFlag = 1.
This control mode is a combination of both local controls, i.e., local constant voltage
and local constant reactive power control. Local coordinated reactive power/voltage
control allows a fast voltage control to restore the voltage quickly after the fault,
followed by a slow reactive power control that brings the PV system back to its
reactive power set point in steady state. This mode is of particular interest if the
grid voltage is rather stiff and cannot be controlled through local constant voltage
control of the PV system.

Table 3.2: Local voltage control modes

Control mode Required models VFlag QFlag


Local Q ctr REGC A + REEC B N/A∗ 0
Local V ctr REGC A + REEC B 0 1
Local Q/V ctr REGC A + REEC B 1 1
∗ The designation “N/A” means that the flag has no effect
on the control mode, and hence may be set to either 0 or 1.
44 Chapter 3. Modelling and control of photovoltaic systems

3.3.4 Plant-level voltage control


The following four control modes involve different plant-level voltage control strate-
gies during normal operation, i.e., Voltage dip = 0. The goal is to investigate the
impact of the rather slow plant control on the dynamic performance of the power
system. The different control modes can be selected via specific flag combinations,
as shown in Table 3.3. The details of the specific control modes are explained in
the sequel.

Plant-level reactive power control (Plant Q ctr): This mode allows constant
reactive power control of the PV system using the lower path in the REPC A 1
part of Fig. 3.5 with RefFlag = 0 and the feed-forward loop in the REEC B 2 part
of Fig. 3.4 setting QFlag = 0.

Plant-level voltage control (Plant V ctr): This mode enables constant voltage
control of the PV system using the upper path in the REPC A 1 part of Fig. 3.5
with RefFlag = 1 and the feed-forward loop in the REEC B 2 part of Fig. 3.4 with
QFlag = 0.

Plant-level reactive power and local coordinated reactive power/voltage


control (Plant Q & local Q/V ctr): This control mode is a combination of
plant-level constant reactive power control in the REPC A 1 part of Fig. 3.5 with
RefFlag = 0, together with local coordinated reactive power/voltage control in the
REEC B 2 part of Fig. 3.4 with QFlag = VFlag = 1. The mode uses the fast local
control for a quick voltage recovery, followed by the rather slow plant-level constant
reactive power control to restore the pre-fault reactive power.

Plant-level voltage and local coordinated reactive power/voltage control


(Plant V & local Q/V ctr): The last mode is a combination of plant-level con-
stant voltage control in the REPC A 1 part of Fig. 3.5 with RefFlag = 1, together
with local coordinated reactive power/voltage control in the REEC B 2 part of
Fig. 3.4 with QFlag = VFlag = 1. This control mode applies the fast local control
that enables a quick voltage support after fault clearing, followed by the rather slow
plant-level constant voltage control to restore the pre-fault voltage.
Table 3.3: Plant-level voltage control modes

Control mode Required models VFlag QFlag RefFlag


Plant Q ctr REGC A + REEC B + REPC A N/A∗ 0 0
Plant V ctr REGC A + REEC B + REPC A N/A∗ 0 1
Plant Q & local Q/V ctr REGC A + REEC B + REPC A 1 1 0
Plant V & local Q/V ctr REGC A + REEC B + REPC A 1 1 1
∗ Thedesignation “N/A” means that the flag has no effect on the control mode, and
hence may be set to either 0 or 1.
3.3. Control 45

3.3.5 Combinations of control modes


The control modes presented in Sections 3.3.1 to 3.3.4, are combined for several
cases. An overview of all cases is given in Table 3.4. The relevant control mode
combinations are highlighted in grey and directly linked to the parameter variations
in Table 3.1. The control modes with respect to the LVRT and DVS capability of PV
systems, as detailed in Section 3.3.1, are summarised in the first part of Table 3.4
(cases A1–A3). The operating modes with regard to the active current recovery
rate of PV systems, as described in detail in Section 3.3.2, are summarised in the
second part of Table 3.4 (cases B1–B3). The control modes considering local voltage
control of PV systems, as detailed in Section 3.3.3, are summarised in the third part
of Table 3.4 (cases C1–C2). The control modes with respect to plant-level voltage
control of PV systems, as described in detail in Section 3.3.4, are summarised in the
fourth and last part of Table 3.4 (cases D1–D4). The presented cases in Table 3.4
are validated in Section 3.6 and investigated in detail considering a one-load infinite-
bus system as well as a multi-machine system, as shown in Sections 4.3 and 5.3,
respectively.

Table 3.4: Combinations of control modes

Plant Q & local Q/V ctr

Plant V & local Q/V ctr


LVRT & block

Local Q/V ctr


LVRT & DVS

Case
Local Q ctr

Plant Q ctr
Local V ctr

Plant V ctr
No LVRT

1000 %/s
100 %/s
10 %/s

20 %/s

A1 × × ×
A2 × × ×
A3 × × ×
B1 × × ×
B2 × × ×
B3 × × ×
C1 × × ×
C2 × × ×
D1 × × × ×
D2 × × × ×
D3 × × × ×
D4 × × × ×
46 Chapter 3. Modelling and control of photovoltaic systems

3.4 Parameters
In general, model parameters are crucial for power system dynamic studies. Be-
cause many grid codes do not define the detailed specification or characteristic of
the inverter control, the model parameters could be different depending on the
manufacturer of the inverter. Even if the model parameters of one inverter can be
identified through field validation, it is almost infeasible to identify the parameters
of all inverters connected to the power system. Therefore, one of the main contri-
butions of this thesis is to provide a set of typical model parameters. In that sense,
typical parameters represent the average behaviour of presently required control
modes and settings according to international guidelines and grid codes. In par-
ticular, the WECC generic PV system model consists of 75 model parameters in
total. The identification process and the specific parameter values for the REGC A,
REEC B and REPC A model are presented in the sequel.

3.4.1 Generator/converter model


The REGC A model of the WECC generic PV system incorporates 15 model pa-
rameters. Realistic parameter values are determined in close collaboration with a
manufacturer and tuned in order to fit field measurements of a typical inverter.
The resulting parameter values of the REGC A model are shown in Table B.1 of
Appendix B. Note that the parameter values of the low voltage power logic Lvplsw
and the maximum rate-of-change of active current rrpwr are of interest in this thesis
and vary for specific control modes, as presented in Table 3.1.

3.4.2 Electrical control model


The REEC B model of the WECC generic PV system includes 29 model parameters.
Realistic parameter values are determined in close collaboration with a manufac-
turer and tuned in order to fit field measurements of a typical inverter. The resulting
parameter values of the REEC B model are given in Table B.2 of Appendix B. Note
that the parameter values of the reactive current injection gain Kqv , the reactive
power flag QFlag, and the voltage flag VFlag are of interest in this thesis and vary
for specific control modes, as shown in Table 3.1.

3.4.3 Plant control model


The REPC A model of the WECC generic PV system involves 31 model parame-
ters. The parameter values are mainly based on the WECC specifications [82, 83].
The resulting parameter values of the REPC A model are shown in Table B.3 of
Appendix B. Note that the parameter value of the plant-level reference control flag
RefFlag is of interest in this thesis and varies for specific control modes, as presented
in Table 3.1.
3.5. Test system 47

3.5 Test system


3.5.1 Overview
For the model validation, which is carried out in Section 3.6, a single-converter
equivalent of the large-scale PV system [91] is considered, as depicted in Fig. 3.11.
It should be noted that this equivalent system is not necessarily appropriate for
small PV units connected to the distribution system. In Fig. 3.11, the PV system
equivalent is connected to the step-up transformer equivalent and via the collector
system equivalent to the substation transformer. The substation transformer is
connected to the external grid, which is represented by a voltage source. It should
be noted that an interconnection transmission line can be optionally used to connect
the substation transformer to the external power system, as depicted in Fig 4.4. The
PV system equivalent represents the total generating capacity of all the inverters of
the PV plant. The step-up transformer equivalent stands for the aggregate effect of
all step-up transformers. The collector system equivalent represents the aggregate
effect of the PV plant collector system. The impedance of the substation transformer
tends to be high compared to the collector system equivalent. Therefore, it should
be modelled explicitly. The measurement and control actions are carried out at
bus 4. This bus is defined as the PCC. The detailed test system parameters are
given in Table B.4 of Appendix B. It should be noted that the parameters do not
necessarily represent a realistic system. The intent here is simply to validate the
implementation of the WECC generic PV system model.

3.5.2 Operating point


For the model validation, the large-scale PV plant is operated at unity power fac-
tor. The active and reactive power injection of the PV system is −0.9 pu and 0 pu,
respectively. It should be noted that the consumer-oriented sign convention is con-
sidered, i.e., consumed active power as well as inductive reactive power are positive
and generated active power and capacitive reactive power are negative.

External Substation Collector Step-up transformer Photovoltaic


grid transformer system equivalent equivalent system equivalent

PV

Bus 4 (PCC) Bus 3 Bus 2 Bus 1


110 kV 20 kV 20 kV 0.6 kV

Figure 3.11: Single-converter equivalent of large-scale photovoltaic system


48 Chapter 3. Modelling and control of photovoltaic systems

3.5.3 Contingency
The contingency that is considered for the model validation, can be characterised
by the fault location, fault impedance, fault type, and fault duration [92]. The fault
is emulated using a reference voltage dip, as depicted in Fig. 3.12. This voltage
dip is created by the voltage source, which represents the external grid in Fig. 3.11.
The dynamic behaviour of the PV system model is tested in response to this voltage
dip. The details to the reference voltage dip are outlined in the sequel.

Fault location: The reference voltage dip represents a fault located in the external
grid, as shown in Fig. 3.11.

Fault impedance: As the fault is emulated, the fault impedance corresponds to


a voltage dip leading to a remaining voltage at bus 4 of 0.5 pu, as presented in
Fig. 3.12.

Fault type: The WECC generic PV system model corresponds to a positive-


sequence stability model. Therefore, the reference voltage dip emulates a three-
phase short circuit.

Fault duration: For the validation of the PV system model, a fault duration of
0.1 s is considered, as depicted in Fig. 3.12. The fault occurs at t = 0.0 s and is
cleared at t = 0.1 s. According to the LVRT characteristic shown in Fig. 3.8, the
PV system needs to stay connected to the grid during the short circuit and ride
through the fault.

1.0

0.8
Voltage [pu]

0.6

0.4

0.2

0.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure 3.12: Reference voltage dip for model validation


3.6. Validation 49

3.6 Validation
The implemented WECC generic PV system model, as presented in Section 3.2, is
validated against the EPRI written REMV tool [93] using the test system described
in Section 3.5. The REMV tool represents the WECC specifications of the generic
PV system model. In general, this tool is applied to validate simulations against
field measurements of a PV park. In this work, the tool is used to validate the
PV system model implemented in DIgSILENT PowerFactory. It is important to
mention that the REMV tool is successfully used for the validation of real measure-
ments by numerous manufacturers [35, 94, 95]. The REMV tool provides active and
reactive power reference signals. The focus of this work are short-term dynamics in
the order of a few seconds up to several seconds. Therefore, current signals of the
PV system are more appropriate instead of power signals. Due to this fact, current
signals are calculated out of power signals using the corresponding voltage. This
conversion could lead to minor deviations between the REMV tool and PowerFac-
tory. In addition, deviations are likely because of different numerical integration
methods of the software simulation tools, as presented in Section 2.3.3. Moreover,
the modifications made within the model implementation, as outlined in Section 3.2,
could also lead to minor deviations. It should be noted that model validation is
generally difficult to quantify [86]. Mathematical norms, such as the sum of squared
error, the Euclidean norm, or the root mean square error, can serve as useful metrics
that are also used, to some extent, for model reduction techniques [96]. However,
the application of these norms are out of scope in this work. The validation of the
WECC generic PV system model, considering the control modes given in Table 3.4,
is presented in detail in the following sections.

3.6.1 Low voltage ride-through and dynamic voltage


support
This section describes the validation of the LVRT and DVS capability. The corre-
sponding control modes of the PV system are detailed in Section 3.3.1. The simu-
lation results of the currents in response to the voltage dip are given in Figs. B.5
to B.8 of Appendix B. The implemented PowerFactory model matches very well
with the REMV tool, and therefore the WECC specifications. Minor deviations
are possible due to the conversion of power signals into current signals, as described
before. Due to this calculation, active and reactive current spikes of the REMV tool
can be observed upon clearing of the voltage dip, as depicted in Figs. B.7 and B.8
of Appendix B, respectively. These spikes can be attributed to the fact that the
corresponding voltage is slightly shifted in time. It should be noted that the dis-
connection of the PV plant (A1) via the under voltage protection function, is not
considered for validation, as the protection system is not included in the REMV
tool.
50 Chapter 3. Modelling and control of photovoltaic systems

3.6.2 Active current recovery rate


This section outlines the validation of the active current recovery rate. The corre-
sponding control modes of the PV system are presented in Section 3.3.2. The sim-
ulation results of the currents in response to the voltage dip are given in Figs. B.9
to B.14 of Appendix B. The curves show that the implemented PowerFactory model
matches very well with the REMV tool, and therefore the WECC specifications.
The observed spikes of the reactive current that exhibit values outside of the cur-
rent limit range, can be attributed to the conversion of power signals into current
signals, as also mentioned in Section 3.6.1.

3.6.3 Local voltage control


This section presents the validation of the local voltage control. The corresponding
control modes of the PV system are presented in Section 3.3.3. The simulation
results of the currents in response to the voltage dip are given in Figs. B.15 to B.18
of Appendix B. In case of local constant voltage control (C1), the set point Qext,
which is the input of the REEC B 2 part, as shown in Fig. 3.4, can be twofold.
Either, the set point takes the initial value of the signal Qgen, or the initial value of
the signal Vt filt, which leads to the simulation results shown in Figs. B.15 and B.16
of Appendix B. The REMV tool uses the initialisation with Qgen, where both curves
match. However, as constant voltage control is required, the initialisation with
Vt filt is considered. In case of local coordinated reactive power/voltage control
(C2), the curves in Figs. B.17 and B.18 of Appendix B show that the implemented
PowerFactory model matches very well with the REMV tool, and therefore the
WECC specifications. The observed current spikes are caused by the conversion of
power signals into current signals, as outlined in the sections before.

3.6.4 Plant-level voltage control


This section describes the validation of the last control mode, i.e., plant-level volt-
age control. The corresponding control modes of the PV system are presented
in detail in Section 3.3.4. The simulation results of the currents in response to
the voltage dip are given in Figs. B.19 to B.26 of Appendix B. The implemented
PowerFactory model matches very well with the REMV tool, and therefore the
WECC specifications. Minor deviations are possible due to the conversion of power
signals into current signals, as mentioned before. Finally, the validation results of
the different control modes show that the overall accuracy is maintained, which
verifies the adequate implementation of the WECC generic PV system model in
DIgSILENT PowerFactory.
3.7. Open source 51

3.7 Open source


In recent years, there has been much effort on the development of validated mod-
els for inverter-based generation. This work has been primarily related to wind
generation. Now, further attention is starting to be devoted to PV systems and
other technologies. In general, there is still a lack of well-validated and generally
accepted dynamic simulation models, particularly for PV systems, for the use in
power system dynamic studies. Moreover, system operators do usually not cre-
ate their own models. They request validated models from manufacturers, either
proprietary or adjusted generic models. This request poses extra work for the man-
ufacturer. Thus, the importance of developing and providing reliable and flexible
generic models should be noted. For wind generation, these models are already
being widely used. However, generic PV system models are not widely used yet.
Although accepted generic models are provided, the model parameters are crucial
for power system dynamic studies. Because many grid codes do not define the de-
tailed specification or characteristics of the inverter control, the model parameters
could be different depending on the manufacturer. Therefore, realistic model pa-
rameters need to be provided. In order to serve this need of academia and industry,
the implemented and validated WECC generic PV system model in DIgSILENT
PowerFactory is provided as open-source software. The model is publicly available
and it can be download free of charge and used for any purpose. The users have
access to the entire source code, which means the model is unencrypted and all
equations and control blocks are accessible. The equations and control blocks can
be modified and adapted to the individual needs. Moreover, the open-source model
provides typical parameters. The parameters are identified in collaboration with
a manufacturer, as outlined in Section 3.4, and set as default values in the open-
source model. Since July 2016, the WECC generic PV system model is publicly
available. Until February 2019, the model was requested 80 times in total, as shown
in the last row of Table 3.5. Out of these 80 requests, 43 model requests came
from academia and 37 requests from industry. It is worth to mention that academia
uses the model mainly for research, whereas the industry is more focused on field
validation of existing PV plants. It should be noted that the WECC generic PV
system model implemented in DIgSILENT PowerFactory is distributed all over the
world and used in 30 countries on 6 continents.
Table 3.5: Diffusion of WECC generic photovoltaic system model

Model distribution Number


Model requests from academia 43
Model requests from industry 37
Model requests from countries 30
Model requests from continents 6
Model requests in total 80
52 Chapter 3. Modelling and control of photovoltaic systems

3.8 Conclusions and future work


Conclusions: This chapter presents the detailed implementation of the positive-
sequence WECC generic PV system model in DIgSILENT PowerFactory. Each
part of the PV system model, i.e., the generator/converter model, the electrical
control model, and the plant control model, is described in detail. The required
modifications, such as the adjusted current limit logic to avoid an algebraic loop,
are presented. The relevant control modes that impact specifically short-term volt-
age stability and FIDVR, are identified and described in detail. Various test cases
are composed by combinations of these control modes. These cases are studied
considering a one-load infinite-bus and a multi-machine system, as outlined in the
following sections. Furthermore, typical model parameters of the PV system are
provided that were determined in consultation with a manufacturer. The imple-
mented model, along with the identified parameters, is successfully validated for
the defined control modes against the EPRI written REMV tool that represents
the WECC specifications. In order to serve the need of academia and industry, the
WECC generic PV system model, which is implemented in DIgSILENT PowerFac-
tory, is provided as an open-source model and can be downloaded free of charge and
used for any purpose.

Future work: Future work includes the validation of additional control modes
and parameters of the WECC generic PV system model. In this context, negative-
sequence control for unbalanced system conditions is of interest. Moreover, the
WECC generic PV system model is highly non-linear as it includes freeze functions
and deadbands. Therefore, the model is not necessarily appropriate for small-signal
stability analysis. However, if linearisation and eigenvalue analysis is of interest,
closed-form equations of the system should be developed based on strict linearisation
of the non-linear model. Furthermore, in weak systems the model may not be
adequate anymore to capture all relevant power system phenomena. To overcome
this problem, the inner-current control loops as well as detailed PLL modelling
should be considered. Moreover, the generator/converter model implementation as
a current source poses numerical challenges for state-of-the-art software simulation
tools, which can be solved using a voltage source model. Grid-forming control of
PV systems is also of interest, as it seems a promising method to improve power
system stability.
53

Stability analysis of a one-load infinite-bus system


4
4.1 Introduction
The content of this chapter is partly based on the following publication:

G. Lammert, J. C. Boemer, D. Premm, O. Glitza, L. D. Pabón Ospina,


D. Fetzer, and M. Braun, “Impact of Fault Ride-Through and Dynamic
Reactive Power Support of Photovoltaic Systems on Short-Term Voltage
Stability”, in IEEE PowerTech, Manchester, 2017.

Motivation: The fundamental concepts and principles of short-term voltage sta-


bility and FIDVR are well known for synchronous generator dominated systems.
However, the impact of inverter-based generation, such as PV systems, on these
phenomena, is not yet fully understood.

Contributions: This chapter provides a comprehensive stability analysis of a one-


load infinite-bus system considering different PV system control modes. The impact
of PV systems on short-term voltage stability is studied and fundamental concepts
are explained in detail. Furthermore, an extensive investigation is outlined on how
PV systems could enhance FIDVR. Moreover, advantages and drawbacks of the
analysis methods, i.e., time-domain simulation, CCT, KL divergence measure and
VRI, are given. Finally, this chapter gives recommendations for the adequate control
of PV systems in order to improve the dynamic performance of the power system.

Structure: The remainder of this chapter is organised as follows. The one-load


infinite-bus test system that is considered for the stability analysis, is presented in
Section 4.2. This section includes details about the load and the PV system model,
the operating point as well as the considered contingency. Section 4.3 presents the
case studies, specifically the results of the impact analysis of different PV system
control modes on short-term voltage stability and FIDVR. Finally, the conclusions
and directions for future work are given in Section 4.4.
54 Chapter 4. Stability analysis of a one-load infinite-bus system

4.2 Test system


In this section, a simple test system is introduced. This test system is used to
present fundamental concepts and principles of power system stability, specifically
short-term voltage stability and FIDVR, considering a high penetration level of
PV generation. The stability of the test system is studied in response to large
disturbances, i.e., severe system faults. The considered test system is based on a
simple but representative example system [52] and adapted for the specific needs
to study short-term voltage stability and FIDVR. Similar test systems [97] are
used to investigate not only short-term voltage stability, but also long-term voltage
stability, transient stability or frequency stability. The details of the considered test
system, i.e., the overview of the single-line diagram, the load model, the modelling
approach of the large-scale PV system, the pre-disturbance operating point, and
the considered contingency, are described in the sequel.

4.2.1 Overview
The single-line diagram of the one-load infinite-bus test system is depicted in Fig. 4.1.
The term “one-load” refers to the one composite load equivalent connected to bus 5,
which represents the distribution system. In order to keep the test system as sim-
ple as possible, while retaining the essential of voltage instability phenomena, the
following modelling assumptions are made, as presented in the sequel. It should be
noted that the load and the PV system model are of particular interest. Therefore,
these models are detailed in Sections 4.2.2 and 4.2.3, respectively.

Exponential
load
External Transmission Step-down
grid line 1 transformer

Transmission Small induction


IM
line 2 motor load

Shunt
Bus 6 capacitor Large induction
220 kV IM
motor load
Bus 7
20 kV
Interconnection
transmission line
Single-converter equivalent of large-
scale photovoltaic system (Fig. 4.4)
Bus 5 (PCC)
220 kV

Figure 4.1: One-load infinite-bus test system


4.2. Test system 55

External grid: The external system corresponds to an infinite bus, which is rep-
resented by a voltage source of constant voltage magnitude and constant frequency,
i.e., 220 kV and 50 Hz, respectively. It should be noted that the distinction between
short-term voltage instability, which is related to induction motors, and transient
instability, which is related to synchronous generators, is not always clear, and as-
pects of both phenomena may exist, as outlined in Section 2.2.5. Therefore, the
detailed dynamic characteristics of the synchronous generators that are included in
the external grid, are neglected in order to focus on the basic mechanism of short-
term voltage stability. This assumption is usually less critical for short-term voltage
stability.

Transmission lines: The transmission lines are represented by π-equivalents and


connect the external system with the load center and the PV plant. The lines
consider typical values for voltages, resistances, reactances and susceptances. It
is important to note that both transmission lines are equal with regard to the
modelling approach and the parameter values. The corresponding parameter values
are based on report [98] and can be found in Table C.1 of Appendix C.

Step-down transformer: The step-down transformer connects the load center to


the rest of the power system. It should be noted that due to the power system
phenomena of interest, i.e., short-term voltage stability and FIDVR, the dynamics
of the load tap changer can be neglected because they have intentional delays of
about 30 s, which is outside of the relevant time scale. Therefore, the transformer
tap ratio can be assumed as fix. The step-down transformer considers typical values
for apparent power, voltage levels, resistance, reactance and tap ratio. The corre-
sponding parameter values are based on report [98] and can be found in Table C.1
of Appendix C.

Shunt capacitor: The shunt capacitor is connected to the load bus in order to
compensate the aggregated load. The value of the shunt capacitor is fix and adjusted
to achieve a voltage value of 1.0 pu at the load bus (bus 5) considering normal
operating conditions.
56 Chapter 4. Stability analysis of a one-load infinite-bus system

4.2.2 Load modelling


Load dynamic response is a key mechanism of power system voltage stability, driv-
ing the dynamic evolution of voltages and, in extreme cases, leading to voltage
collapse [52]. Therefore, accurate load modelling is crucial for the investigation of
short-term voltage stability and delayed voltage recovery, as also mentioned in Sec-
tions 2.2.1 and 2.2.4. Power system loads are generally complex in composition and
dynamic behaviour, time varying, different in various parts of the system, statis-
tical in nature, and may include a myriad of continuous and discrete controls and
protections [99]. According to the well-known reports [100, 101], the load model
can be categorised into a static and dynamic part. The static part can be charac-
terised by algebraic equations, whereas the dynamic part is described by differential
equations due to its time variant behaviour. Considering the one-load infinite-bus
system depicted in Fig. 4.1, the load center is connected via the step-down trans-
former to the rest of the power system. The load composition with respect to
the static and dynamic part, is 70 % and 30 %, respectively. The details of the
load model are outlined in the sequel. Furthermore, the modelling assumptions are
briefly discussed.

Static part: The static part of the load is represented by the well-known exponen-
tial load model that is defined according to:
α
Vbus
Pexp = P0 (4.1)
V0

β
Vbus
Qexp = Q0 (4.2)
V0
where Pexp and Qexp are the active and reactive power, respectively, consumed by
the load at the bus voltage Vbus . P0 and Q0 are the active and reactive power,
respectively, under the reference voltage V0 , obtained from the initial operating
conditions. Three particular cases for the load exponents are noteworthy:

• α = β = 0: constant power load

• α = β = 1: constant current load

• α = β = 2: constant impedance load

Other exponents can be used to represent the aggregate effect of different load
components. Since the dynamic part is represented explicitly, a constant admittance
is assumed. The parameter values of the static load part are shown in Table C.2 of
Appendix C. It should be noted that the performed power system dynamic studies
consider rather small frequency deviations. Therefore, the frequency dependency of
loads is not addressed in this work.
4.2. Test system 57

Dynamic part: The dynamic part consists of two equivalent induction motor
loads. The induction motor is a crucial component in short-term voltage stability
and FIDVR studies [52], because:
• it is a fast restoring load in the time frame of a second;
• it is a low power factor load with a high demand of reactive power; and
• it is prone to stalling, when voltage is significantly depressed.
In power system dynamic studies, aggregated induction motor models are usually
considered, i.e., one motor representing a large number of similar motors fed by
the same substation [52]. Because of significantly different characteristics, it may
be desirable to model equivalents for both small and large induction motors [101].
Therefore, the dynamic part of the load model consists of two equivalent, single-cage
induction motors. Each equivalent represents 15 % of the load consumption, i.e.,
15 % of small and 15 % of large industrial motors, which is in total 30 % of load
consumption. Typical parameter values of the small and large induction motors
are considered and shown in Tables C.3 and C.4 of Appendix C, respectively. The
values are based on the well-known report [101]. Moreover, a compensation shunt
capacitor is connected in parallel to the induction motor in order to match the
reactive power consumption of the entire load in steady state. Each equivalent
induction motor is represented by a third-order model. The induction motor model
uses the rotor winding fluxes and the rotor speed as state variables. Assuming a
three-phase induction motor with constant resistance, the well-known equivalent
circuit is shown in Fig. 4.2. The induction motor rotates at a speed ωimr = ωims
determined by the induction motor slip, according to:
ωims − ωimr
sim = (4.3)
ωims
with the speed of the stator and rotor ωims and ωimr , respectively. In Fig. 4.2, the
induction motor stator and rotor leakage reactances are denoted as Xims and Ximr ,
respectively, whereas the magnetising reactance is denoted as Ximm . The induc-
tion motor stator and rotor resistances are denoted as Rims and Rimr , respectively.
The terminal voltage of the induction motor is referred to as Vim . Considering the

Rims Xims Ximr

Vim Ximm Rimr /sim

Figure 4.2: Equivalent circuit of induction motor [52]


58 Chapter 4. Stability analysis of a one-load infinite-bus system

equivalent circuit in Fig. 4.2, and the substitution by:

jXimm (Rims + jXims )


Rime + jXime = (4.4)
Rims + j(Xims + Ximm )

the well-known slip–torque characteristic is obtained:


Rimr
Vim
2
Ximm
2
sim
Te (Vim , sim ) = 2 (4.5)
Rimr
Rime + sim
+ (Xime + Ximr )2 (Rims
2
+ (Xims + Ximm )2 )

The differential equation of the rotor motion dynamics has the form of:
dωimr
2Him = Te (Vim , ωimr ) − Tm (ωimr ) (4.6)
dt
with the induction motor inertia constant Him and the electrical and mechanical
torques Te and Tm , respectively. The mechanical torque is assumed constant with:

Tm (ωimr ) = T0 (4.7)

where T0 is the torque value determined from the initial power flow computation, as-
suming the motor is operating in steady state. The characteristic given in Eq. (4.5),
is illustrated in Fig. 4.3 for different terminal voltages Vim . The stable equilibria
are the intersection points close to sim = 0 pu, where the mechanical torque Tm is

3.0
Te (Vim = 1.0 pu)
Te (Vim = 0.75 pu)
2.5 Te (Vim = 0.5 pu)
Tm
2.0
Torque [pu]

Unstable
1.5

1.0

0.5
Stable

0.0
1.0 0.8 0.6 0.4 0.2 0.0
Slip [pu]

Figure 4.3: Slip–torque characteristic of induction motor


4.2. Test system 59

equal to the electrical torque Te . If a voltage dip occurs, e.g., to Vim = 0.75 pu or
Vim = 0.5 pu, the electrical torque is reduced and the induction motor decelerates.
As long as the mechanical torque is less or equal than the electrical torque, the
motor is attracted by its stable equilibrium and able to reaccelerate. However,
if the mechanical torque is larger than the electrical torque, the induction motor
decelerates beyond the unstable equilibrium and stalls. The lower the voltage at
the terminals of the induction motor or the longer the fault duration, the higher
the risk of motor stalling.

Discussion on modelling assumptions: Accurate load modelling assumptions


are crucial for power system dynamic studies, specifically for voltage stability stud-
ies, as outlined in Section 2.2.1. The increased integration of demand-side controls
and distributed generation, particularly distributed PV systems, further compli-
cates load characteristics and poses additional challenges to load modelling [102].
There are still no widely accepted dynamic models that adequately represent dis-
tributed generation in bulk power system stability studies. Therefore, the majority
of academia and industry uses a load representation that typically consists of a
static and dynamic part [103]. The impact of this modelling approach as well as
the change of the load characteristic, on power system stability, are briefly discussed.

• Sensitivity to load composition: A composite load model is considered for the


power system dynamic studies, which consists of a static and dynamic part.
The higher the percentage of the dynamic part, the more the power system is
prone to short-term voltage instability [104].

• Sensitivity to exponential load active and reactive power exponent α, β: A


constant admittance is assumed for the stability studies. With the increase of
battery energy systems in the grid, e.g., in the form of electric vehicles, the
load exponents of active and reactive power would decrease. This would, in
turn, deteriorate the dynamic performance of the power system. It should be
noted that the constant power load is the most susceptible type of load to
voltage stability problems [105].

• Sensitivity to induction motor mechanical torque Tm : The mechanical torque


is assumed constant for the stability studies. This is usually more critical for
short-term voltage stability [52]. It should be noted that at steady state, a
constant torque motor model is equivalent to a constant power load [52].

• Sensitivity to induction motor inertia constant Him : The inertia constants of


the dynamic loads correspond to small and large industrial motors. The lower
the inertia constants are, the quicker the induction motors decelerate [104].
60 Chapter 4. Stability analysis of a one-load infinite-bus system

4.2.3 Modelling of photovoltaic system


The PV system is modelled according to Chapter 3. The model is generic and
based on the WECC specifications [82, 83]. Furthermore, the model represents a
large-scale PV plant connected to the high or extra high voltage level. An overview
of the PV system model as well as its implementation is given in Section 3.2. The
control modes that are studied considering the one-load infinite-bus system are de-
tailed in Section 3.3. Typical parameters that represent the average behaviour of
presently required control modes and settings according to international guidelines
and grid codes, are determined in collaboration with a manufacturer. The consid-
ered parameter values of the PV system model are shown in Tables B.1 to B.3 of
Appendix B. The single-converter equivalent system of the entire large-scale PV
plant is depicted in Fig. 4.4. The PV system equivalent is connected to the step-up
transformer equivalent and via the collector system equivalent to the substation
transformer. The substation transformer is connected to the rest of the power sys-
tem via the interconnection transmission line. It should be noted that within the
PV plant, no shunt compensation is considered. The measurement and control ac-
tions are carried out at the PCC (bus 5). Parameter values of the single-converter
equivalent system are based on reports [98, 106] and can be found in Table C.5 of
Appendix C. The entire equivalent system of the large-scale PV plant, as shown
in Fig. 4.4, is similar to the system used for the model validation, as depicted in
Fig. 3.11. Only the interconnection transmission line is added that connects the
substation transformer to the rest of the power system. The entire single-converter
equivalent of the large-scale PV system is connected to the one-load infinite-bus
test system, as depicted in Fig. 4.1.

Interconnection Substation Collector Step-up transformer Photovoltaic


transmission line transformer system equivalent equivalent system equivalent

PV

Bus 5 (PCC) Bus 4 Bus 3 Bus 2 Bus 1


220 kV 220 kV 20 kV 20 kV 0.6 kV

Figure 4.4: Single-converter equivalent of large-scale photovoltaic system


4.2. Test system 61

4.2.4 Operating point


The demand of the load center is covered by two parts, as shown in Fig. 4.1. One
part of the power is provided by the external grid through rather long transmission
lines. The other part of the power is provided by the large-scale PV plant. In
particular, the active and reactive power consumption and generation of the ele-
ments in the one-load infinite-bus system is shown in Table 4.1. Please note that
the consumer-oriented sign convention is used in this work, i.e., consumed active
power as well as inductive reactive power are positive and generated active power
and capacitive reactive power are negative. In normal operation, half of the active
power demand of the load center is covered by the large-scale PV system that op-
erates at unity power factor. Hence, a PV penetration level PV pen of nearly 50 %
is reached. The penetration level is calculated according to:

PPV
PV pen =   (4.8)
PPV + PSG

where PPV is the sum of the PV system active power generation, divided by
the total active power generation in the system. Thisincludes also the sum of
the generation by synchronous generators, denoted as PSG . It should be noted
that the active power provided by the external grid Pext , as shown in Fig. 4.1,
 also
refers to the sum of the active power generation by synchronous generators PSG .
Considering the operating points of load and generation detailed in Table 4.1, the
power flow computation of the one-load infinite-bus test system leads to the power
flow results depicted in Table C.6 of Appendix C.

Table 4.1: Load and generation of one-load infinite-bus system

Description Symbol Value Unit


Photovoltaic system active power PPV −300 [MW]
Photovoltaic system reactive power QPV 0 [Mvar]
Shunt reactive power Qsh −370 [Mvar]
External grid active power Pext −315 [MW]
External grid reactive power Qext −17 [Mvar]
Exponential load active power Pexp 420 [MW]
Exponential load reactive power Qexp 137 [Mvar]
Small induction motor load active power Psim 90 [MW]
Small induction motor load reactive power Qsim 30 [Mvar]
Large induction motor load active power Plim 90 [MW]
Large induction motor load reactive power Qlim 30 [Mvar]
62 Chapter 4. Stability analysis of a one-load infinite-bus system

4.2.5 Contingency
The contingency that is considered to study the impact of PV system control on
power system stability, specifically short-term voltage stability and delayed voltage
recovery, can be characterised by the fault location, fault impedance, fault type,
and fault duration [92], as detailed in the sequel.

Fault location: The fault is located on transmission line 2, as shown in Fig. 4.1.
It takes place exactly in the middle of the line between bus 5 and bus 6, i.e., 50 %
of the line length. Therefore, the fault location highly influences the power transfer
from the external grid to the load center.

Fault impedance: The fault impedance is zero, which corresponds to a solid


short circuit. This impedance leads to a severe voltage dip that highly influences
the dynamic behaviour of PV systems. Moreover, the resulting voltage dip also
affects the operation of the small and large induction motor load.

Fault type: The fault type is a three-phase short circuit that is considered for the
analysis. This type of fault leads to a severe voltage sag that, again, highly impacts
the operation of the large-scale PV system as well as the small and large induction
motor load.

Fault duration: The last characteristic of the contingency is the fault duration.
For the stability analysis of the one-load infinite-bus system, a fault duration of 0.1 s
is considered. The fault is cleared by opening the transmission line, which remains
open. The fault duration is typical for protection systems according to international
grid codes. Moreover, the large-scale PV system needs to stay connected to the grid
during the short circuit and ride through the fault. This requirement is defined by
the LVRT characteristic shown in Fig. 3.8.
4.3. Case studies 63

4.3 Case studies


This section analyses the impact of different PV system control modes on power
system stability, specifically short-term voltage stability and FIDVR, considering
the one-load infinite-bus system. The relevant control modes are combined and
yield a set of cases, as shown in Table 3.4. All of the cases are studied based on the
analysis methods described in Chapter 2, i.e., time-domain simulation, CCT, KL
divergence measure, and VRI. The results of the case studies are presented in the
sequel.

4.3.1 Low voltage ride-through and dynamic voltage


support
This section investigates the impact of the LVRT and DVS capability of the PV
system on short-term voltage stability and delayed voltage recovery. The corre-
sponding cases aim at comparing the disconnection of the large-scale PV system
(A1) with the blocking mode (A2) as well as with the LVRT and DVS capability
(A3), as shown in Table 3.4. The aforementioned control modes are described in
detail in Section 3.3.1.
The time-domain results in response to the fault are gathered in Figs. 4.5 to 4.10.
Due to the disconnection of the PV system (A1), the small and large induction
motor load connected to bus 7 stall because their rotor slips exceed the unstable
equilibrium values, as revealed by the slips shown in Figs. 4.9 and 4.10. This causes
short-term voltage instability, as shown by the voltage at the PCC and the load bus
in Figs. 4.5 and 4.6, respectively. If the PV system remains connected to the grid
but reduces its active and reactive current to zero during the fault-on period (A2),
short-term voltage instability still takes place, as shown by the load bus voltage
in Fig. 4.6. After the fault is cleared at t = 0.1 s, the voltage is locked at a low
level. Again, the cause for the voltage collapse is the stalling of the small and
large induction motor loads, as depicted in Figs. 4.9 and 4.10, respectively. The
slips show that the LVRT capability of the PV system in this case (A2), slightly
delays the stalling effect of the motors due to the delayed active current reduction.
For both cases (A1 and A2), the WECC criterion, as defined in Section 2.5.1, is
significantly violated, which is illustrated in Fig. 4.6. It should be noted that due
to the depressed voltage caused by the small and large induction motor stalling, the
active current of the PV system remains zero. The best dynamic performance is
achieved with the LVRT and DVS capability of the PV system (A3), as shown by
the load bus voltage in Fig. 4.6. In this case, the PV system rides through the fault
and feeds in an additional capacitive reactive current during the fault-on period, as
shown by the reactive current in Fig. 4.8. As a result, short-term voltage instability
can be avoided and FIDVR is significantly improved. Moreover, using the LVRT
and DVS capability (A3), the WECC voltage criterion is barely violated, as shown
in Fig. 4.6.
64 Chapter 4. Stability analysis of a one-load infinite-bus system

1.0

0.8
Voltage [pu]

0.6

0.4 LVRT characteristic

0.2
A1
A2
0.0 A3

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure 4.5: Impact of low voltage ride-through and dynamic voltage support
 Voltage at bus 5 (point of common coupling)

1.0

0.8

WECC criterion
Voltage [pu]

0.6

0.4

0.2
A1
A2
0.0 A3

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure 4.6: Impact of low voltage ride-through and dynamic voltage support
 Voltage at bus 7 (load bus)
4.3. Case studies 65

0.0

−0.2
Active current [pu]

−0.4
A1
−0.6 A2
A3

−0.8

−1.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure 4.7: Impact of low voltage ride-through and dynamic voltage support
 Active current of photovoltaic system connected to bus 5

0.0

−0.2
Reactive current [pu]

−0.4

−0.6

−0.8
A1
A2
−1.0 A3

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure 4.8: Impact of low voltage ride-through and dynamic voltage support
 Reactive current of photovoltaic system connected to bus 5
66 Chapter 4. Stability analysis of a one-load infinite-bus system

1.0

0.8

0.6
Slip [pu]

A1
0.4 A2
A3

0.2

0.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure 4.9: Impact of low voltage ride-through and dynamic voltage support
 Slip of small induction motor load connected to bus 7

1.0

0.8

0.6
Slip [pu]

A1
0.4 A2
A3

0.2

0.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure 4.10: Impact of low voltage ride-through and dynamic voltage support
 Slip of large induction motor load connected to bus 7
4.3. Case studies 67

The stability margin is determined in terms of the CCT and the results are
shown in Table 4.2. The best dynamic performance is highlighted in grey. The
CCT is significantly increased from 0.07 s to 0.16 s comparing the disconnection of
the PV system (A1) with the LVRT and DVS capability of the PV system (A3).
Considering the blocking mode of the PV system (A2), the same CCT is calculated
as for the disconnection case (A1), i.e., 0.07 s.
The KL divergence measure is applied to analyse the voltage recovery and give
information about the effectiveness of short-term voltage controls. The results are
shown in Table 4.2. In the same manner, the KL divergence measure K7 that is
calculated from the voltage at the load bus (bus 7), shows the best performance
using the LVRT and DVS capability of the PV system (A3), which yields a value
of 3.49, as highlighted in grey. For the disconnection (A1) and the blocking mode
(A2) of the PV system, the same K7 value of 102.84 each, is obtained. Therefore,
no distinction can be made between these two short-term control modes of the PV
system.
The last metric that is used to evaluate the impact of LVRT and DVS, is the
VRI and the results are shown in Table 4.2. The best dynamic performance is
highlighted in grey. The same tendency as for the other metrics is observed. The
disconnection of the PV system (A1) results in an unacceptable V7 value of −0.13,
considering the voltage recovery at the load bus (bus 7). On the other hand, the best
performance is achieved using the LVRT and DVS capability of the PV system (A3),
with a VRI of 0.58. If only the LVRT capability of the PV system is incorporated
(A2), the calculation of the VRI yields −0.12. Compared to the KL divergence
measure, the disconnection of the PV system (A1) and the blocking mode (A2) can
be distinguished using the VRI, which shows the superiority of this metric.

Table 4.2: Metrics of low voltage ride-through and dynamic voltage support

Case tCCT [s] K7 [–] V7 [–]


A1 0.07 102.84 −0.13
A2 0.07 102.84 −0.12
A3 0.16 3.49 0.58
68 Chapter 4. Stability analysis of a one-load infinite-bus system

4.3.2 Active current recovery rate


This section studies the impact of the active current recovery rate of the PV sys-
tem on delayed voltage recovery. The corresponding cases differ by the ramp rate
rrpwr of the active current. The active current ramp rate is investigated for the
values 10 %/s (B1), 20 %/s (B2), 100 %/s (A3) and 1000 %/s (B3), as presented
in Table 3.4. The details to the aforementioned control modes can be found in
Section 3.3.2.
The time-domain results in response to the fault are illustrated in Figs. 4.11
to 4.14. It should be noted that due to the specific control modes considered in
this section, the simulation time is extended until 10.0 s in order to capture the
relevant dynamic phenomena. The very slow active current recovery rate of 10 %/s
(B1) leads to significantly delayed voltage recoveries with voltages as low as 0.87 pu
at 1.0 s, as shown by the load bus voltage in Fig. 4.12. Higher ramp rates, such
as 1000 %/s (B3), accelerate the voltage recovery, leading, for instance, to 0.97 pu
voltage at 1.0 s, as illustrated in the same figure. It is noteworthy that due to the
low voltages caused by the slow active current ramping of 10 %/s (B1) and 20 %/s
(B2), the additional reactive current injection of the PV system during the fault
is extended until 2.0 s and 1.0 s, respectively, to support the voltage, even though
the fault has been already cleared. This phenomenon is illustrated in Fig. 4.14.
Moreover, using the very slow active current ramp rate of 10 %/s (B1), leads to
a violation of the WECC voltage criterion, as shown in Fig. 4.12. In general,
considering the overall voltage dynamics, it can be concluded that the higher the
active current ramp rate, the better the delayed voltage recovery. However, it
should be noted that fast active current recovery rates could also lead to frequency
problems, as outlined in Section 5.3.3. It is important to mention that due to
the opening of transmission line 2 caused by the fault, the impedance between the
external grid and the load center increases. This leads to a voltage decrease at the
PCC (bus 5) and the load bus (bus 7). In order to keep the active power injection of
the PV system constant, the voltage decrease is encountered with an active current
increase, as presented in Fig. 4.13. The pre-fault active current injection is −0.94 pu
for t < 0 s, whereas the post-fault active current injection in steady state is −0.98 pu
for t > 10.0 s, as illustrated in the same figure. Finally, it should be noted that for
all control modes investigated in this section, no short-term voltage instability is
present, which is also depicted by the slips of the small and large induction motor
loads in Figs. C.1 and C.2 of Appendix C, respectively.
4.3. Case studies 69

1.0

0.8
LVRT characteristic
Voltage [pu]

0.6
1.00
0.95
0.4
0.90
0.85
0.2 0.80 B1
B2
0.0 0.5 1.0 1.5 2.0 2.5 3.0 A3
0.0 B3

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0
Time [s]

Figure 4.11: Impact of active current recovery rate


 Voltage at bus 5 (point of common coupling)

1.0

0.8 WECC criterion


Voltage [pu]

0.6 1.00
0.95
0.4 0.90
0.85
0.80
0.2 B1
0.75
B2
0.0 0.5 1.0 1.5 2.0 2.5 3.0 A3
0.0 B3

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0
Time [s]

Figure 4.12: Impact of active current recovery rate


 Voltage at bus 7 (load bus)
70 Chapter 4. Stability analysis of a one-load infinite-bus system

0.0 B1
B2
A3
−0.2 B3
Active current [pu]

−0.4

−0.6

−0.8

−1.0

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0
Time [s]

Figure 4.13: Impact of active current recovery rate


 Active current of photovoltaic system connected to bus 5

0.0

−0.2
Reactive current [pu]

−0.4 0.05
0.00
−0.6
−0.05
−0.10
−0.8 B1
−0.15 B2
0.0 0.5 1.0 1.5 2.0 2.5 3.0 A3
−1.0 B3

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0
Time [s]

Figure 4.14: Impact of active current recovery rate


 Reactive current of photovoltaic system connected to bus 5
4.3. Case studies 71

The CCT is calculated in order to determine the stability margin of the system.
The results are presented in Table 4.3. The best dynamic performance is highlighted
in grey. In general, the same tendency is observed as for the time-domain results.
The best performance is achieved with a medium and fast active current recovery
rate of 100 %/s (A3) and 1000 %/s (B3), respectively, which leads to a CCT of
0.16 s. Slower ramp rates, such as 10 %/s (B1) and 20 %/s (B2), lead to a smaller
CCT of 0.15 s, respectively.
The delayed voltage recovery is analysed using the KL divergence measure and
the results are given in Table 4.3. Again, the same trend is observed as for the CCT
and the time-domain results. The fastest active current ramp rate of 1000 %/s
(B3) yields the best KL divergence measure K7 that is calculated from the load bus
voltage (bus 7), with a value of 3.13, as highlighted in grey. On the other hand, the
lowest performance is reached if the PV system injects the active current with a
ramp rate of 10 %/s (B1), which results in a K7 value of 6.94. Furthermore, the KL
divergence measure for the slow active current recovery of 20 %/s (B2) is higher,
with a value of 4.82, compared to the medium active current recovery of 100 %/s
(A3), with a value of 3.49. Hence, the latter results in a better quality of voltage
recovery.
Finally, the impact of the active current recovery rate on the phenomenon of
FIDVR is assessed using the newly defined VRI and the results are shown in Ta-
ble 4.3. The best dynamic performance is highlighted in grey. Generally speaking,
a similar tendency is observed as for the aforementioned metrics. In the same man-
ner, the fastest active current recovery rate of 1000 %/s (B3) results in a high V7
value of 0.60, considering the voltage recovery at the load bus (bus 7). The worst
VRI is calculated considering the very slow active current recovery rate of 10 %/s,
which yields a V7 value of 0.15. If only a slow (B2) or medium (A3) active current
recovery rate of the PV system is incorporated, the calculation of the VRI yields
0.26 or 0.58, respectively.

Table 4.3: Metrics of active current recovery rate

Case tCCT [s] K7 [–] V7 [–]


B1 0.15 6.94 0.15
B2 0.15 4.82 0.26
A3 0.16 3.49 0.58
B3 0.16 3.13 0.60
72 Chapter 4. Stability analysis of a one-load infinite-bus system

4.3.3 Local voltage control


This section investigates the impact of the slower local voltage control of the PV
system on voltage dynamics in the power system. The corresponding cases aim
at comparing different local voltage control strategies, i.e., local constant reactive
power control (A3), local constant voltage control (C1), and local coordinated re-
active power/voltage control (C2), as shown in Table 3.4. Further details to the
specific control modes can be found in Section 3.3.3.
The time-domain results in response to the large disturbance are shown in
Figs. 4.15 to 4.18. Similar to the investigation of the active current recovery rate in
Section 4.3.2, the simulation time is extended until 10.0 s in order to capture the rel-
evant dynamic phenomena that are influenced by the different local voltage control
modes of the PV system. In the case of local constant reactive power control (A3)
of the PV system, the reactive power in steady state is kept at the value given by
the initial power flow computation, namely zero, which is shown in Fig. 4.18. This
control method results in a post-fault steady-state voltage deviation, as depicted by
the PCC voltage in Fig. 4.15. The pre-fault voltage at the PCC is 1.0 pu for t < 0 s,
whereas the post-fault voltage in steady state is 0.96 pu for t > 10.0 s, as illustrated
in the same figure. On the other hand, in the case of local constant voltage control
(C1), the PV system restores its pre-fault voltage by injecting reactive current into
the grid, as depicted in Fig. 4.18. In this case (C1), the local constant voltage
control acts fast and brings the voltage back to its pre-fault value, namely 1.0 pu,
as shown by the PCC voltage in Fig. 4.15. In order to keep the voltage constant,
the post-fault reactive current injection in steady state for t > 2.0 s takes a value
of −0.08 pu in order to keep the PCC voltage at its initial power flow value. The
third and last control mode in this section is the coordinated reactive power/voltage
control (C2). This mode of operation combines constant reactive power (A3) with
constant voltage control (C1). The coordinated reactive power/voltage control (C2)
allows a fast voltage control to restore the voltage quickly after the fault, as shown
by the PCC voltage in Fig. 4.15, followed by a slow reactive power control that
brings the PV system back to its reactive power set point in steady state, which
is depicted by the reactive current in Fig. 4.18. In other words, for dynamics up
to 1.0 s, the coordinated reactive power/voltage control (C2) is similar to constant
voltage control (C1), whereas for dynamics up to 10.0 s, the control is related to
constant reactive power control (A3). Finally, it should be noted that for all con-
trol modes investigated in this section, no short-term voltage instability is present,
which is also illustrated by the slips of the small and large induction motor loads
in Figs. C.3 and C.4 of Appendix C, respectively.
4.3. Case studies 73

1.0

0.8
LVRT characteristic
Voltage [pu]

0.6
1.00
0.95
0.4
0.90
0.85
0.2 0.80
A3
0.0 0.5 1.0 1.5 2.0 2.5 3.0 C1
0.0 C2

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0
Time [s]

Figure 4.15: Impact of local voltage control


 Voltage at bus 5 (point of common coupling)

1.0

0.8 WECC criterion


Voltage [pu]

0.6 1.00
0.95
0.4 0.90
0.85
0.80
0.2
0.75 A3
0.0 0.5 1.0 1.5 2.0 2.5 3.0 C1
0.0 C2

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0
Time [s]

Figure 4.16: Impact of local voltage control


 Voltage at bus 7 (load bus)
74 Chapter 4. Stability analysis of a one-load infinite-bus system

0.0 A3
C1
C2
−0.2
Active current [pu]

−0.4

−0.6

−0.8

−1.0

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0
Time [s]

Figure 4.17: Impact of local voltage control


 Active current of photovoltaic system connected to bus 5

0.0

−0.2
Reactive current [pu]

−0.4
0.00
−0.05
−0.6
−0.10
−0.15
−0.8 −0.20
A3
0.0 0.5 1.0 1.5 2.0 2.5 3.0 C1
−1.0 C2

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0
Time [s]

Figure 4.18: Impact of local voltage control


 Reactive current of photovoltaic system connected to bus 5
4.3. Case studies 75

The stability margin of the system is determined using the CCT and the results
are shown in Table 4.4. The best dynamic performance is highlighted in grey.
Compared to the time-domain simulations discussed before, the same trend for the
CCT can be observed. The best dynamic performance is achieved using either local
constant voltage control (C1) or local coordinated reactive power/voltage control
(C2), leading to CCT values of 0.17 s each. Considering local constant reactive
power control (A3), leads to the worst performance with a CCT value of 0.16 s.
The KL divergence measure is applied in order to assess the impact of the local
voltage control methods of the PV system on the delayed voltage recovery. The
results are given in Table 4.4. Similar to the analysis methods applied before, the
same trend is observed. Local constant voltage control (C1) of the PV system
results in the best KL divergence measure K7 that is calculated from the load bus
voltage (bus 7), with a value of 2.32, as highlighted in grey. On the other hand,
the lowest performance is reached if the PV system applies local constant reactive
power control (A3), which results in a K7 value of 3.49. The combination of both
local controls, i.e., coordinated reactive power/voltage control (C2), leads to a KL
divergence measure K7 of 2.41.
The last metric that is used to analyse the impact of local voltage control strate-
gies is the VRI. The results are given in Table 4.4 and the best performance is
highlighted in grey. In the same manner, the local constant voltage control (C1)
results in the best VRI with a V7 value of 0.95, considering the voltage recovery
at the load bus (bus 7). The worst VRI is calculated considering local constant
reactive power control (A3) of the PV system, which yields a V7 value of 0.58. If
coordinated reactive power/voltage control (C2) is applied, the calculation of the
VRI yields 0.75.

Table 4.4: Metrics of local voltage control

Case tCCT [s] K7 [–] V7 [–]


A3 0.16 3.49 0.58
C1 0.17 2.32 0.95
C2 0.17 2.41 0.75
76 Chapter 4. Stability analysis of a one-load infinite-bus system

4.3.4 Plant-level voltage control


The last section analyses the impact of the relatively slow plant-level voltage control
of the PV system on voltage dynamics. The associated cases compare different
voltage control strategies at plant level, i.e., plant-level constant reactive power
control (D1), plant-level constant voltage control (D2), plant-level constant reactive
power and local coordinated reactive power/voltage control (D3) as well as plant-
level constant voltage and local coordinated reactive power/voltage control (D4),
as presented in Table 3.4. Further details to the aforementioned control modes are
presented in Section 3.3.4.
The results of the time-domain simulations in response to the fault are provided
in Figs. 4.19 to 4.22. In order to capture the relevant dynamics that are influenced
by the plant-level voltage control of the PV system, the simulation time is extended
until 20.0 s. Focusing on dynamics in the order of a few seconds, i.e., 0 s < t < 3 s,
plant-level constant reactive power control (D1) and plant-level constant voltage
control (D2) yield very similar behaviour, as shown by the voltage at the PCC in
Fig. 4.19. In the same time scale, plant-level constant reactive power and local coor-
dinated reactive power/voltage control (D3) as well as plant-level constant voltage
and local coordinated reactive power/voltage control (D4) act similarly and lead to
a less oscillatory bus voltage during the recovery period, as also shown in Fig. 4.19.
The aforementioned correlations of the voltage dynamics are caused by the sim-
ilarities of the active and reactive current control of the PV system, as depicted
in Figs. 4.21 and 4.22, respectively. On the contrary, focusing on dynamics up to
20 s, different control mode similarities are observed. Plant-level constant reactive
power control (D1) and plant-level constant reactive power and local coordinated
reactive power/voltage control (D3) act in a similar manner and bring the reactive
power back to its pre-fault value, i.e., zero, as shown by the reactive current of the
PV system in Fig. 4.22. On the other hand, plant-level constant voltage control
(D2) and plant-level constant voltage and local coordinated reactive power/voltage
control (D4) try to restore the pre-fault voltage, as shown by the slow injection
of reactive current in Fig. 4.22 and the resulting voltage at the PCC in Fig. 4.19.
From the aforementioned observations it can be concluded, the best performance
considering both voltage dynamics is achieved using plant-level constant voltage
and local coordinated reactive power/voltage control (D4) of the PV system. Fi-
nally, it should be noted that for all plant-level control modes investigated in this
section, no short-term voltage instability is present, which is also shown by the slips
of the small and large induction motor loads in Figs. C.5 and C.6 of Appendix C,
respectively.
4.3. Case studies 77

1.0

0.8
LVRT characteristic
Voltage [pu]

0.6
1.00
0.95
0.4
0.90
0.85
0.2 0.80 D1
D2
0.0 0.5 1.0 1.5 2.0 2.5 3.0 D3
0.0 D4

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0 20.0
Time [s]

Figure 4.19: Impact of plant-level voltage control


 Voltage at bus 5 (point of common coupling)

1.0

0.8 WECC criterion


Voltage [pu]

0.6 1.00
0.95
0.4 0.90
0.85
0.80
0.2 D1
0.75 D2
0.0 0.5 1.0 1.5 2.0 2.5 3.0 D3
0.0 D4

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0 20.0
Time [s]

Figure 4.20: Impact of plant-level voltage control


 Voltage at bus 7 (load bus)
78 Chapter 4. Stability analysis of a one-load infinite-bus system

0.0 D1
D2
D3
−0.2 D4
−0.90
Active current [pu]

−0.4
−0.95

−0.6
−1.00
0.0 0.5 1.0 1.5 2.0 2.5 3.0
−0.8

−1.0

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0 20.0
Time [s]

Figure 4.21: Impact of plant-level voltage control


 Active current of photovoltaic system connected to bus 5

0.0

−0.2
Reactive current [pu]

−0.4 0.05
0.00
−0.6
−0.05
−0.10
−0.8 D1
−0.15 D2
0.0 0.5 1.0 1.5 2.0 2.5 3.0 D3
−1.0 D4

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0 20.0
Time [s]

Figure 4.22: Impact of plant-level voltage control


 Reactive current of photovoltaic system connected to bus 5
4.3. Case studies 79

The calculation of the CCT determines the stability margin of the system. The
results are given in Table 4.5 and the best dynamic performance is highlighted in
grey. Compared to the time-domain simulations discussed before, a similar trend
for the CCT can be observed. The best CCT is achieved considering either plant-
level constant reactive power and local coordinated reactive power/voltage control
(D3) or plant-level constant voltage and local coordinated reactive power/voltage
control (D4), leading to values of 0.17 s, respectively. In contrast, using plant-level
constant reactive power control (D1) or plant-level constant voltage control (D2) of
the PV system, the CCT is worsened with values of 0.16 s, respectively.
The impact on FIDVR is analysed using the KL divergence measure. The re-
sults are shown in Table 4.5 and the best performance is highlighted in grey. Plant-
level constant reactive power and local coordinated reactive power/voltage control
(D3) or plant-level constant voltage and local coordinated reactive power/voltage
control (D4), lead to the same K7 value of 2.41 each. Hence, no distinction can
be made between these two control modes, which is a disadvantage of the KL
divergence measure. The lowest performance is reached if plant-level constant re-
active power control (D1) is used, leading to a value of 3.49. It should be noted
that local constant reactive power control (A3), as presented in Section 4.3.3, and
plant-level constant reactive power control (D1), lead to the same KL divergence
measures. The same correlation exists between coordinated reactive power/voltage
control (C2) and plant-level constant reactive power and local coordinated reactive
power/voltage control (D3). The similarities are caused due to the same behaviour
of short-term controls of the PV system.
Finally, the results of the VRI are given in Table 4.5. The best VRI is achieved
considering plant-level constant voltage and local coordinated reactive power/voltage
control (D4), which results in a V7 value of 0.76, as highlighted in grey. Compared
to the KL divergence measure, the former control mode (D4) and plant-level con-
stant reactive power and local coordinated reactive power/voltage control (D3) can
be distinguished using the VRI, which reveals the superiority of this metric. The
worst performance value is calculated if plant-level constant reactive power control
(D1) is applied, leading to a V7 value of 0.58. Note that the same relations between
local and plant-level voltage control are observed for the VRI, as described for the
KL divergence measure in the paragraph before.

Table 4.5: Metrics of plant-level voltage control

Case tCCT [s] K7 [–] V7 [–]


D1 0.16 3.49 0.58
D2 0.16 2.95 0.65
D3 0.17 2.41 0.75
D4 0.17 2.41 0.76
80 Chapter 4. Stability analysis of a one-load infinite-bus system

4.4 Conclusions and future work


Conclusions: This chapter presents a comprehensive stability analysis of a one-
load infinite-bus system considering different control modes of a PV system. The
impact of these modes on power system stability, specifically short-term voltage
stability and FIDVR, is investigated based on various analysis methods, i.e., time-
domain simulation, CCT, KL divergence measure, and VRI. The results show that
without the LVRT capability of the PV system, the system is short-term voltage
unstable. Even the blocking mode leads to short-term voltage instability. Only the
LVRT and DVS capability of the PV system helps avoiding instability and improves
FIDVR. Considering the delayed voltage recovery, a fast active current recovery rate
of 1000 %/s shows the best behaviour. In case of local constant voltage control,
not only FIDVR, but also the post-fault voltage in steady state is substantially
enhanced. Based on the control at plant level and its influence on voltage dynamics
up to 20 s, the best performance can be achieved using plant-level constant volt-
age and local coordinated reactive power/voltage control. The considered analysis
methods reveal advantages and drawbacks. The numerical time-domain simulations
show accurate results and allow a detailed analysis of the dynamic performance of
the power system. The calculation of the CCT is very helpful to determine the sta-
bility margin of the system. However, a smaller step size would lead to somewhat
more accurate values of the CCT. Furthermore, it is shown that the application of
the KL divergence measure is limited for the analysis of FIDVR. However, the new
metric, namely the VRI, provides useful information about the voltage recovery and
helps to compare different short-term voltage controls more systematically. In gen-
eral, with the adequate control of the PV system, the results show the significant
improvement of power system dynamic performance.

Future work: Future work considers variations of the load model that is used
in the one-load infinite-bus test system. In particular, the mechanical torque of
the induction motor load can be modified to, e.g., quadratic, or a combination of
different torque characteristics to form composite mechanical loads. Another aspect
that is of interest for future research, is the investigation of additional operating
points in the test system considering a higher PV penetration level. Moreover, the
studied contingency could be varied in terms of fault location, fault impedance, fault
type or fault duration. Other stability criteria, such as long-term voltage stability
or small-signal stability, are also of interest.
81

Stability analysis of a multi-machine system


5
5.1 Introduction
The content of this chapter is partly based on the following publication:
G. Lammert, D. Premm, L. D. Pabón Ospina, J. C. Boemer, M. Braun,
and T. Van Cutsem, “Control of Photovoltaic Systems for Enhanced
Short-Term Voltage Stability and Recovery”, IEEE Transactions on En-
ergy Conversion, vol. 34, no. 1, pp. 243–254, 2019.
Motivation: The impact of traditional technologies on power system stability in a
multi-machine system is well known. However, the interactions of PV systems with
synchronous generators and induction motors in a multi-machine voltage stability
test system, are not yet fully understood.

Contributions: This chapter provides a comprehensive stability analysis of a


multi-machine system, namely the Nordic test system, considering different PV
system control modes. The impact of PV systems not only on short-term voltage
stability and FIDVR, but also transient stability and frequency dynamics, is inves-
tigated in detail. The dynamic performance of the Nordic test system is analysed
using different methods, i.e., time-domain simulation, CCT, KL divergence measure
and VRI. Finally, this chapter gives recommendations for the adequate control of
PV systems in order to improve the overall dynamic behaviour of the power system.

Structure: The remainder of this chapter is organised as follows. The multi-


machine test system that is considered for the stability analysis, is presented in
Section 5.2. This section includes details about the load model, the substitution
of synchronous generators, the PV system model, the operating point as well as
the considered contingencies. Section 5.3 presents the case studies, specifically the
results of the impact analysis of different PV system control modes on short-term
voltage stability and FIDVR as well as transient and frequency stability. Finally,
the conclusions and directions for future work are given in Section 5.4.
82 Chapter 5. Stability analysis of a multi-machine system

5.2 Test system


In this section, an IEEE multi-machine voltage stability test system, namely the
Nordic test system, is introduced. The Nordic test system is a variation of the
so-called Nordic32 system, which was created to illustrate the voltage collapse in
Sweden that happened in 1983 [107]. The proposed test system was initially devel-
oped to investigate long-term dynamics. Nevertheless, the Nordic test system is not
limited to long-term phenomena. Its applicability to short-term dynamics, such as
short-term voltage stability and transient stability as well as small-signal stability
is recognised [108]. The considered test system is based on report [98] and adapted
to investigate specifically short-term voltage stability and FIDVR. Moreover, the
influence of PV systems on transient and frequency stability is studied briefly. The
details of the Nordic test system, i.e., the overview of the single-line diagram, the
load model, the replacement of synchronous generators, the modelling approach of
the large-scale PV systems, the pre-disturbance operating point, and the considered
contingencies, are described in the sequel.

5.2.1 Overview
The single-line diagram of the Nordic test system is depicted in Fig. 5.1. The system
consists of four areas, namely:
• “North” with hydro generation and some load,
• “Central” with thermal generation and the load center,
• “Equiv” connected to “North”, which includes a simple equivalent of an ex-
ternal system, and
• “South” with thermal generation, rather loosely connected to the rest of the
system.
The test network is composed by 32 transmission and subtransmission buses with
nominal voltages of 400 kV, 220 kV and 130 kV. In addition, 22 buses are at distri-
bution level with a voltage of 20 kV. The distribution systems are represented by
aggregated loads, as described in detail in Section 5.2.2. It should be noted that
the distribution system can be represented explicitly, if, for instance, interactions
between the transmission and distribution system are studied, and dispersed gen-
eration is considered [109, 110]. The nominal frequency of the system is 50 Hz.
The frequency is controlled through the speed governors of the hydro generators
in the “North” and “Equiv” areas only. In order to keep the test system as sim-
ple as possible, while retaining the essential of voltage instability phenomena, the
following modelling assumptions are made. It should be noted that the details of
the load and the PV system model are of particular interest, and therefore given
in Sections 5.2.2 and 5.2.4, respectively. Additional information about the general
system description can be found in report [98].
5.2. Test system 83

g19 g9 g2

4071 4011 1011 1013

71 11 13
12
EQUIV. 4012 1012 1014

4072
g10 g1 g3
g5 g11
72 g20 NORTH
400 kV
1021 1022 4022 4021
220 kV
g12 130 kV
22  
g4     
2032 2031 4031 4032

32 31
g8
42
PV
4042
4041
CENTRAL
g14
6&
41 g7 PV PV
g13 4044 4043 4046
4
3
4061 43
1043 g6 46
PV
PV PV
1044
6161
1042

g17 PV
1041 1045 2 g15 g15b
62 PV PV
47
PV 1 5 g16 g16b

4062 4045 5151


4047
4063 4051

g18
g18 g18b 63
63 SOUTH
Figure 5.1: Nordic test system [98]
84 Chapter 5. Stability analysis of a multi-machine system

Synchronous machines: The synchronous machines are represented by a stan-


dard model with three rotor windings for the salient pole machines of hydro power
plants, and four rotor windings for the round-rotor machines of thermal power
plants. The model of the hydro power plant includes the exciter, the over exci-
tation limiter, the automatic voltage regulator and the power system stabiliser.
Moreover, the hydro power plants are involved in frequency control. Therefore, the
speed governor and the hydro turbine are also represented. On the other hand, the
thermal power plants are modelled with exciter, automatic voltage regulator, over
excitation limiter as well as power system stabiliser. However, the thermal power
plants do not control the frequency. Hence, no speed governor and no hydro turbine
are represented. It should be noted that the synchronous machine 13 is a condenser
and modelled as a salient pole machine. Furthermore, the “Equiv” area includes
the synchronous generators 19 and 20, which are equivalent generators, with a large
participation in primary frequency control. Therefore, the synchronous machines
13, 19 and 20 do not include power system stabilisers. All dynamic models and the
corresponding parameter values are detailed in report [98].

Transmission and subtransmission lines: The transmission and subtransmis-


sion lines are represented by π-equivalents. Five transmission lines are equipped
with series compensation. The corresponding parameter values of the transmission
and subtransmission lines are presented in detail in report [98].

Step-up and step-down transformers: The step-up transformers of all gen-


erators and the step-down transformers of all loads are represented explicitly. In
the initial power flow computation, the step-down transformers have their ratios
adjusted according to Table D.1 of Appendix D, in order to receive the distribu-
tion bus voltage of 1.0 pu. It should be noted that all distribution transformers are
equipped with load tap changers in order to keep the distribution voltage within the
deadband. However, due to the power system phenomena of interest, i.e., short-
term voltage stability and FIDVR, the dynamics of the load tap changer can be
neglected because they have intentional delays of about 30 s, which is outside of
the relevant time scale. Therefore, the transformer tap ratio can be assumed as fix.
Additional parameter values of the step-up and step-down transformers are detailed
in report [98].

Shunt capacitors and inductors: Shunt capacitors and inductors are connected
to the system in order to compensate the reactive power and control the voltage.
Compared to report [98], the values of the shunt capacitors and inductors need to
be adjusted due to the substitution of synchronous generators, which is presented
in Section 5.2.3.
5.2. Test system 85

5.2.2 Load modelling


This section describes the considered load model, which is a key element in power
system voltage stability studies. Similar to the considered load model in the one-
load infinite-bus system, as outlined in Section 4.2.2, the load is divided into a static
and dynamic part, which is different to the modelling approach in report [98]. The
static part is represented by the well-known exponential model. The parameter
values of the exponential load model are shown in Table D.2 of Appendix D. The
dynamic part consists of two aggregated induction motor models. One represents
a large equivalent industrial motor, the other a small equivalent industrial motor.
The parameter values of the small and large induction motor models are shown in
Tables D.3 and D.4 of Appendix D, respectively. Further details to the load model
are outlined in Section 4.2.2. Similar to that section, the load is composed by a 70 %
static and 30 % dynamic part. Note that the load model in the “Equiv” area of the
test system in Fig. 5.1 is not modified and represented according to report [98].

5.2.3 Substitution of synchronous generators


The initial configuration of the Nordic test system is without PV systems, as detailed
in report [98], and serves as a reference. To investigate the impact of PV plants on
power system stability, specifically short-term voltage stability and FIDVR, syn-
chronous generators are replaced. In particular, synchronous generators 6 and 7
are replaced by 5, generator 14 by 3, and generator 17 by 2 PV plants, as detailed
in Tables 5.1 and 5.2. In total, four synchronous generators are substituted by ten
large-scale PV systems that have in sum a similar nominal power. This specific
substitution is selected to show the impact of PV system control on loads that are
electrical close. The disconnected synchronous generators as well as the connected
PV plants are highlighted in grey in the test system shown in Fig. 5.1.
Table 5.1: Disconnected synchronous generators

Number of Number of Connected Size of each


Area
substitution synchronous generators to bus synchronous generator
1 Central 2 (generator 6 and 7) 1042, 1043 400 MVA, 200 MVA
2 Central 1 (generator 14) 4042 700 MVA
3 South 1 (generator 17) 4062 600 MVA

Table 5.2: Connected photovoltaic systems

Number of Number of Connected Size of each


Area
substitution photovoltaic systems to bus photovoltaic system
1 Central 5 1041–1045 130 MVA
2 Central 3 4042, 4043, 4046 240 MVA
3 South 2 4061, 4062 300 MVA
86 Chapter 5. Stability analysis of a multi-machine system

5.2.4 Modelling of photovoltaic systems


The PV systems are modelled according to Chapter 3. The model is generic and
based on the WECC specifications [82, 83]. Furthermore, the model represents a
large-scale PV plant connected to the high or extra high voltage level. An overview
of the PV system model as well as its implementation is given in Section 3.2. The
control modes that are investigated considering the multi-machine system are de-
tailed in Section 3.3. Typical parameters that represent the average behaviour of
presently required control modes and settings according to international guidelines
and grid codes, are determined in collaboration with a manufacturer. The con-
sidered parameter values of the PV system model are given in Tables B.1 to B.3
of Appendix B. The single-converter equivalent system of the large-scale PV plant
is depicted in Fig. 5.2. The PV system equivalent is connected to the step-up
transformer equivalent and via the collector system equivalent to the substation
transformer. The substation transformer is connected to the rest of the power sys-
tem via the interconnection transmission line. It should be noted that within the
PV plant, no shunt compensation is considered. The measurement and control ac-
tions are carried out at the PCC. Parameter values of the single-converter equivalent
system are based on reports [98, 106] and can be found in Table D.5 of Appendix D.
The considered equivalent system of the large-scale PV plant illustrated in Fig. 5.2,
is similar to the system used for the stability analysis of the one-load infinite-bus
system presented in Fig. 4.4. The entire single-converter equivalent system of the
large-scale PV plant is connected to the Nordic test system in Fig 5.1, according to
the substitutions outlined in Table 5.2.
Interconnection Substation Collector Step-up transformer Photovoltaic
transmission line transformer system equivalent equivalent system equivalent

PV

130 kV or 130 kV or 20 kV 20 kV 0.6 kV


400 kV (PCC) 400 kV

Figure 5.2: Single-converter equivalent of large-scale photovoltaic system


5.2. Test system 87

5.2.5 Operating point


Compared to operating point A, in which the system cannot withstand some N − 1
contingencies, the system is made secure by some modifications, as detailed in
report [98], which lead to operating point B. The modifications are twofold, namely:

• In parallel with the synchronous generator 16 and its step-up transformer, an


identical generator and step-up transformer are connected. The new generator
is named g16b, as shown in Fig. 5.1. Generator 16b produces the same active
power under the same terminal voltage.

• The synchronous generators 15 and 18 are replaced by two identical generators


with half of the nominal apparent power, turbine power and production. The
new generators are named g15b and g18b, as shown in Fig. 5.1. Each generator
is connected to the rest of the system by a step-up transformer with half of
the nominal apparent power. Both generators have the same terminal voltage
as the generator they replace.

Further details to operating point B and the necessary modifications can be found
in report [98]. Considering this operating point, the demand of the load center in
the “Central” area is mainly covered by the generation in the “North” area, which
leads to large transfers. An overview of the load and generation in the Nordic
test system with respect to the area, is shown in Table 5.3. Please note that
the consumer-oriented sign convention is used in this work, i.e., consumed active
power is positive and generated active power is negative. In particular, the load
consumption is detailed in Table D.6 of Appendix D. Furthermore, all PV systems
are operated at unity power factor and their generation is detailed in Table D.7
of Appendix D. In order to match operating point B as described in report [98],
a new shunt is connected to bus 1042 and the other existing shunts are adjusted,
as shown in Table D.8 of Appendix D. Considering the load and generation shown
in Table 5.3, the PV penetration level is calculated according to Eq. (4.8), which
leads to 15 % for the entire system. For the “Central” area in particular, a PV
penetration level of 41 % is reached.

Table 5.3: Load and generation of Nordic test system

Generation
 [MW]  Load
 [MW] 
Area PV pen [%]
PSG PPV Pexp Psim + Plim
North −4628.5 0 826 354 0
Central −1680 −1170 4333 1857 41
South −1050 −540 973 417 34
Equiv −2437.4 0 2300 0 0
Total −9795.9 −1710 8432 2628 15
88 Chapter 5. Stability analysis of a multi-machine system

5.2.6 Contingencies
Two contingencies, named fault 1 and fault 2, are considered to investigate the im-
pact of PV system control on power system stability, specifically short-term voltage
stability and delayed voltage recovery. The contingencies can be characterised by
the fault location, fault impedance, fault type, and fault duration [92], as detailed
in the sequel.

Fault location: Fault 1 is located on transmission line 4032–4044, between the


“North” and the “Central” area and close to bus 4044, i.e., 80 % of the line length,
as shown in Fig. 5.1. Fault 2 takes place on one of the transmission lines 4045–
4051, within the “Central” area and close to bus 4051, i.e., 10 % of the line length,
as depicted in the same figure. The fault locations highly influence PV systems,
induction motors and synchronous generators.

Fault impedance: Fault 1 and fault 2 have a fault impedance of zero, which
corresponds to a solid short circuit. This impedance leads to a severe voltage
dip that highly influences the dynamic behaviour of PV systems. Moreover, the
resulting voltage dip affects the operation of small and large induction motor loads
as well as synchronous generators.

Fault type: Fault 1 and fault 2 are three-phase short circuits. This type of fault
leads to a severe voltage sag that, again, highly impacts the operation of large-
scale PV systems, small and large induction motor loads as well as synchronous
generators.

Fault duration: Fault 1 and fault 2 have a fault duration of 0.1 s, respectively. The
faults are cleared by opening the corresponding transmission line, which remains
open. The fault duration is typical for protection systems according to international
grid codes. Moreover, the large-scale PV systems need to stay connected to the grid
during the short circuit and ride through the fault. This requirement is defined by
the LVRT characteristic shown in Fig. 3.8.
5.3. Case studies 89

5.3 Case studies


This section analyses the impact of different PV system control modes on power
system stability considering the Nordic test system depicted in Fig. 5.1. The relevant
cases are shown in Table 3.4 and studied based on the analysis methods described
in Chapter 2, i.e., time-domain simulation, CCT, KL divergence measure, and VRI.
The results of the case studies are presented in the sequel.

5.3.1 Without photovoltaic systems


This section studies the initial configuration of the Nordic test system, which is
without PV systems. This configuration is the base case and serves as a reference.
The time-domain results of the voltage at the load buses, in response to fault 1 and
fault 2, are given in Fig. 5.3. The corresponding metrics are shown in Table 5.4.
The CCT determines the stability margin, whereas the KL divergence measure and
the VRI analyse the voltage recovery. Considering fault 2, for instance, leads to
tCCT2 , K2 and V2 values of 0.12 s, 0.24 and 0.87, respectively.

1.0

0.8
Voltage [pu]

0.6

0.4

WECC criterion
0.2

V4 (fault 1)
0.0 V2 (fault 2)

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure 5.3: Base case without photovoltaic systems


 Voltage at bus 2 and bus 4 (load buses)

Table 5.4: Metrics of base case without photovoltaic systems

Fault 1 Fault 2
Case
tCCT1 [s] K4 [–] V4 [–] tCCT2 [s] K2 [–] V2 [–]
Base 0.15 1.47 0.74 0.12 0.24 0.87
90 Chapter 5. Stability analysis of a multi-machine system

5.3.2 Low voltage ride-through and dynamic voltage


support
This section analyses the impact of the LVRT and DVS capability of PV systems on
power system stability. The corresponding cases aim at comparing the disconnection
of large-scale PV systems (A1) with the blocking mode (A2) as well as with the
LVRT and DVS capability (A3), as shown in Table 3.4. The aforementioned control
modes are described in detail in Section 3.3.1.
The time-domain results in response to fault 2 are gathered in Figs. 5.4 to 5.11.
Due to the disconnection of PV systems (A1), the small and large induction motors
connected to buses 1–5 and 51 stall because their rotor slips exceed the unstable
equilibrium values, as revealed by the slips shown in Figs. 5.8 and 5.9. This causes
short-term voltage instability, as shown by the voltage at the PCC and the load bus
in Figs. 5.4 and 5.5, respectively. As a result of the voltage collapse, synchronous
generators 16 and 16b lose synchronism, i.e., become transient unstable, which is
shown by the relative rotor angle in Figs. 5.10 and 5.11. As a result of the repeated
pole slips for t > 2.0 s, the voltage oscillates, as shown in Figs. 5.4 and 5.5. In fact,
the slips of the small and large induction motors connected to bus 2 and the relative
rotor angle of synchronous generator 16 show that short-term voltage instability and
transient instability go hand in hand and appear in the same time scale. If the PV
systems remain connected to the grid but reduce their active and reactive current to
zero during the fault-on period (A2), short-term voltage instability still takes place,
as shown by the load bus voltage in Fig. 5.5. After the fault is cleared at t = 0.1 s,
the voltage is locked at a low level. Again, the cause for the voltage collapse is the
stalling of the large induction motor load, as depicted in Fig. 5.9. However, due to
the LVRT capability of the PV systems in this case (A2), the dynamic performance
of the overall power system is improved in so far, as less induction motors stall
and no transient instability takes place. For both cases (A1 and A2), the WECC
criterion, as defined in Section 2.5.1, is significantly violated, which is illustrated in
Fig. 5.5. It should be noted that due to the depressed voltage caused by the blocking
mode (A2), PV systems may disconnect according to the LVRT requirements [7],
as shown in Fig. 3.8. However, this is not considered in this investigation. The
best dynamic performance is achieved with the LVRT and DVS capability of PV
systems (A3), as shown by the load bus voltage in Fig. 5.5. In this case, PV systems
ride through the fault and feed in an additional capacitive reactive current during
the fault-on period, as shown by the reactive current in Fig. 5.7. As a result,
short-term voltage instability can be avoided and FIDVR is significantly improved.
Moreover, using the LVRT and DVS capability (A3), the WECC voltage criterion
is not violated anymore, as shown in Fig. 5.5. Further time-domain results that
show the impact of the LVRT and DVS capability of PV systems on the dynamic
performance of the Nordic test system, are given in Figs. D.1 to D.6 of Appendix D.
5.3. Case studies 91

1.0

0.8
Voltage [pu]

0.6

0.4

0.2 LVRT characteristic


A1
A2
0.0 A3

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure 5.4: Impact of low voltage ride-through and dynamic voltage support (fault 2)
 Voltage at bus 1042 (point of common coupling)

1.0

0.8
Voltage [pu]

0.6

0.4

WECC criterion
0.2
A1
A2
0.0 A3

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure 5.5: Impact of low voltage ride-through and dynamic voltage support (fault 2)
 Voltage at bus 2 (load bus)
92 Chapter 5. Stability analysis of a multi-machine system

0.0
A1
−0.2 A2
A3
Active current [pu]

−0.4

−0.6

−0.8

−1.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure 5.6: Impact of low voltage ride-through and dynamic voltage support (fault 2)
 Active current of photovoltaic system connected to bus 1042

0.0

−0.2
Reactive current [pu]

−0.4

−0.6

−0.8
A1
A2
−1.0 A3

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure 5.7: Impact of low voltage ride-through and dynamic voltage support (fault 2)
 Reactive current of photovoltaic system connected to bus 1042
5.3. Case studies 93

1.0

0.8

0.6
Slip [pu]

A1
0.4 A2
A3

0.2

0.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure 5.8: Impact of low voltage ride-through and dynamic voltage support (fault 2)
 Slip of small induction motor load connected to bus 2

1.0

0.8

0.6
Slip [pu]

A1
0.4 A2
A3

0.2

0.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure 5.9: Impact of low voltage ride-through and dynamic voltage support (fault 2)
 Slip of large induction motor load connected to bus 2
94 Chapter 5. Stability analysis of a multi-machine system

200.0

100.0
Rotor angle [◦ ]

0.0

−100.0
A1
A2
A3
−200.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Time [s]

Figure 5.10: Impact of low voltage ride-through and dynamic voltage support (fault 2)
 Rotor angle of synchronous generator 16

1.4 A1
A2
1.2 A3

t = 0.5 s
Active power [pu]

1.0

0.8

0.6

0.4

0.2 t=0s
t = 0.1 s
0.0
−40.0 −20.0 0.0 20.0 40.0 60.0 80.0 100.0 120.0 140.0
Rotor angle [◦ ]

Figure 5.11: Impact of low voltage ride-through and dynamic voltage support (fault 2)
 Active power as function of rotor angle of synchronous generator 16
5.3. Case studies 95

The stability margin is determined using the CCT and the results are shown in
Table 5.5. The best dynamic performance is highlighted in grey. Considering fault 2,
for instance, the CCT is significantly increased from 0.05 s to 0.12 s comparing
the disconnection of PV systems (A1) with the LVRT and DVS capability (A3).
The blocking mode of PV systems (A2) leads to a CCT of 0.09 s, which is higher
compared to the disconnection case (A1). The base case without PV systems leads
to the same stability margin, i.e., a CCT of 0.12 s. The calculation of the CCT for
fault 1 reveals the same tendency. The best dynamic performance is achieved using
the LVRT and DVS capability of PV plants (A3), which leads to a CCT of 0.12 s.
The worst CCT is calculated if PV systems disconnect (A1), which yields a value
of 0.02 s.
The KL divergence measure is applied to analyse the voltage recovery and give
information about the effectiveness of short-term voltage controls of PV systems.
The results are shown in Table 5.5 and the best dynamic performance is highlighted
in grey. In the same manner, the KL divergence measure K2 that is calculated
from the voltage at the load bus (bus 2), shows the best performance using the
LVRT and DVS capability of PV systems (A3), which yields a value of 0.84. The
lowest performance is reached if PV systems do not have any LVRT capability (A1),
with a K2 value of 159.03. A better voltage recovery is obtained considering the
blocking mode (A2), with a KL divergence measure of 36.12. The base case without
PV systems yields a value of 0.24 due to the higher voltage support capability of
synchronous generators during the fault-on period. The same trend, as described
before, can be observed for fault 1.
Finally, the impact of the LVRT and DVS capability on the phenomenon of
FIDVR is analysed using the VRI and the results are presented in Table 5.5. The
best dynamic performance is highlighted in grey. The same tendency as for the
other metrics is observed. With respect to fault 2, for instance, the disconnection
of PV systems (A1) results in an unacceptable V2 value of −0.28, considering the
voltage recovery at bus 2. On the other hand, the best performance is achieved
using the LVRT and DVS capability of PV systems (A3), with a VRI of 0.64. If
only the LVRT capability of PV systems is incorporated (A2), the calculation of
the VRI yields −0.02. The base case leads to a VRI of 0.87 due to the higher
voltage support capability of synchronous generators. Compared to fault 2, the
same tendency of the calculated metrics can be observed for fault 1.

Table 5.5: Metrics of low voltage ride-through and dynamic voltage support

Fault 1 Fault 2
Case
tCCT1 [s] K4 [–] V4 [–] tCCT2 [s] K2 [–] V2 [–]
A1 0.02 55.71 −0.04 0.05 159.03 −0.28
A2 0.05 31.53 −0.01 0.09 36.12 −0.02
A3 0.12 2.82 0.66 0.12 0.84 0.64
Base 0.15 1.47 0.74 0.12 0.24 0.87
96 Chapter 5. Stability analysis of a multi-machine system

5.3.3 Active current recovery rate


This section studies the impact of the active current recovery rate of PV systems on
the delayed voltage recovery. The corresponding cases differ by the ramp rate of the
active current. The active current ramp rate is investigated for the values 10 %/s
(B1), 20 %/s (B2), 100 %/s (A3) and 1000 %/s (B3), as presented in Table 3.4.
The details to the aforementioned control modes can be found in Section 3.3.2.
The time-domain results in response to fault 1 are illustrated in Figs. 5.12 to 5.16.
It should be noted that due to the specific control modes considered in this section,
the simulation time is extended until 14.0 s in order to capture the relevant dynamic
phenomena. The very slow active current recovery rate of 10 %/s (B1) leads to
significantly delayed voltage recoveries with voltages as low as 0.83 pu at 1.2 s, as
shown by the load bus voltage in Fig. 5.13. Higher ramp rates, such as 1000 %/s
(B3), accelerate the voltage recovery, leading to 0.93 pu voltage at 1.2 s, as depicted
in the same figure. It is noteworthy that due to the low voltages caused by the slow
active current ramping of 10 %/s (B1) or 20 %/s (B2), the PV systems inject an
additional reactive current at 1.2 s and 2.5 s, to support the voltage, even though
the fault has been already cleared. This phenomenon is illustrated by the reactive
current evolution of the PV system connected to bus 1044, as shown in Fig. 5.15.
Moreover, the active current ramp rates of 10 %/s (B1) and 20 %/s (B2) violate the
WECC voltage criterion, as shown in Fig. 5.13. In general, considering the overall
voltage dynamics, it can be concluded that the higher the active current ramp rate,
the better the voltage recovery. It is important to mention that due to the opening
of transmission line 4032–4044 caused by the fault, the impedance between the
“North” and the “Central” area increases. This leads to a voltage decrease at the
PCC of the PV system connected to bus 1044. In order to keep the active power
injection of the PV system constant, the voltage decrease is encountered with an
active current increase, as presented in Fig. 5.14. In particular, the pre-fault active
current injection is −0.88 pu for t < 0 s, whereas the post-fault active current
injection in steady state is −0.91 pu for t > 14.0 s, as illustrated in the same figure.
Finally, it should be noted that for all PV system control modes investigated in
this section, no short-term voltage instability is present, as shown by the voltage
at the PCC and the load bus voltage in Figs. 5.12 and 5.13, respectively. Further
results of the time-domain simulations that show the impact of the active current
recovery rate of PV systems on the dynamic performance of the Nordic test system,
are given in Figs. D.7 to D.10 of Appendix D.
5.3. Case studies 97

1.0

0.8
LVRT characteristic
Voltage [pu]

0.6
1.05
1.00
0.4
0.95
0.90 B1
0.2
0.85 B2
0.0 0.5 1.0 1.5 2.0 2.5 3.0 A3
0.0 B3

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0


Time [s]

Figure 5.12: Impact of active current recovery rate (fault 1)


 Voltage at bus 1044 (point of common coupling)

1.0

0.8
WECC criterion
Voltage [pu]

0.6 1.05
1.00
0.4 0.95
0.90
0.2 0.85 B1
B2
0.0 0.5 1.0 1.5 2.0 2.5 3.0 A3
0.0 B3

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0


Time [s]

Figure 5.13: Impact of active current recovery rate (fault 1)


 Voltage at bus 4 (load bus)
98 Chapter 5. Stability analysis of a multi-machine system

0.0 B1
B2
A3
−0.2 B3
Active current [pu]

−0.4

−0.6

−0.8

−1.0

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0


Time [s]

Figure 5.14: Impact of active current recovery rate (fault 1)


 Active current of photovoltaic system connected to bus 1044

0.0

−0.2
Reactive current [pu]

−0.4
0.00
−0.05
−0.6
−0.10
−0.15
−0.8 −0.20 B1
B2
0.0 0.5 1.0 1.5 2.0 2.5 3.0 A3
−1.0 B3

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0


Time [s]

Figure 5.15: Impact of active current recovery rate (fault 1)


 Reactive current of photovoltaic system connected to bus 1044
5.3. Case studies 99

Apart from the voltage dynamics discussed before, the active current recovery
rate of PV systems also influences frequency dynamics. Therefore, the frequency of
the center of inertia of the Nordic test system is determined. This frequency is a
weighted sum of all synchronous machine speeds in the system. The frequency of
the center of inertia is depicted in Fig. 5.16. At 0.6 s, the frequency rises due to the
load sensitivity, i.e., a lower voltage leads to a lower active power consumption, as
shown in Eq. (4.1). However, at 2.2 s, the active current ramp rates of 10 %/s (B1)
and 20 %/s (B2) lead to a frequency nadir of 49.8 Hz and 49.85 Hz, respectively.
On the other hand, considering an active current ramp rate of 1000 %/s (B3),
the frequency rises to a value higher than 50.2 Hz at 0.8 s and 1.8 s, respectively.
According to old German grid code requirements, PV systems would disconnect
at this frequency and cause a further frequency drop in the system, which is well
known as the 50.2 Hz problem [12]. Therefore, the medium active current recovery
rate of 100 %/s (A3) shows the best frequency performance. If both dynamics are
taken into account, i.e., voltage and frequency dynamics, as depicted in Figs. 5.13
and 5.16, respectively, the best performance is achieved with the medium active
current recovery rate of 100 %/s (A3).

50.2

50.1
Frequency [Hz]

50.0

49.9 B1
B2
A3
B3
49.8
0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0
Time [s]

Figure 5.16: Impact of active current recovery rate (fault 1)


 Frequency of center of inertia
100 Chapter 5. Stability analysis of a multi-machine system

The CCT is calculated in order to determine the stability margin of the system.
The results are presented in Table 5.6 and the best dynamic performance is high-
lighted in grey. With respect to fault 1, for instance, the CCT is not influenced
by the different ramp rates, leading to 0.12 s for all cases. The base case leads to
a CCT of 0.15 s. Considering fault 2, the same tendency is observed as for the
time-domain results. The best performance is achieved with a medium and fast
active current recovery rate of 100 %/s (A3) and 1000 %/s (B3), which lead to a
CCT of 0.12 s, respectively. Slower ramp rates, such as 10 %/s (B1) and 20 %/s
(B2), lead to a smaller CCT of 0.11 s each.
The voltage recovery is analysed using the KL divergence measure and the results
are given in Table 5.6. The best performance is highlighted in grey. Again, the same
trend is observed as for the CCT and the time-domain results. The fastest active
current ramp rate of 1000 %/s (B3) yields the best KL divergence measure K4 that
is calculated from the load bus voltage (bus 4), with a value of 2.10 for fault 1. The
lowest performance is reached if PV systems inject active current with a ramp rate
of 10 %/s (B1), which results in a K4 value of 7.19. Furthermore, the KL divergence
measure for the slow active current recovery of 20 %/s (B2) is higher, with a value
of 5.65, compared to the medium active current recovery of 100 %/s (A3), with
a value of 2.82. Hence, the latter results in a better quality of voltage recovery.
The base case yields a value of 1.47 due to the higher voltage support capability of
synchronous generators. The same trend is observed for fault 2.
The last metric that is used to analyse the impact of the active current recovery
rate on FIDVR, is the VRI and the results are shown in Table 5.6. The best
dynamic performance is highlighted in grey. Generally speaking, a similar tendency
is observed as for the aforementioned metrics. In the same manner, the fastest
active current recovery rate of 1000 %/s (B3) results in a high V4 value of 0.68,
considering the voltage recovery at bus 4 for fault 1. The worst VRI is calculated
considering the very slow active current recovery rate of 10 %/s (B1), which yields
a V4 value of 0.22. If only a slow (B2) or medium (A3) active current recovery rate
of the PV system is incorporated, the calculation of the VRI yields 0.29 or 0.66,
respectively. Again, the same tendency is observed for fault 2.

Table 5.6: Metrics of active current recovery rate

Fault 1 Fault 2
Case
tCCT1 [s] K4 [–] V4 [–] tCCT2 [s] K2 [–] V2 [–]
B1 0.12 7.19 0.22 0.11 1.22 0.58
B2 0.12 5.65 0.29 0.11 1.07 0.60
A3 0.12 2.82 0.66 0.12 0.84 0.64
B3 0.12 2.10 0.68 0.12 0.82 0.66
Base 0.15 1.47 0.74 0.12 0.24 0.87
5.3. Case studies 101

5.3.4 Local voltage control


This section investigates the impact of the local voltage control of PV systems on the
dynamic performance of the power system. The corresponding cases aim at com-
paring different local voltage control strategies, i.e., local constant reactive power
control (A3), local constant voltage control (C1), and local coordinated reactive
power/voltage control (C2), as shown in Table 3.4. Further details to the specific
control modes can be found in Section 3.3.3.
The time-domain results in response to fault 1 are shown in Figs. 5.17 to 5.22.
Similar to the investigation of the active current recovery rate in Section 5.3.3, the
simulation time is extended until 14.0 s in order to capture the relevant dynamic
phenomena that are influenced by the different local voltage control modes of PV
systems. In case of local constant reactive power control (A3) of PV systems, the
reactive power in steady state is kept at the value given by the initial power flow
computation, namely zero, which is shown in Fig. 5.20. This control method results
in a post-fault steady-state voltage deviation, as depicted by the PCC voltage in
Fig. 5.17. The pre-fault voltage at the PCC is 1.01 pu for t < 0 s, whereas the
post-fault voltage in steady state is 0.98 pu for t > 14.0 s, as illustrated in the same
figure. On the other hand, in case of local constant voltage control (C1), the PV
system acts fast and tries to restore its pre-fault voltage by injecting reactive current
into the grid, as depicted in Fig. 5.20. Therefore, the active current is reduced, as
shown in Fig. 5.19. However, due to the rather stiff grid behaviour, restoring grid
voltage is impossible and the system collapses, as shown by the load bus voltage
in Fig. 5.18 and the small and large induction motor slips in Figs. 5.21 and 5.22,
respectively. The Nordic test system collapses owing to the lack of active power
produced by the PV systems. The third and last control mode in this section, is the
local coordinated reactive power/voltage control (C2). This control mode combines
local constant reactive power (A3) with local constant voltage control (C1). The
local coordinated reactive power/voltage control (C2) allows a fast voltage control
to restore the voltage quickly after the fault, as shown by the PCC voltage in
Fig. 5.17, followed by a slow reactive power control that brings the PV system back
to its reactive power set point in steady state, which is depicted by the reactive
current in Fig. 5.20. In other words, for dynamics in the order of a few seconds, the
local coordinated reactive power/voltage control (C2) is similar to local constant
voltage control (C1), whereas for dynamics in the order of several seconds, the
control is related to local constant reactive power control (A3). This control mode
(C2) shows the best overall dynamic performance as it results in a less oscillatory
voltage evolution compared to the local constant reactive power control (A3), as
depicted in Fig. 5.17. Although counterintuitive, the time-domain results show
that a high reactive power injection is not always the best to counteract a voltage
collapse. This is in fact a matter of relative effects of active and reactive current
injections. Further time-domain results that show the impact of local voltage control
strategies of PV systems on the dynamic performance of the Nordic test system,
are given in Figs. D.11 to D.16 of Appendix D.
102 Chapter 5. Stability analysis of a multi-machine system

1.0

0.8
Voltage [pu]

0.6

0.4

0.2 LVRT characteristic


A3
C1
0.0 C2

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0


Time [s]

Figure 5.17: Impact of local voltage control (fault 1)


 Voltage at bus 1044 (point of common coupling)

1.0

0.8
Voltage [pu]

0.6

0.4

WECC criterion
0.2
A3
C1
0.0 C2

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0


Time [s]

Figure 5.18: Impact of local voltage control (fault 1)


 Voltage at bus 4 (load bus)
5.3. Case studies 103

0.0
A3
−0.2 C1
C2
Active current [pu]

−0.4

−0.6

−0.8

−1.0

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0


Time [s]

Figure 5.19: Impact of local voltage control (fault 1)


 Active current of photovoltaic system connected to bus 1044

0.4 A3
C1
0.2 C2
Reactive current [pu]

0.0

−0.2

−0.4

−0.6

−0.8

−1.0
0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0
Time [s]

Figure 5.20: Impact of local voltage control (fault 1)


 Reactive current of photovoltaic system connected to bus 1044
104 Chapter 5. Stability analysis of a multi-machine system

1.0

0.8

0.6
Slip [pu]

A3
0.4 C1
C2

0.2

0.0

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0


Time [s]

Figure 5.21: Impact of local voltage control (fault 1)


 Slip of small induction motor load connected to bus 4

1.0

0.8

0.6
Slip [pu]

A3
0.4 C1
C2

0.2

0.0

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0


Time [s]

Figure 5.22: Impact of local voltage control (fault 1)


 Slip of large induction motor load connected to bus 4
5.3. Case studies 105

The stability margin of the system is determined using the CCT and the results
are shown in Table 5.7. The best dynamic performance is highlighted in grey.
Compared to the time-domain simulations discussed before, a different trend for
the CCT can be observed owing to the relative effects of active and reactive current
injections of PV systems. Using local constant reactive power control (A3), leads to
the best performance with a CCT of 0.12 s for fault 1. If, for the same fault, local
constant voltage control (C1) or local coordinated reactive power/voltage control
(C2) is applied, the CCT results in a value of 0.11 s, respectively. The base case
leads to a CCT of 0.15 s due to the higher voltage support capability of synchronous
generators. The same tendency is observed for fault 2.
The KL divergence measure is applied in order to assess the delayed voltage
recovery. The results are given in Table 5.7 and the best performance is highlighted
in grey. The same trend is observed as for the CCT. Local constant reactive power
control (A3) results in the best KL divergence measure K4 that is calculated from
the load bus voltage (bus 4), with a value of 2.82 for fault 1. The lowest performance
is reached, if local constant voltage control (C1) is applied, which results in a K4
value of 71.99 due to the system collapse. The combination of both local controls,
i.e., local coordinated reactive power/voltage control (C2), leads to a KL divergence
measure K4 of 3.78. Compared to the time-domain results, the KL divergence mea-
sure of local constant reactive power control (A3) is better than local coordinated
reactive power/voltage control (C2) due to the higher amount of voltage samples
that are closer to the pre-fault voltage. The same trend is observed for fault 2.
Finally, the VRI is used to analyse the impact of local voltage control strategies.
The results are given in Table 5.7 and the best performance is highlighted in grey.
The same trend is observed as for the CCT and the KL divergence measure. Local
constant reactive power control (A3) results in the best VRI with a V4 value of
0.66 for fault 1. The worst VRI is calculated considering local constant voltage
control (C1), which yields 0.02 due to the system collapse. If coordinated reactive
power/voltage control (C2) is applied, the calculation of the VRI results in 0.49,
which is worse than local constant reactive power control (A3) due to the reason
described in the paragraph before. The base case leads to a VRI of 0.74. Again,
the same trend is observed for fault 2.
Table 5.7: Metrics of local voltage control

Fault 1 Fault 2
Case
tCCT1 [s] K4 [–] V4 [–] tCCT2 [s] K2 [–] V2 [–]
A3 0.12 2.82 0.66 0.12 0.84 0.64
C1 0.11 71.99 0.02 0.09 62.44 0.04
C2 0.11 3.78 0.49 0.10 2.66 0.55
Base 0.15 1.47 0.74 0.12 0.24 0.87
106 Chapter 5. Stability analysis of a multi-machine system

5.3.5 Plant-level voltage control


The last section analyses the impact of the relatively slow plant-level voltage control
of PV systems on voltage dynamics. The associated cases compare different voltage
control strategies at plant level, i.e., plant-level constant reactive power control (D1),
plant-level constant voltage control (D2), plant-level constant reactive power and
local coordinated reactive power/voltage control (D3) as well as plant-level constant
voltage and local coordinated reactive power/voltage control (D4), as presented
in Table 3.4. Further details to the aforementioned control modes are given in
Section 3.3.4.
The time-domain results in response to fault 1 are provided in Figs. 5.23 to 5.26.
In order to capture the relevant dynamics that are influenced by the plant-level
voltage control of PV systems, the simulation time is extended until 20.0 s. Fo-
cusing on dynamics in the order of a few seconds, i.e., 0 s < t < 3 s, plant-level
constant reactive power control (D1) and plant-level constant voltage control (D2)
yield very similar behaviour, as shown by the voltage at the PCC in Fig. 5.23. In
the same time scale, plant-level constant reactive power and local coordinated re-
active power/voltage control (D3) as well as plant-level constant voltage and local
coordinated reactive power/voltage control (D4) act similarly and lead to a less
oscillatory bus voltage during the recovery period, as also shown in Fig. 5.23. The
aforementioned correlations of the voltage dynamics are caused by the similarities
of the active and reactive current control of PV systems, as depicted in Figs. 5.25
and 5.26, respectively. On the contrary, focusing on dynamics up to 20 s, different
control mode similarities are observed. Plant-level constant reactive power con-
trol (D1) and plant-level constant reactive power and local coordinated reactive
power/voltage control (D3) act in a similar manner and bring the reactive power
back to its pre-fault value, i.e., zero, as shown by the reactive current of the PV sys-
tem in Fig. 5.26. On the other hand, plant-level constant voltage control (D2) and
plant-level constant voltage and local coordinated reactive power/voltage control
(D4) try to restore the pre-fault voltage, as shown by the slow injection of reactive
current in Fig. 5.26 and the resulting voltage at the PCC in Fig. 5.23. From the
aforementioned observations it can be concluded, the best performance considering
both voltage dynamics is achieved using plant-level constant voltage and local co-
ordinated reactive power/voltage control (D4). It should be noted that due to the
reactive power limitation and the rather slow behaviour of the plant control, the
plant-level constant voltage control modes, i.e., plant-level constant voltage control
(D2) and plant-level constant voltage and local coordinated reactive power/voltage
control (D4), can avoid the system collapse that was experienced using local con-
stant voltage control (C1), as described in Section 5.3.4. Finally, it should be noted
that for all investigated plant-level control modes, no short-term voltage instability
is present, which is shown by the voltage at the PCC and the load bus voltage in
Figs. 5.23 and 5.24, respectively. Further time-domain results that show the impact
of plant-level voltage control of PV systems on the dynamic performance of the
Nordic test system, are given in Figs. D.17 to D.20 of Appendix D.
5.3. Case studies 107

1.0

0.8
LVRT characteristic
Voltage [pu]

0.6
1.05
0.4
1.00

0.2 0.95 D1
D2
0.0 0.5 1.0 1.5 2.0 2.5 3.0 D3
0.0 D4

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0 20.0
Time [s]

Figure 5.23: Impact of plant-level voltage control (fault 1)


 Voltage at bus 1044 (point of common coupling)

1.0

0.8
WECC criterion
Voltage [pu]

0.6 1.05

0.4 1.00

0.95
0.2 D1
D2
0.0 0.5 1.0 1.5 2.0 2.5 3.0 D3
0.0 D4

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0 20.0
Time [s]

Figure 5.24: Impact of plant-level voltage control (fault 1)


 Voltage at bus 4 (load bus)
108 Chapter 5. Stability analysis of a multi-machine system

0.0 D1
D2
D3
−0.2 D4
−0.85
Active current [pu]

−0.4
−0.90

−0.6
−0.95
0.0 0.5 1.0 1.5 2.0 2.5 3.0
−0.8

−1.0

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0 20.0
Time [s]

Figure 5.25: Impact of plant-level voltage control (fault 1)


 Active current of photovoltaic system connected to bus 1044

0.4 D1
D2
0.2 D3
D4
Reactive current [pu]

0.0

−0.2

−0.4

−0.6

−0.8

−1.0
0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0 20.0
Time [s]

Figure 5.26: Impact of plant-level voltage control (fault 1)


 Reactive current of photovoltaic system connected to bus 1044
5.3. Case studies 109

The calculation of the CCT determines the stability margin of the system. The
results are given in Table 5.8 and the best dynamic performance is highlighted in
grey. The best CCT is achieved considering either plant-level constant reactive
power control (D1) or plant-level constant voltage control (D2), leading to values
of 0.12 s for fault 1, respectively. In contrast, using plant-level constant reactive
power and local coordinated reactive power/voltage control (D3) or plant-level con-
stant voltage and local coordinated reactive power/voltage control (D4), the CCT is
worsened with values of 0.11 s, respectively. The base case leads to 0.15 s due to the
higher voltage support capability of synchronous generators. The same tendency is
observed for fault 2.
The impact on FIDVR is analysed using the KL divergence measure. The results
are shown in Table 5.8 and the best performance is highlighted in grey. Plant-level
constant voltage control (D2) leads to the best voltage recovery, with a K4 value of
2.77 for fault 1. The lowest performance is reached if plant-level constant reactive
power and local coordinated reactive power/voltage control (D3) is used, leading to
a value of 3.78. It should be noted that local constant reactive power control (A3),
as presented in Section 5.3.4, and plant-level constant reactive power control (D1),
lead to the same KL divergence measures. The same correlation exists between local
coordinated reactive power/voltage control (C2) and plant-level constant reactive
power and local coordinated reactive power/voltage control (D3). The similarities
are caused due to the same behaviour of the short-term controls of PV systems.
The same trend is observed for fault 2.
Finally, the VRI is used to analyse FIDVR. The results are given in Table 5.8
and the best performance is highlighted in grey. Plant-level constant voltage control
(D2) leads to the best VRI, with a V4 value of 0.67 for fault 1. On the contrary, the
worst VRI is calculated if plant-level constant reactive power and local coordinated
reactive power/voltage control (D3) is applied, leading to a V4 value of 0.49. The
same trend is observed for fault 2. It should be noted that similar correlations
between local and plant-level voltage control are observed for the VRI, as described
for the KL divergence measure in the paragraph before.

Table 5.8: Metrics of plant-level voltage control

Fault 1 Fault 2
Case
tCCT1 [s] K4 [–] V4 [–] tCCT2 [s] K2 [–] V2 [–]
D1 0.12 2.82 0.66 0.12 0.84 0.64
D2 0.12 2.77 0.67 0.12 0.81 0.66
D3 0.11 3.78 0.49 0.10 2.66 0.55
D4 0.11 3.67 0.50 0.10 2.19 0.60
Base 0.15 1.47 0.74 0.12 0.24 0.87
110 Chapter 5. Stability analysis of a multi-machine system

5.4 Conclusions and future work


Conclusions: This chapter presents a comprehensive analysis of an IEEE multi-
machine voltage stability test system, considering different control modes of PV
systems. The impact of these modes on short-term voltage stability and FIDVR
is investigated based on various analysis methods, i.e., time-domain simulation,
CCT, KL divergence measure, and VRI. Moreover, the influence on transient and
frequency stability is studied briefly. The results show that without the LVRT
capability of PV systems, the system is short-term voltage and transient unsta-
ble. Even the blocking mode of PV systems leads to short-term voltage instability.
Only the LVRT and DVS capability of PV systems helps avoiding instability and
improves FIDVR. Considering voltage and frequency dynamics, an intermediate
active current recovery rate of 100 %/s shows the best performance. In case of
local constant voltage control, the system collapses due to the lack of active power
produced by the PV systems. To overcome this problem, adequate reactive power
limitation is required or a different current limit logic needs to be selected. More-
over, local coordinated reactive power/voltage control can be used that combines
a fast voltage control with a slow reactive power control. Based on voltage dy-
namics up to 20 s, the best performance can be achieved using plant-level constant
voltage and local coordinated reactive power/voltage control. It should be noted
that the base case without PV systems leads to a better power system dynamic
performance due to a higher voltage support capability of synchronous generators.
The considered analysis methods reveal advantages and drawbacks. The numerical
time-domain simulations show accurate results and allow a detailed analysis of the
dynamic performance of the power system for the considered PV penetration level.
The calculation of the CCT is very helpful to determine the stability margin of the
system. However, a smaller step size would lead to somewhat more accurate values
of the CCT. Furthermore, the delayed voltage recovery is analysed using the KL
divergence measure as well as the newly defined VRI. Both metrics help to compare
different short-term voltage controls. In general, with the adequate control of PV
systems, the results show the significant improvement of power system dynamic
performance.

Future work: Future work should consider variations of the load model. In par-
ticular, the mechanical torque of the induction motor load can be modified to, e.g.,
quadratic, or a combination of different torque characteristics to form composite
mechanical loads. Another aspect that is of interest for future research is the inves-
tigation of additional operating points in the test system considering a higher PV
penetration level. Moreover, the studied contingencies could be varied in terms of
fault location, fault impedance, fault type or fault duration. Furthermore, the im-
pact of the plant control of PV systems on long-term voltage stability is of interest.
111

Conclusions and future work


6
6.1 Conclusions
This section summarises the main conclusions and highlights the scientific contri-
butions of the presented research work. The conclusions are separated according to
the main chapters of this thesis, as presented in the sequel.

Stability definitions and analysis methods: The dynamic performance of the


power system is analysed based on its different categories and the corresponding
methods. The considered analysis methods indicate advantages and drawbacks.
• Time-domain simulation: For the considered PV penetration level and the
investigated phenomena, RMS time-domain simulations show accurate results
and allow a detailed analysis of the dynamic performance of the power system.
• CCT: The CCT determines the stability margin of the system and refers
to both, short-term voltage stability and transient stability. The CCT is
calculated using a trial-and-error strategy, which yields adequate results in
most of the cases. However, in some cases the calculation of the CCT was
not sufficiently precise in order to compare different control modes of the PV
system.
• KL divergence measure: The phenomenon of FIDVR is quantified using the
recently developed KL divergence measure. This measure provides reasonable
results for the analysis of short-term controls of PV systems in most of the
cases. However, in some cases the application of the KL divergence measure
is limited.
• VRI: In order to overcome the drawbacks of the KL divergence measure, a
novel enhanced metric, namely the VRI, is introduced to quantify the phe-
nomenon of FIDVR more systematically. The VRI provides useful information
about the delayed voltage recovery and helps to compare different short-term
voltage controls of PV systems.
112 Chapter 6. Conclusions and future work

Modelling and control of PV systems: A generic PV system model is consid-


ered for the stability analysis. It represents a large-scale PV plant connected to the
high or extra high voltage level.

• Model: The positive-sequence WECC generic PV system model is imple-


mented in DIgSILENT PowerFactory. The implementation procedure and
the required modifications are detailed.

• Control: The relevant PV system control modes that impact the dynamic
performance of the power system, specifically short-term dynamics, are de-
termined. The considered control modes are the LVRT and DVS capability,
the active current recovery rate, the local voltage control and the plant-level
voltage control.

• Parameters: Typical model parameters of the PV system are provided that


were determined in collaboration with a manufacturer. The parameters rep-
resent the average behaviour of presently required control modes and settings
according to international guidelines and grid codes.

• Validation: The implemented model, along with the identified parameters, is


successfully validated for the defined control modes against the EPRI written
REMV tool, which represents the WECC specifications.

• Open source: To serve the need of academia and industry, the DIgSILENT
PowerFactory model of the WECC generic PV system is open-source software
and can be downloaded free of charge and used for any purpose.

Stability analysis: A stability analysis of a one-load infinite-bus and an IEEE


multi-machine voltage stability test system, i.e., the Nordic test system, is per-
formed considering different PV system control modes. The results show the im-
provement of the power system dynamic performance using an adequate control of
PV systems.

• LVRT and DVS capability: The time-domain results show that without the
LVRT capability of PV systems, the system is unstable. Even the blocking
mode of PV systems is not sufficient to prevent short-term voltage instability.
Only the LVRT and DVS capability of PV systems helps avoiding instability
and improves FIDVR. The CCT that determines the stability margin of the
system, shows the same tendency. The CCT is significantly improved using
the LVRT and DVS capability of PV systems. In the same manner, the KL
divergence measure as well as the VRI indicate the enhancement of the voltage
recovery applying the latter control mode.
6.1. Conclusions 113

• Active current recovery rate: The time-domain results show that the best
voltage recovery is achieved using the fastest active current recovery rate of
PV systems. However, this recovery rate also leads to overfrequencies in the
system that could cause disconnection of generation. Therefore, the active
current recovery rate of PV systems should be tuned considering both, volt-
age and frequency dynamics. The CCT shows the best performance with
the fastest active current recovery rate. The same trend is observed for the
analysis of FIDVR using the KL divergence measure and the VRI.
• Local voltage control: The results for the multi-machine system with a medium
PV penetration level are different compared to the one-load infinite-bus sys-
tem with a high PV penetration level. The time-domain simulations for the
multi-machine system show that in case of local constant voltage control, PV
systems try to restore their pre-fault voltage by increasing their reactive cur-
rent. However, due to the rather stiff grid behaviour, restoring grid voltage
is impossible. The active current is reduced to zero due to the missing re-
active power limitation, and the system collapses owing to the lack of active
power produced by the PV systems. Although counterintuitive, the results
show that a high reactive power injection is not always the best to counteract
a voltage collapse. Therefore, adequate reactive power limitation is required
or a different current limit logic of the PV system needs to be selected. The
stability margin in terms of the CCT as well as the KL divergence measure
and the VRI that quantify FIDVR, show the best performance considering
local constant reactive power control. The time-domain simulations of the
one-load infinite-bus system yield the best results in case of local constant
voltage control of PV systems. The same holds true for the CCT, the KL
divergence measure and the VRI, considering this test system.
• Plant-level voltage control: If plant control of PV systems is used, the time-
domain results show the best performance using plant-level constant volt-
age and local coordinated reactive power/voltage control. This control mode
combines a fast voltage control that injects a fast reactive current to improve
FIDVR, together with a slow voltage control that slowly injects reactive power
to restore the grid voltage. It is shown that the fast voltage control leads to
a less oscillatory voltage evolution during the recovery period. The enhanced
short-term dynamics are also confirmed by the CCT, the KL divergence mea-
sure and the VRI for the one-load infinite-bus system. On the contrary, for
the multi-machine system, the latter metrics are worsened using plant-level
constant voltage and local coordinated reactive power/voltage control owing
to the rather stiff grid behaviour. This leads to a higher deviation to the
pre-fault voltage, which deteriorates the KL divergence measure and the VRI.
Finally, it is important to mention that the plant control of the PV system
model needs to be considered for short-term voltage stability and FIDVR
studies because it can significantly influence short-term dynamics.
114 Chapter 6. Conclusions and future work

6.2 Future work


This section draws the directions for future research based on the results of this
work. The future work is separated according to the main chapters of this thesis,
as presented in the sequel.

Stability definitions and analysis methods: Improvements of the applied


power system dynamic performance analysis methods are elaborated. In addition,
other promising analysis methods for short-term voltage stability are identified.

• Time-domain simulation: For very high PV penetration levels and weak sys-
tem conditions, RMS time-domain simulations might not be accurate or suit-
able anymore. Instead, combined RMS and EMT, or pure EMT analysis is
required.

• CCT: The calculation of the CCT is very helpful to determine the stability
margin of the power system. However, a smaller step size for the simulations
would lead to somewhat more accurate values of the CCT.

• KL divergence measure: If the actual voltage recovery is better than the


reference voltage recovery, in terms of the PDF, the KL divergence measure
increases. Thus, it suggests a lower quality of recovery, which is contradictory.
Hence, the PDF of the reference voltage recovery should be modified, specifi-
cally the concentration parameter of the density near the pre-fault voltage.

• VRI: Different weighting functions for overvoltages, i.e., voltages higher than
the normal operating voltage, could be introduced for the calculation of the
VRI.

• Lyapunov exponents: The application of Lyaponov exponents seems to be a


promising method for the analysis of short-term voltage stability as it is a
data-driven, model-free approach based on a finite time interval of voltage
samples. Its suitability for the analysis of short-term controls of PV systems
is of interest.

• Transient P–V curves: Transient P–V curves could be used to determine the
stability boundary of the load bus including PV generation for the analysis of
short-term voltage stability. It is an analytical method and the P–V curves
represent the power transfer capability to the load at a given time.
6.2. Future work 115

Modelling and control of PV systems: The level of detail of the PV system


model is crucial. For very high PV penetration levels and weak system conditions,
the WECC generic PV system model with its control might not be adequate any-
more to capture all relevant power system phenomena. Therefore, further modelling
and control aspects need to be considered.
• Model: The inner-current control loops as well as detailed PLL modelling of
the PV plant should be considered to study weak system conditions. More-
over, the generator/converter model implementation as a current source poses
numerical challenges for state-of-the-art software simulation tools, which can
be solved using a voltage source model.
• Control: Additional control modes of the WECC generic PV system model
should be considered. For instance, negative-sequence control or grid-forming
control are of interest. Moreover, the current limit logic could be changed,
i.e., active current priority could be selected.

Stability analysis: The results of the stability analysis of the one-load infinite-
bus system and the multi-machine system, namely the Nordic test system, indicate
further directions for future research.
• Load modelling: Additional variations of the load model could be considered.
In particular, the mechanical torque of the induction motor load can be mod-
ified to, e.g., quadratic, or a combination of different torque characteristics
to form composite mechanical loads, which would affect short-term voltage
stability and FIDVR.
• Operating point: Additional operating points in the test system should be con-
sidered, e.g., PV systems that do not operate at unity power factor. Moreover,
higher PV penetration levels are also of interest.
• Contingency: The studied contingencies could be varied in terms of fault loca-
tion, fault impedance, fault type or fault duration in order to affect different
PV systems, induction motors and synchronous generators in the test system.
• Long-term voltage stability: The impact of the plant control of PV systems
on long-term voltage stability is of interest. Therefore, the test system should
include the associated controls of synchronous generators as well as the load
tap changer control of step-down transformers.
• Small-signal stability: The impact of PV system control on small-signal sta-
bility is of interest. The WECC generic PV system model is highly non-linear
due to freeze functions and deadbands. Therefore, it is not necessarily ap-
propriate for linearisation and eigenvalue analysis. However, if small-signal
stability analysis is of interest, closed-form equations of the system should be
developed based on strict linearisation of the non-linear model.
117

Appendices
119

Stability definitions and analysis methods


A
A.1 Kullback-Leibler divergence measure
A.1.1 Voltage violation criterion

Table A.1: Parameters of WECC voltage violation criterion

Description Symbol Value Unit


Voltage limit 1 of WECC violation criterion V1 0.75 [pu]
Voltage limit 2 of WECC violation criterion V2 0.8 [pu]
Voltage limit 3 of WECC violation criterion V3 0.95 [pu]
Time limit 1 of WECC violation criterion t1 0.8 [s]
Time limit 2 of WECC violation criterion t2 2.5 [s]

A.1.2 Probability density function of voltage sample

Table A.2: Parameters for construction of probability density function

Description Symbol Value Unit


Minimum voltage Vmin 0 [pu]
Maximum voltage Vmax 1.2 [pu]
Time at the fault clearing instant tcl 0.1 [s]
Final observation time tf 3.0 [s]
Total number of voltage sub-intervals L 100 [–]
Total number of voltage limits M 3 [–]
120

A.2
A.2.1

Table A.3: Parameters of weighting coefficients ———————

Description Symbol Value Unit


Weighting function

Rewarding weighting coefficient for voltage limit 1 of WECC violation criterion λ+


1 70 [–]
Rewarding weighting coefficient for voltage limit 2 of WECC violation criterion λ+
2 150 [–]
Rewarding weighting coefficient for voltage limit 3 of WECC violation criterion λ+
3 3000 [–]
Voltage recovery index

Penalising weighting coefficient for voltage limit 1 of WECC violation criterion λ−


1 11 [–]
Penalising weighting coefficient for voltage limit 2 of WECC violation criterion λ−
2 8 [–]
Penalising weighting coefficient for voltage limit 3 of WECC violation criterion λ−
3 5 [–]
Appendix A. Stability definitions and analysis methods
121

Modelling and control of photovoltaic systems


B
B.1 Model
B.1.1 Electrical control model
Anti-windup limiters: The anti-windup function is used to prevent the integrator
from winding up. When the control limits are reached, the integration has to
be stopped, otherwise the integrator would continue to integrate, which results in
a substantial overshoot. The basic mechanism is explained using three different
configurations of integrator models, as shown in Fig. B.1. The first integrator is
without any limits, as depicted in Fig. B.1 (a), and the system equation is defined
as:
dx
=y (B.1)
dt
The second integrator is with windup limits, as shown in Fig. B.1 (b), and the
system equation is defined as:
dx̃
=y (B.2)
dt
which is similar to Eq. (B.1). However, the additional windup limits can be defined

xmax xmax

y 1 x y 1 x̃ x y 1 x
s s s

(a) (b) xmin (c) xmin

Figure B.1: Block diagrams of integrator models with different limiters


(a) Integrator without limits
(b) Integrator with windup limits
(c) Integrator with anti-windup limits
122 Appendix B. Modelling and control of photovoltaic systems

according to: ⎧

⎨xmax if x̃ ≥ xmax
x = x̃ if xmin < x̃ < xmax (B.3)


xmin if x̃ ≤ xmin
The third integrator shown in Fig. B.1 (c) includes anti-windup limits. The system
equation has the form of:
dx
=y (B.4)
dt
and the anti-windup limits are defined as:
dx dx
if x ≥ xmax and ≥ 0 → x = xmax and =0
dt dt
dx dx
if x ≤ xmin and ≤ 0 → x = xmin and =0 (B.5)
dt dt
dx
otherwise → =y
dt
The windup effect of the three configurations is illustrated in Fig. B.2 with y(t) =
sin(t), xmax = 0.5 and xmin = −0.5. For x̃ > xmax at t = 1.05 s the output x is
locked and remains locked until x̃ < xmax at t = 5.23 s. This is the typical windup
effect of the integrator. The anti-windup function solves this problem, i.e., x starts
decreasing at t = 3.14 s, which is the instant for which dx/dt becomes negative.
The anti-windup limiter of the PI controller in DIgSILENT PowerFactory uses a
semi-implicit formulation [111].

2.0

1.5

1.0
Signals

0.5

0.0
y(t)
−0.5 x(t) (without limits)
x(t) (windup)
x(t) (anti-windup)
−1.0
0.0 1.0 2.0 3.0 4.0 5.0 6.0
Time [s]

Figure B.2: Responses of integrator models with different limiters


B.1. Model 123

Freeze function: If a fault in the system occurs, the function freezes the state
variables of the control blocks that are used for normal operating conditions. The
function is activated only if the PV system turns from the normal operating condi-
tion into the fault condition, defined by the Voltage dip signal shown in Eq. (3.4).
The freeze function is implemented as follows:

dx y if Voltage dip = 0 (normal operating condition)


= (B.6)
dt 0 if Voltage dip = 1 (fault condition)

A simple integrator model with anti-windup limits and the Voltage dip signal is
shown in Fig. B.3. A numerical example of the freeze function with the input signal
y(t) = sin(t) and the limits xmax = 0.5 and xmin = −0.5 is shown in Fig. B.4.
The freeze function is activated from 3 s to 5 s. During the fault-on period, i.e.,
Voltage dip = 1, the state variable x is frozen and remains at the value 0.5. At
t = 5 s, the model returns to the normal operating condition and x starts decreasing.
Voltage dip
xmax

y 1 x
s

xmin

Figure B.3: Block diagram of integrator model with anti-windup limits and freeze func-
tion

1.0

0.5
Signals

0.0

−0.5 y(t)
Voltage dip
x(t) (without freeze function)
x(t) (with freeze function)
−1.0
0.0 1.0 2.0 3.0 4.0 5.0 6.0
Time [s]

Figure B.4: Responses of integrator models with and without freeze function
124

B.2
Table B.1: Parameters of generator/converter model —————— B.2.1
Model Description Symbol Value Unit
REGC A 0 High voltage clamp logic acceleration factor Khv 2 [–]
REGC A 0 Voltage limit for high voltage clamp logic Volim 1.2 [pu]
Parameters

REGC A 0 Inverter reactive current regulator lag time constant Tgq 0.02 [s]
REGC A 0 Current limit for high voltage clamp logic Iolim −1 [pu]
REGC A 0 Maximum rate-of-change of reactive current Iqrmax 20 [pu/s]
REGC A 0 Minimum rate-of-change of reactive current Iqrmin −20 [pu/s]
REGC A 1 Terminal voltage filter time constant Tfltr 0.01 [s]
REGC A 1 Inverter active current regulator lag time constant Tgp 0.02 [s]
REGC A 1 Enable (1) or disable (0) low voltage power logic Lvplsw [0 1]∗ [–]
Generator/converter model

REGC A 1 Maximum rate-of-change of active current rrpwr [0.1 0.2 1 10]∗ [pu/s]
REGC A 1 Low voltage power logic zero crossing zerox 0.7 [pu]
REGC A 1 Low voltage power logic breakpoint brkpt 0.9 [pu]
REGC A 1 Low voltage power logic gain breakpoint Lvpl1 1 [pu/pu]
REGC A 1 Low voltage active current management breakpoint 0 lvpnt0 0.01 [pu]
REGC A 1 Low voltage active current management breakpoint 1 lvpnt1 0.05 [pu]
∗ The value of this parameter varies according to the selected control mode.
Appendix B. Modelling and control of photovoltaic systems
B.2.2
Table B.2: Parameters of electrical control model ————————-

Model Description Symbol Value Unit


REEC B 0 Priority to reactive current (0) or active current (1) PqFlag 0 [–]
B.2. Parameters

REEC B 0 Maximum apparent current Imax 1 [pu]


REEC B 1 Terminal bus voltage filter time constant Trv 0.01 [s]
REEC B 1 Over voltage deadband for reactive current injection dbd1 −0.1 [pu]
REEC B 1 Under voltage deadband for reactive current injection dbd2 0.1 [pu]
REEC B 1 Reactive current injection gain Kqv [0 4]∗ [pu]
REEC B 1 Low voltage condition trigger Vdip 0.9 [pu]
REEC B 1 High voltage condition trigger Vup 1.1 [pu]
REEC B 1 Maximum reactive current injection Iqh1 1.3 [pu]
REEC B 1 Minimum reactive current injection Iql1 −1.3 [pu]
Electrical control model

REEC B 2 Bypass (0) or engage (1) inner voltage regulator loop QFlag [0 1]∗ [–]
REEC B 2 Local voltage regulator proportional gain Kvp 12 [pu/pu]
REEC B 2 Local voltage regulator integral gain Kvi 300 [pu/pu-s]
REEC B 2 Reactive current regulator lag time constant Tiq 0.01 [s]
REEC B 2 Local reactive power regulator proportional gain Kqp 0.05 [pu/pu]
REEC B 2 Local reactive power regulator integral gain Kqi 0.1 [pu/pu-s]
REEC B 2 Local reactive power (0) or voltage (1) control VFlag [0 1]∗ [–]
REEC B 2 Maximum voltage at inverter terminal bus Vmax 1.1 [pu]
REEC B 2 Minimum voltage at inverter terminal bus Vmin 0.9 [pu]
REEC B 2 Maximum reactive power if VFlag = 1 Qmax 0.4 [pu]
REEC B 2 Minimum reactive power if VFlag = 1 Qmin −0.4 [pu]
∗ The value of this parameter varies according to the selected control mode.
125
126

Table B.2: Parameters of electrical control model (continued)—————————

Model Description Symbol Value Unit


REEC B 3 Inverter power order lag time constant Tpord 0.01 [s]
REEC B 3 Active power down-ramp limit dPmin −1 [pu/s]
REEC B 3 Active power up-ramp limit dPmax 1 [pu/s]
REEC B 3 Maximum active power Pmax 1 [pu]
REEC B 3 Minimum active power Pmin 0 [pu]
REEC B 4 Active power filter time constant Tp 0.01 [s]
REEC B 4 Constant reactive power (0) or power factor (1) control PfFlag 0 [–]
Appendix B. Modelling and control of photovoltaic systems
B.2.3

Table B.3: Parameters of plant control model ———————


B.2. Parameters

Model Description Symbol Value Unit


REPC A 0 Line drop compensation resistance if VcmpFlag = 1 Rc 0 [pu]
REPC A 0 Line drop compensation reactance if VcmpFlag = 1 Xc 0 [pu]
REPC A 0 Reactive droop if VcmpFlag = 0 Kc 0 [pu]
REPC A 0 Reactive droop (0) or line drop compensation (1) VcmpFlag 0 [–]
REPC A 1 Voltage filter time constant Tvfltr 0.02 [s]
REPC A 1 Reactive power filter time constant Tqfltr 0.02 [s]
Plant control model

REPC A 1 Plant-level reactive power (0) or voltage (1) control RefFlag [0 1]∗ [–]
Reactive power deadband if RefFlag = 0 or
REPC A 1 dbd 0.005 [pu]
voltage deadband if RefFlag = 1
REPC A 1 Volt/var regulator proportional gain Kp 0.36 [pu/pu]
REPC A 1 Volt/var regulator integral gain Ki 0.18 [pu/pu-s]
REPC A 1 Voltage for freezing volt/var regulator integrator Vfrz 0.9 [pu]
REPC A 1 Plant controller reactive power output lead time constant Tft 0 [s]
REPC A 1 Plant controller reactive power output lag time constant Tfv 0.1 [s]
REPC A 1 Maximum volt/var error emax 0.1 [pu]
REPC A 1 Minimum volt/var error emin −0.1 [pu]
REPC A 1 Maximum plant reactive power Qmax plant 0.4 [pu]
REPC A 1 Minimum plant reactive power Qmin plant −0.4 [pu]
∗ The value of this parameter varies according to the selected control mode.
127
128

Table B.3: Parameters of plant control model (continued) ————————

Model Description Symbol Value Unit


REPC A 2 Frequency reference Freq ref 1 [pu]
REPC A 2 Over frequency deadband for governor response fdbd1 0 [pu]
REPC A 2 Under frequency deadband for governor response fdbd2 0 [pu]
REPC A 2 Down regulation droop Ddn 20 [pu/pu]
REPC A 2 Up regulation droop Dup 10 [pu/pu]
REPC A 2 Active power filter time constant Tp 0.02 [s]
REPC A 2 Droop regulator proportional gain Kpg 1 [pu/pu]
REPC A 2 Droop regulator integral gain Kig 10 [pu/pu-s]
REPC A 2 Plant controller active power output lag time constant Tlag 0.1 [s]
REPC A 2 Governor response disable (0) or enable (1) FrqFlag 0 [–]
REPC A 2 Maximum power error in droop regulator femax 999 [pu]
REPC A 2 Minimum power error in droop regulator femin −999 [pu]
REPC A 2 Maximum plant active power Pmax plant 1 [pu]
REPC A 2 Minimum plant active power Pmin plant 0 [pu]
Appendix B. Modelling and control of photovoltaic systems
B.3. Test system 129

B.3 Test system


B.3.1 Overview

Table B.4: Parameters of single-converter equivalent system

Description Symbol Value Unit


Step-up transformer equivalent nominal apparent power Streq 110 [MVA]
Step-up transformer equivalent voltage on high voltage side Vtreqh 20 [kV]
Step-up transformer equivalent voltage on low voltage side Vtreql 0.6 [kV]
Step-up transformer equivalent resistance Rtreq 0 [pu]
Step-up transformer equivalent reactance Xtreq 0.0001 [pu]
Step-up transformer equivalent tap ratio ntreq 1 [pu/pu]
Collector system equivalent nominal voltage Vcoleq 20 [kV]
Collector system equivalent resistance Rcoleq 0 [pu]
Collector system equivalent reactance Xcoleq 0.001 [pu]
Collector system equivalent susceptance Bcoleq 0.032 [pu]
Substation transformer nominal apparent power Ssubtr 110 [MVA]
Substation transformer voltage on high voltage side Vsubtrh 110 [kV]
Substation transformer voltage on low voltage side Vsubtrl 20 [kV]
Substation transformer resistance Rsubtr 0 [pu]
Substation transformer reactance Xsubtr 0.0001 [pu]
Substation transformer tap ratio nsubtr 1 [pu/pu]
130 Appendix B. Modelling and control of photovoltaic systems

B.4 Validation
B.4.1 Low voltage ride-through and dynamic voltage
support

0.0 A2 (REMV)
A2 (PowerFactory)

−0.2
Active current [pu]

−0.4

−0.6

−0.8

−1.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]
Figure B.5: Validation of low voltage ride-through and dynamic voltage support
 Active current of photovoltaic system

0.0

−0.2
Reactive current [pu]

−0.4

−0.6

−0.8

A2 (REMV)
−1.0 A2 (PowerFactory)

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]
Figure B.6: Validation of low voltage ride-through and dynamic voltage support
 Reactive current of photovoltaic system
B.4. Validation 131

0.0 A3 (REMV)
A3 (PowerFactory)

−0.2
Active current [pu]

−0.4

−0.6

−0.8

−1.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure B.7: Validation of low voltage ride-through and dynamic voltage support
 Active current of photovoltaic system

0.0

−0.2
Reactive current [pu]

−0.4

−0.6

−0.8

A3 (REMV)
−1.0 A3 (PowerFactory)

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure B.8: Validation of low voltage ride-through and dynamic voltage support
 Reactive current of photovoltaic system
132 Appendix B. Modelling and control of photovoltaic systems

B.4.2 Active current recovery rate

0.0 B1 (REMV)
B1 (PowerFactory)

−0.2
Active current [pu]

−0.4

−0.6

−0.8

−1.0

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0
Time [s]

Figure B.9: Validation of active current recovery rate


 Active current of photovoltaic system

0.0

−0.2
Reactive current [pu]

−0.4

−0.6

−0.8

B1 (REMV)
−1.0 B1 (PowerFactory)

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0
Time [s]

Figure B.10: Validation of active current recovery rate


 Reactive current of photovoltaic system
B.4. Validation 133

0.0 B2 (REMV)
B2 (PowerFactory)

−0.2
Active current [pu]

−0.4

−0.6

−0.8

−1.0

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0
Time [s]

Figure B.11: Validation of active current recovery rate


 Active current of photovoltaic system

0.0

−0.2
Reactive current [pu]

−0.4

−0.6

−0.8

B2 (REMV)
−1.0 B2 (PowerFactory)

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0
Time [s]

Figure B.12: Validation of active current recovery rate


 Reactive current of photovoltaic system
134 Appendix B. Modelling and control of photovoltaic systems

0.0 B3 (REMV)
B3 (PowerFactory)

−0.2
Active current [pu]

−0.4

−0.6

−0.8

−1.0

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0
Time [s]

Figure B.13: Validation of active current recovery rate


 Active current of photovoltaic system

0.0

−0.2
Reactive current [pu]

−0.4

−0.6

−0.8

B3 (REMV)
−1.0 B3 (PowerFactory)

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0
Time [s]

Figure B.14: Validation of active current recovery rate


 Reactive current of photovoltaic system
B.4. Validation 135

B.4.3 Local voltage control

0.0

−0.2
C1 (REMV)
Active current [pu]

C1 (PowerFactory–Qgen)
−0.4 C1 (PowerFactory–Vt filt)

−0.6

−0.8

−1.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure B.15: Validation of local voltage control


 Active current of photovoltaic system

1.0

0.5 C1 (REMV)
Reactive current [pu]

C1 (PowerFactory–Qgen)
C1 (PowerFactory–Vt filt)
0.0

−0.5

−1.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Time [s]

Figure B.16: Validation of local voltage control


 Reactive current of photovoltaic system
136 Appendix B. Modelling and control of photovoltaic systems

0.0 C2 (REMV)
C2 (PowerFactory)

−0.2
Active current [pu]

−0.4

−0.6

−0.8

−1.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure B.17: Validation of local voltage control


 Active current of photovoltaic system

1.0
C2 (REMV)
C2 (PowerFactory)

0.5
Reactive current [pu]

0.0

−0.5

−1.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Time [s]

Figure B.18: Validation of local voltage control


 Reactive current of photovoltaic system
B.4. Validation 137

B.4.4 Plant-level voltage control

0.0 D1 (REMV)
D1 (PowerFactory)

−0.2
Active current [pu]

−0.4

−0.6

−0.8

−1.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure B.19: Validation of plant-level voltage control


 Active current of photovoltaic system

0.0

−0.2
Reactive current [pu]

−0.4

−0.6

−0.8

D1 (REMV)
−1.0 D1 (PowerFactory)

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure B.20: Validation of plant-level voltage control


 Reactive current of photovoltaic system
138 Appendix B. Modelling and control of photovoltaic systems

0.0 D2 (REMV)
D2 (PowerFactory)

−0.2
Active current [pu]

−0.4

−0.6

−0.8

−1.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure B.21: Validation of plant-level voltage control


 Active current of photovoltaic system

0.0

−0.2
Reactive current [pu]

−0.4

−0.6

−0.8

D2 (REMV)
−1.0 D2 (PowerFactory)

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure B.22: Validation of plant-level voltage control


 Reactive current of photovoltaic system
B.4. Validation 139

0.0 D3 (REMV)
D3 (PowerFactory)

−0.2
Active current [pu]

−0.4

−0.6

−0.8

−1.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure B.23: Validation of plant-level voltage control


 Active current of photovoltaic system

0.0

−0.2
Reactive current [pu]

−0.4

−0.6

−0.8

D3 (REMV)
−1.0 D3 (PowerFactory)

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure B.24: Validation of plant-level voltage control


 Reactive current of photovoltaic system
140 Appendix B. Modelling and control of photovoltaic systems

0.0 D4 (REMV)
D4 (PowerFactory)

−0.2
Active current [pu]

−0.4

−0.6

−0.8

−1.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure B.25: Validation of plant-level voltage control


 Active current of photovoltaic system

0.0

−0.2
Reactive current [pu]

−0.4

−0.6

−0.8

D4 (REMV)
−1.0 D4 (PowerFactory)

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure B.26: Validation of plant-level voltage control


 Reactive current of photovoltaic system
141

Stability analysis of a one-load infinite-bus system


C
C.1 Test system
C.1.1 Overview

Table C.1: Parameters of one-load infinite-bus system

Description Symbol Value Unit


Step-down transformer nominal apparent power Sdntr 600 [MVA]
Step-down transformer voltage on high voltage side Vdntrh 220 [kV]
Step-down transformer voltage on low voltage side Vdntrl 20 [kV]
Step-down transformer resistance Rdntr 0 [pu]
Step-down transformer reactance Xdntr 0.1 [pu]
Step-down transformer tap ratio ndntr 1.04 [pu/pu]
Transmission line 1 nominal voltage Vtl1 220 [kV]
Transmission line 1 resistance Rtl1 0.012 [pu]
Transmission line 1 reactance Xtl1 0.09 [pu]
Transmission line 1 susceptance Btl1 0.0152 [pu]
Transmission line 2 nominal voltage Vtl2 220 [kV]
Transmission line 2 resistance Rtl2 0.012 [pu]
Transmission line 2 reactance Xtl2 0.09 [pu]
Transmission line 2 susceptance Btl2 0.0152 [pu]
142 Appendix C. Stability analysis of a one-load infinite-bus system

C.1.2 Load modelling

Table C.2: Parameters of exponential load

Description Symbol Value Unit


Exponential load active power exponent α 2 [–]
Exponential load reactive power exponent β 2 [–]
Exponential load power factor PF exp 0.95ind [–]

Table C.3: Parameters of small induction motor load

Description Symbol Value Unit


Small induction motor nominal apparent power Ssim 150 [MVA]
Small induction motor stator resistance Rsims 0.031 [pu]
Small induction motor stator leakage reactance Xsims 0.1 [pu]
Small induction motor magnetising reactance Xsimm 3.2 [pu]
Small induction motor rotor resistance Rsimr 0.018 [pu]
Small induction motor rotor leakage reactance Xsimr 0.18 [pu]
Small induction motor inertia constant Hsim 0.7 [s]
Small induction motor power factor PF sim 0.6ind [–]

Table C.4: Parameters of large induction motor load

Description Symbol Value Unit


Large induction motor nominal apparent power Slim 112.5 [MVA]
Large induction motor stator resistance Rlims 0.013 [pu]
Large induction motor stator leakage reactance Xlims 0.067 [pu]
Large induction motor magnetising reactance Xlimm 3.8 [pu]
Large induction motor rotor resistance Rlimr 0.009 [pu]
Large induction motor rotor leakage reactance Xlimr 0.17 [pu]
Large induction motor inertia constant Hlim 1.5 [s]
Large induction motor power factor PF lim 0.8ind [–]
C.1. Test system 143

C.1.3 Modelling of photovoltaic system

Table C.5: Parameters of single-converter equivalent system

Description Symbol Value Unit


Step-up transformer equivalent nominal apparent power Streq 330 [MVA]
Step-up transformer equivalent voltage on high voltage side Vtreqh 20 [kV]
Step-up transformer equivalent voltage on low voltage side Vtreql 0.6 [kV]
Step-up transformer equivalent resistance Rtreq 0.0088 [pu]
Step-up transformer equivalent reactance Xtreq 0.0593 [pu]
Step-up transformer equivalent tap ratio ntreq 1 [pu/pu]
Collector system equivalent nominal voltage Vcoleq 20 [kV]
Collector system equivalent resistance Rcoleq 0.016 [pu]
Collector system equivalent reactance Xcoleq 0.013 [pu]
Collector system equivalent susceptance Bcoleq 0.032 [pu]
Substation transformer nominal apparent power Ssubtr 350 [MVA]
Substation transformer voltage on high voltage side Vsubtrh 220 [kV]
Substation transformer voltage on low voltage side Vsubtrl 20 [kV]
Substation transformer resistance Rsubtr 0 [pu]
Substation transformer reactance Xsubtr 0.15 [pu]
Substation transformer tap ratio nsubtr 1.05 [pu/pu]
Interconnection transmission line nominal voltage Vitl 220 [kV]
Interconnection transmission line resistance Ritl 0.002 [pu]
Interconnection transmission line reactance Xitl 0.014 [pu]
Interconnection transmission line susceptance Bitl 0.03 [pu]

C.1.4 Operating point

Table C.6: Power flow results of one-load infinite-bus system

Bus Vbus [kV] ϕbus [◦ ]


1 0.578 5.6
2 19.133 2.2
3 18.840 1.5
4 219.594 −5.8
5 219.631 −8.2
6 220.000 0
7 20.011 −13.8
144 Appendix C. Stability analysis of a one-load infinite-bus system

C.2 Case studies


C.2.1 Active current recovery rate

·10−2

B1
8 ·10−2 B2
1.8 A3
B3
6 1.6

1.4
Slip [pu]

4 1.2
0.0 0.5 1.0 1.5 2.0 2.5 3.0

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0
Time [s]
Figure C.1: Impact of active current recovery rate
 Slip of small induction motor load connected to bus 7

·10−2

B1
8 ·10−2 B2
1.5 A3
B3
6
1
Slip [pu]

4 0.5
0.0 0.5 1.0 1.5 2.0 2.5 3.0

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0
Time [s]
Figure C.2: Impact of active current recovery rate
 Slip of large induction motor load connected to bus 7
C.2. Case studies 145

C.2.2 Local voltage control

·10−2

A3
8 ·10−2 C1
1.6 C2
1.4
6
1.2
Slip [pu]

1
4 0.8
0.0 0.5 1.0 1.5 2.0 2.5 3.0

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0
Time [s]

Figure C.3: Impact of local voltage control


 Slip of small induction motor load connected to bus 7

·10−2

A3
8 ·10−2 C1
C2
1
6
0.5
Slip [pu]

4
0
0.0 0.5 1.0 1.5 2.0 2.5 3.0

0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0 9.0 10.0
Time [s]

Figure C.4: Impact of local voltage control


 Slip of large induction motor load connected to bus 7
146 Appendix C. Stability analysis of a one-load infinite-bus system

C.2.3 Plant-level voltage control

·10−2

D1
8 ·10−2 D2
1.6 D3
1.4 D4
6
1.2
Slip [pu]

1
4 0.8
0.0 0.5 1.0 1.5 2.0 2.5 3.0

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0 20.0
Time [s]

Figure C.5: Impact of plant-level voltage control


 Slip of small induction motor load connected to bus 7

·10−2

D1
8 ·10−2 D2
D3
1 D4
6
0.5
Slip [pu]

4
0
0.0 0.5 1.0 1.5 2.0 2.5 3.0

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0 20.0
Time [s]

Figure C.6: Impact of plant-level voltage control


 Slip of large induction motor load connected to bus 7
147

Stability analysis of a multi-machine system


D
D.1 Test system
D.1.1 Overview

Table D.1: Parameters of step-down transformers

From bus To bus ndntr [pu/pu]


1 1041 1.01
2 1042 1.01
3 1043 1.02
4 1044 1.00
5 1045 1.00
11 1011 1.06
12 1012 1.06
13 1013 1.04
22 1022 1.08
31 2031 1.07
32 2032 1.07
41 4041 1.09
42 4042 1.08
43 4043 1.07
46 4046 1.06
47 4047 1.06
51 4051 1.08
61 4061 1.05
62 4062 1.05
63 4063 1.03
71 4071 1.03
72 4072 1.05
148 Appendix D. Stability analysis of a multi-machine system

D.1.2 Load modelling

Table D.2: Parameters of exponential loads

Bus α [–] β [–] PF exp [–]


1 2 2 0.97ind
2 2 2 0.98ind
3 2 2 0.95ind
4 2 2 0.96ind
5 2 2 0.97ind
11 2 2 0.95ind
12 2 2 0.95ind
13 2 2 0.95ind
22 2 2 0.96ind
31 2 2 0.97ind
32 2 2 0.98ind
41 2 2 0.97ind
42 2 2 0.95ind
43 2 2 0.96ind
46 2 2 0.96ind
47 2 2 0.92ind
51 2 2 0.95ind
61 2 2 0.97ind
62 2 2 0.96ind
63 2 2 0.91ind
71 1 2 0.96ind
72 1 2 0.98ind
Table D.3: Parameters of small induction motor loads ————————-

Bus Ssim [MVA] Rsims [pu] Xsims [pu] Xsimm [pu] Rsimr [pu] Xsimr [pu] Hsim [s] PF sim [–]
D.1. Test system

1 150 0.031 0.1 3.2 0.018 0.18 0.7 0.6ind


2 83 0.031 0.1 3.2 0.018 0.18 0.7 0.6ind
3 65 0.031 0.1 3.2 0.018 0.18 0.7 0.6ind
4 210 0.031 0.1 3.2 0.018 0.18 0.7 0.6ind
5 180 0.031 0.1 3.2 0.018 0.18 0.7 0.6ind
11 50 0.031 0.1 3.2 0.018 0.18 0.7 0.6ind
12 75 0.031 0.1 3.2 0.018 0.18 0.7 0.6ind
13 25 0.031 0.1 3.2 0.018 0.18 0.7 0.6ind
22 70 0.031 0.1 3.2 0.018 0.18 0.7 0.6ind
31 25 0.031 0.1 3.2 0.018 0.18 0.7 0.6ind
32 50 0.031 0.1 3.2 0.018 0.18 0.7 0.6ind
41 135 0.031 0.1 3.2 0.018 0.18 0.7 0.6ind
42 100 0.031 0.1 3.2 0.018 0.18 0.7 0.6ind
43 225 0.031 0.1 3.2 0.018 0.18 0.7 0.6ind
46 175 0.031 0.1 3.2 0.018 0.18 0.7 0.6ind
47 25 0.031 0.1 3.2 0.018 0.18 0.7 0.6ind
51 200 0.031 0.1 3.2 0.018 0.18 0.7 0.6ind
61 125 0.031 0.1 3.2 0.018 0.18 0.7 0.6ind
62 75 0.031 0.1 3.2 0.018 0.18 0.7 0.6ind
63 148 0.031 0.1 3.2 0.018 0.18 0.7 0.6ind
149
150

Table D.4: Parameters of large induction motor loads —————————

Bus Slim [MVA] Rlims [pu] Xlims [pu] Xlimm [pu] Rlimr [pu] Xlimr [pu] Hlim [s] PF lim [–]
1 113 0.013 0.067 3.8 0.009 0.17 1.5 0.8ind
2 62 0.013 0.067 3.8 0.009 0.17 1.5 0.8ind
3 49 0.013 0.067 3.8 0.009 0.17 1.5 0.8ind
4 158 0.013 0.067 3.8 0.009 0.17 1.5 0.8ind
5 135 0.013 0.067 3.8 0.009 0.17 1.5 0.8ind
11 38 0.013 0.067 3.8 0.009 0.17 1.5 0.8ind
12 56 0.013 0.067 3.8 0.009 0.17 1.5 0.8ind
13 19 0.013 0.067 3.8 0.009 0.17 1.5 0.8ind
22 53 0.013 0.067 3.8 0.009 0.17 1.5 0.8ind
31 19 0.013 0.067 3.8 0.009 0.17 1.5 0.8ind
32 38 0.013 0.067 3.8 0.009 0.17 1.5 0.8ind
41 101 0.013 0.067 3.8 0.009 0.17 1.5 0.8ind
42 75 0.013 0.067 3.8 0.009 0.17 1.5 0.8ind
43 169 0.013 0.067 3.8 0.009 0.17 1.5 0.8ind
46 131 0.013 0.067 3.8 0.009 0.17 1.5 0.8ind
47 19 0.013 0.067 3.8 0.009 0.17 1.5 0.8ind
51 150 0.013 0.067 3.8 0.009 0.17 1.5 0.8ind
61 94 0.013 0.067 3.8 0.009 0.17 1.5 0.8ind
62 56 0.013 0.067 3.8 0.009 0.17 1.5 0.8ind
63 111 0.013 0.067 3.8 0.009 0.17 1.5 0.8ind
Appendix D. Stability analysis of a multi-machine system
D.1. Test system 151

D.1.3 Modelling of photovoltaic systems

Table D.5: Parameters of single-converter equivalent system

Bus Streq [MVA] Vtreqh [kV] Vtreql [kV] Rtreq [pu] Xtreq [pu] ntreq [pu/pu]
1041 130 20 0.6 0.0088 0.0593 1
1042 130 20 0.6 0.0088 0.0593 1
1043 130 20 0.6 0.0088 0.0593 1
1044 130 20 0.6 0.0088 0.0593 1
1045 130 20 0.6 0.0088 0.0593 1
4042 240 20 0.6 0.0088 0.0593 1
4043 240 20 0.6 0.0088 0.0593 1
4046 240 20 0.6 0.0088 0.0593 1
4061 300 20 0.6 0.0088 0.0593 1
4062 300 20 0.6 0.0088 0.0593 1

Table D.5: Parameters of single-converter equivalent system (continued)

Bus Vcoleq [kV] Rcoleq [pu] Xcoleq [pu] Bcoleq [pu]


1041 20 0.016 0.013 0.032
1042 20 0.016 0.013 0.032
1043 20 0.016 0.013 0.032
1044 20 0.016 0.013 0.032
1045 20 0.016 0.013 0.032
4042 20 0.016 0.013 0.032
4043 20 0.016 0.013 0.032
4046 20 0.016 0.013 0.032
4061 20 0.016 0.013 0.032
4062 20 0.016 0.013 0.032
152 Appendix D. Stability analysis of a multi-machine system

Table D.5: Parameters of single-converter equivalent system (continued)

Bus Ssubtr [MVA] Vsubtrh [kV] Vsubtrl [kV] Rsubtr [pu] Xsubtr [pu] nsubtr [pu/pu]
1041 130 130 20 0.003 0.15 1.05
1042 130 130 20 0.003 0.15 1.05
1043 130 130 20 0.003 0.15 1.05
1044 130 130 20 0.003 0.15 1.05
1045 130 130 20 0.003 0.15 1.05
4042 250 400 20 0 0.15 1.05
4043 250 400 20 0 0.15 1.05
4046 250 400 20 0 0.15 1.05
4061 300 400 20 0 0.15 1.05
4062 300 400 20 0 0.15 1.05

Table D.5: Parameters of single-converter equivalent system (continued)

Bus Vitl [kV] Ritl [pu] Xitl [pu] Bitl [pu]


1041 130 0.008 0.03 0.0085
1042 130 0.008 0.03 0.0085
1043 130 0.008 0.03 0.0085
1044 130 0.008 0.03 0.0085
1045 130 0.008 0.03 0.0085
4042 400 0.014 0.112 0.0143
4043 400 0.014 0.112 0.0143
4046 400 0.014 0.112 0.0143
4061 400 0.014 0.112 0.0143
4062 400 0.014 0.112 0.0143
D.1. Test system 153

D.1.4 Operating point

Table D.6: Load consumption

Bus Pexp [MW] Qexp [Mvar] Psim [MW] Qsim [Mvar] Plim [MW] Qlim [Mvar]
1 420 103.7 90 22.2 90 22.2
2 231 49.7 49 10.6 49 10.6
3 182 58.6 39 12.5 39 12.5
4 588 176.4 126 37.8 126 37.8
5 504 133.2 108 28.5 108 28.5
11 140 48.1 30 10.3 30 10.3
12 210 58.6 45 12.5 45 12.5
13 70 24.0 15 5.1 15 5.1
22 196 55.9 42 11.9 42 11.9
31 70 17.2 15 3.7 15 3.7
32 140 27.7 30 5.9 30 5.9
41 378 91.9 81 19.7 81 19.7
42 280 89.1 60 19.1 60 19.1
43 630 178.2 135 38.1 135 38.1
46 490 148.2 105 31.7 105 31.7
47 70 30.8 15 6.6 15 6.6
51 560 180.7 120 38.7 120 38.7
61 350 85.7 75 18.3 75 18.3
62 210 58.6 45 12.5 45 12.5
63 413 185.2 88 39.6 88 39.6
71 300 83.8 – – – –
72 2000 396.1 – – – –

Table D.7: Photovoltaic generation

Bus PPV [MW] QPV [Mvar]


1041 −108 0
1042 −108 0
1043 −108 0
1044 −108 0
1045 −108 0
4042 −270 0
4043 −270 0
4046 −210 0
4061 −210 0
4062 −210 0
154 Appendix D. Stability analysis of a multi-machine system

Table D.8: Shunt compensation

Bus Qsh [Mvar]


1022 −50
1041 −230
1042 −150
1043 −220
1044 −180
1045 −200
4012 100
4041 −50
4043 −170
4046 −70
4051 −100
4071 400
D.2. Case studies 155

D.2 Case studies


D.2.1 Low voltage ride-through and dynamic voltage
support

1.2

1.0

0.8
Voltage [pu]

0.6

0.4

LVRT characteristic
0.2
A1
A2
0.0 A3

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]
Figure D.1: Impact of low voltage ride-through and dynamic voltage support (fault 2)
 Voltage at all points of common coupling

1.2

1.0

0.8
Voltage [pu]

0.6

0.4

WECC criterion
0.2
A1
A2
0.0 A3

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]
Figure D.2: Impact of low voltage ride-through and dynamic voltage support (fault 2)
 Voltage at all load buses
156 Appendix D. Stability analysis of a multi-machine system

0.0
A1
−0.2 A2
A3
Active current [pu]

−0.4

−0.6

−0.8

−1.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure D.3: Impact of low voltage ride-through and dynamic voltage support (fault 2)
 Active current of all photovoltaic systems

0.0

−0.2
Reactive current [pu]

−0.4

−0.6

−0.8
A1
A2
−1.0 A3

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure D.4: Impact of low voltage ride-through and dynamic voltage support (fault 2)
 Reactive current of all photovoltaic systems
D.2. Case studies 157

1.0 A1
A2
A3
0.8

0.6
Slip [pu]

0.4

0.2

0.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure D.5: Impact of low voltage ride-through and dynamic voltage support (fault 2)
 Slip of all small induction motor loads

1.0 A1
A2
A3
0.8

0.6
Slip [pu]

0.4

0.2

0.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0


Time [s]

Figure D.6: Impact of low voltage ride-through and dynamic voltage support (fault 2)
 Slip of all large induction motor loads
158 Appendix D. Stability analysis of a multi-machine system

D.2.2 Active current recovery rate

1.0

0.8
LVRT characteristic
Voltage [pu]

0.6 1.15
1.10
1.05
0.4 1.00
0.95
0.90 B1
0.2
0.85 B2
0.0 0.5 1.0 1.5 2.0 2.5 3.0 A3
0.0 B3

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0


Time [s]

Figure D.7: Impact of active current recovery rate (fault 1)


 Voltage at all points of common coupling

1.0

0.8
WECC criterion
Voltage [pu]

0.6 1.10
1.05
1.00
0.4
0.95
0.90
0.2 B1
0.85
B2
0.0 0.5 1.0 1.5 2.0 2.5 3.0 A3
0.0 B3

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0


Time [s]

Figure D.8: Impact of active current recovery rate (fault 1)


 Voltage at all load buses
D.2. Case studies 159

0.0 B1
B2
A3
−0.2 B3
Active current [pu]

−0.4

−0.6

−0.8

−1.0

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0


Time [s]

Figure D.9: Impact of active current recovery rate (fault 1)


 Active current of all photovoltaic systems

0.0

−0.2
Reactive current [pu]

0.00
−0.4 −0.05
−0.10
−0.15
−0.6 −0.20
−0.25
−0.30
−0.8 −0.35 B1
0.0 0.5 1.0 1.5 2.0 2.5 3.0 B2
A3
−1.0 B3

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0


Time [s]

Figure D.10: Impact of active current recovery rate (fault 1)


 Reactive current of all photovoltaic systems
160 Appendix D. Stability analysis of a multi-machine system

D.2.3 Local voltage control

1.2

1.0

0.8
Voltage [pu]

0.6

0.4

LVRT characteristic
0.2
A3
C1
0.0 C2

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0


Time [s]

Figure D.11: Impact of local voltage control (fault 1)


 Voltage at all points of common coupling

1.2

1.0

0.8
Voltage [pu]

0.6

0.4

0.2 WECC criterion


A3
C1
0.0 C2

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0


Time [s]

Figure D.12: Impact of local voltage control (fault 1)


 Voltage at all load buses
D.2. Case studies 161

0.0
A3
−0.2 C1
C2
Active current [pu]

−0.4

−0.6

−0.8

−1.0

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0


Time [s]

Figure D.13: Impact of local voltage control (fault 1)


 Active current of all photovoltaic systems

1.0 A3
C1
C2
0.5
Reactive current [pu]

0.0

−0.5

−1.0
0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0
Time [s]

Figure D.14: Impact of local voltage control (fault 1)


 Reactive current of all photovoltaic systems
162 Appendix D. Stability analysis of a multi-machine system

1.0

0.8

0.6
Slip [pu]

A3
0.4 C1
C2

0.2

0.0

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0


Time [s]

Figure D.15: Impact of local voltage control (fault 1)


 Slip of all small induction motor loads

1.0

0.8

0.6
Slip [pu]

A3
0.4 C1
C2

0.2

0.0

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0


Time [s]

Figure D.16: Impact of local voltage control (fault 1)


 Slip of all large induction motor loads
D.2. Case studies 163

D.2.4 Plant-level voltage control

1.0

0.8
LVRT characteristic
Voltage [pu]

0.6 1.15
1.10
0.4 1.05
1.00
D1
0.2 0.95
D2
0.0 0.5 1.0 1.5 2.0 2.5 3.0 D3
0.0 D4

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0 20.0
Time [s]

Figure D.17: Impact of plant-level voltage control (fault 1)


 Voltage at all points of common coupling

1.0

0.8
WECC criterion
Voltage [pu]

0.6 1.10
1.05
0.4 1.00
0.95
0.2 0.90 D1
D2
0.0 0.5 1.0 1.5 2.0 2.5 3.0 D3
0.0 D4

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0 20.0
Time [s]

Figure D.18: Impact of plant-level voltage control (fault 1)


 Voltage at all load buses
164 Appendix D. Stability analysis of a multi-machine system

0.0 D1
D2
D3
−0.2 D4
−0.80
Active current [pu]

−0.85
−0.4
−0.90
−0.95
−0.6
−1.00
0.0 0.5 1.0 1.5 2.0 2.5 3.0
−0.8

−1.0

0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0 20.0
Time [s]

Figure D.19: Impact of plant-level voltage control (fault 1)


 Active current of all photovoltaic systems

1.0 D1
0.8 D2
D3
0.6
D4
Reactive current [pu]

0.4
0.2
0.0
−0.2
−0.4
−0.6
−0.8
−1.0
0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0 20.0
Time [s]

Figure D.20: Impact of plant-level voltage control (fault 1)


 Reactive current of all photovoltaic systems
165

Nomenclature

Abbreviations
CCT Critical Clearing Time
DAE Differential Algebraic Equation
DVS Dynamic Voltage Support
EMT Electro-Magnetic Transient
EPRI Electric Power Research Institute
FACTS Flexible Alternating Current Transmission System
FIDVR Fault-Induced Delayed Voltage Recovery
FRT Fault Ride-Through
HVDC High Voltage Direct Current
IM Induction Motor
KL Kullback-Leibler
LVRT Low Voltage Ride-Through
PCC Point of Common Coupling
PDF Probability Density Function
PI Proportional Integral
PLL Phase-Locked Loop
PV PhotoVoltaic
REEC Renewable Energy Electrical Control
REGC Renewable Energy Generator/Converter
REMV Renewable Energy Model Validation
REPC Renewable Energy Plant Control
RMS Root Mean Square
SC Synchronous Condenser
SG Synchronous Generator
VRI Voltage Recovery Index
WECC Western Electricity Coordinating Council
166 NOMENCLATURE

Symbols and units


α Exponential load active power exponent [–]
β Exponential load reactive power exponent [–]
γ Density parameter for probability density function of reference [–]
voltage recovery
Δt Time step [s]
Δt1 Time deviation 1 of Western Electricity Coordinating Council [s]
violation criterion
Δt2 Time deviation 2 of Western Electricity Coordinating Council [s]
violation criterion
Δt3 Time deviation 3 of Western Electricity Coordinating Council [s]
violation criterion
Δt4 Time deviation 4 of Western Electricity Coordinating Council [s]
violation criterion
ΔV Voltage deviation [pu]
η Weighting function [–]
η1+ Rewarding weighting function for voltage limit 1 of Western [–]
Electricity Coordinating Council violation criterion
η2+ Rewarding weighting function for voltage limit 2 of Western [–]
Electricity Coordinating Council violation criterion
η3+ Rewarding weighting function for voltage limit 3 of Western [–]
Electricity Coordinating Council violation criterion
ηj+ Rewarding weighting function for j-th voltage limit [–]
ηji
+
Rewarding weighting function for j-th voltage limit in i-th [–]
sub-interval
η1− Penalising weighting function for voltage limit 1 of Western [–]
Electricity Coordinating Council violation criterion
η2− Penalising weighting function for voltage limit 2 of Western [–]
Electricity Coordinating Council violation criterion
η3− Penalising weighting function for voltage limit 3 of Western [–]
Electricity Coordinating Council violation criterion
ηj− Penalising weighting function for j-th voltage limit [–]

ηji Penalising weighting function for j-th voltage limit in i-th [–]
sub-interval
λ+
1 Rewarding weighting coefficient for voltage limit 1 of Western [–]
Electricity Coordinating Council violation criterion
λ+
2 Rewarding weighting coefficient for voltage limit 2 of Western [–]
Electricity Coordinating Council violation criterion
λ+
3 Rewarding weighting coefficient for voltage limit 3 of Western [–]
Electricity Coordinating Council violation criterion
λ+
j Rewarding weighting coefficient for j-th voltage limit [–]
NOMENCLATURE 167

λ−
1 Penalising weighting coefficient for voltage limit 1 of Western [–]
Electricity Coordinating Council violation criterion
λ−
2 Penalising weighting coefficient for voltage limit 2 of Western [–]
Electricity Coordinating Council violation criterion
λ−
3 Penalising weighting coefficient for voltage limit 3 of Western [–]
Electricity Coordinating Council violation criterion
λ−j Penalising weighting coefficient for j-th voltage limit [–]
σ Threshold for accuracy of weighting function [–]
ϕbus Bus angle [◦ ]
ωimr Induction motor rotor speed [pu]
ωims Induction motor stator speed [pu]
K Kullback-Leibler divergence measure [–]
K2 Kullback-Leibler divergence measure of voltage at bus 2 [–]
K4 Kullback-Leibler divergence measure of voltage at bus 4 [–]
K7 Kullback-Leibler divergence measure of voltage at bus 7 [–]
Kbad Kullback-Leibler divergence measure of bad voltage recovery [–]
Kgood Kullback-Leibler divergence measure of good voltage recovery [–]
Kideal Kullback-Leibler divergence measure of ideal voltage recovery [–]
Kosc Kullback-Leibler divergence measure of oscillatory voltage [–]
recovery
KWECC Kullback-Leibler divergence measure of Western Electricity [–]
Coordinating Council voltage violation criterion
Number of time-domain simulation [–]
P KL Probability density function for Kullback-Leibler divergence [–]
measure
Pbad
KL
Probability density function for Kullback-Leibler divergence [–]
measure of bad voltage recovery
Pgood
KL
Probability density function for Kullback-Leibler divergence [–]
measure of good voltage recovery
PiKL Number of voltage samples in i-th sub-interval for Kullback- [–]
Leibler divergence measure
Pideal
KL
Probability density function for Kullback-Leibler divergence [–]
measure of ideal voltage recovery
Posc
KL
Probability density function for Kullback-Leibler divergence [–]
measure of oscillatory voltage recovery
Pref
KL
Probability density function for Kullback-Leibler divergence [–]
measure of reference voltage recovery
Pref,i
KL
Number of voltage samples in i-th sub-interval of reference voltage [–]
recovery for Kullback-Leibler divergence measure
PWECC
KL
Probability density function for Kullback-Leibler divergence [–]
measure of Western Electricity Coordinating Council voltage
violation criterion
P VRI Probability density function for voltage recovery index [–]
168 NOMENCLATURE

Pbad
VRI
Probability density function for voltage recovery index of bad [–]
voltage recovery
Pgood
VRI
Probability density function for voltage recovery index of good [–]
voltage recovery
PiVRI Number of voltage samples in i-th sub-interval for voltage [–]
recovery index
P VRI+ Rewarding probability density function for voltage recovery [–]
index
P1i
VRI+
Number of rewarding voltage samples for voltage limit 1 in [–]
i-th sub-interval for voltage recovery index
P2i
VRI+
Number of rewarding voltage samples for voltage limit 2 in [–]
i-th sub-interval for voltage recovery index
P3i
VRI+
Number of rewarding voltage samples for voltage limit 3 in [–]
i-th sub-interval for voltage recovery index
PiVRI+ Number of rewarding voltage samples in i-th sub-interval for [–]
voltage recovery index
PjiVRI+ Number of rewarding voltage samples for j-th voltage limit in [–]
i-th sub-interval for voltage recovery index
P VRI− Penalising probability density function for voltage recovery [–]
index
P1i
VRI−
Number of penalising voltage samples for voltage limit 1 in [–]
i-th sub-interval for voltage recovery index
P2i
VRI−
Number of penalising voltage samples for voltage limit 2 in [–]
i-th sub-interval for voltage recovery index
P3i
VRI−
Number of penalising voltage samples for voltage limit 3 in [–]
i-th sub-interval for voltage recovery index
PiVRI− Number of penalising voltage samples in i-th sub-interval for [–]
voltage recovery index
PjiVRI− Number of penalising voltage samples for j-th voltage limit in [–]
i-th sub-interval for voltage recovery index
V Voltage recovery index [–]
V2 Voltage recovery index of voltage at bus 2 [–]
V4 Voltage recovery index of voltage at bus 4 [–]
V7 Voltage recovery index of voltage at bus 7 [–]
Vbad Voltage recovery index of bad voltage recovery [–]
Vgood Voltage recovery index of good voltage recovery [–]
A Total number of voltage samples [–]
brkpt Low voltage power logic breakpoint [pu]
Bcoleq Collector system equivalent susceptance [pu]
Bitl Interconnection transmission line susceptance [pu]
Btl1 Transmission line 1 susceptance [pu]
Btl2 Transmission line 2 susceptance [pu]
NOMENCLATURE 169

dbd Reactive power deadband if RefFlag = 0 or voltage [pu]


deadband if RefFlag = 1
dbd1 Over voltage deadband for reactive current injection [pu]
dbd2 Under voltage deadband for reactive current injection [pu]
dPmax Active power up-ramp limit [pu/s]
dPmin Active power down-ramp limit [pu/s]
Ddn Down regulation droop [pu/pu]
Dup Up regulation droop [pu/pu]
emax Maximum volt/var error [pu]
emin Minimum volt/var error [pu]
fdbd1 Over frequency deadband for governor response [pu]
fdbd2 Under frequency deadband for governor response [pu]
femax Maximum power error in droop regulator [pu]
femin Minimum power error in droop regulator [pu]
Freq Frequency [pu]
Freq ref Frequency reference [pu]
FrqFlag Enable (1) or disable (0) governor response [–]
Him Induction motor inertia constant [s]
Hlim Large induction motor inertia constant [s]
Hsim Small induction motor inertia constant [s]
i Number of voltage sub-interval [–]
Ibranch Branch current for line drop compensation [pu]
Imax Maximum apparent current [pu]
Iolim Current limit for high voltage clamp logic [pu]
Ip Active current [pu]
Ipcmd Desired active current [pu]
Ipcmd  Desired active current before current limitation [pu]
Ipmax Maximum active current [pu]
Ipmin Minimum active current [pu]
Iq Reactive current [pu]
Iqcmd Desired reactive current [pu]
Iqcmd  Desired reactive current before current limitation [pu]
Iqh1 Maximum reactive current injection [pu]
Iqinj Supplementary reactive current injection [pu]
Iql1 Minimum reactive current injection [pu]
Iqmax Maximum reactive current [pu]
Iqmin Minimum reactive current [pu]
Iqrmax Maximum rate-of-change of reactive current [pu/s]
Iqrmin Minimum rate-of-change of reactive current [pu/s]
j Number of voltage limit [–]
Kc Reactive droop if VcmpFlag = 0 [pu]
Khv High voltage clamp logic acceleration factor [pu/pu]
Ki Volt/var regulator integral gain [pu/pu-s]
170 NOMENCLATURE

Kig Droop regulator integral gain [pu/pu-s]


Kp Volt/var regulator proportional gain [pu/pu]
Kpg Droop regulator proportional gain [pu/pu]
Kqi Local reactive power regulator integral gain [pu/pu-s]
Kqp Local reactive power regulator proportional gain [pu/pu]
Kqv Reactive current injection gain [pu]
Kvi Local voltage regulator integral gain [pu/pu-s]
Kvp Local voltage regulator proportional gain [pu/pu]
lvpnt0 Low voltage active current management breakpoint 0 [pu]
lvpnt1 Low voltage active current management breakpoint 1 [pu]
L Total number of voltage sub-intervals [–]
LVPL Active current limit from low voltage power logic [pu]
Lvpl1 Low voltage power logic gain breakpoint [pu/pu]
Lvplsw Enable (1) or disable (0) low voltage power logic [–]
M Total number of voltage limits [–]
n Total number of under voltage protection functions [–]
ndntr Step-down transformer tap ratio [pu/pu]
nsubtr Substation transformer tap ratio [pu/pu]
ntreq Step-up transformer equivalent tap ratio [pu/pu]
N Total number of power system elements [–]
pfaref Inverter initial power factor angle reference from power [◦ ]
flow solution
P0 Exponential load active power under reference voltage [MW]
Pexp Exponential load active power [MW]
Pext External grid active power [MW]
Plim Large induction motor load active power [MW]
PPV Photovoltaic system active power [MW]
PSG Synchronous generator active power [MW]
Psim Small induction motor load active power [MW]
Pbranch Branch active power for plant active power regulation [pu]
Pe Inverter active power [pu]
PF exp Exponential load power factor [–]
PF lim Large induction motor power factor [–]
PF sim Small induction motor power factor [–]
PfFlag Constant reactive power (0) or power factor (1) control [–]
Plant pref Initial branch active power from power flow solution [pu]
Pmax Maximum active power [pu]
Pmax plant Maximum plant active power [pu]
Pmin Minimum active power [pu]
Pmin plant Minimum plant active power [pu]
PqFlag Priority to reactive current (0) or active current (1) [–]
Pref Inverter active power reference from power flow solution [pu]
or from plant control model
NOMENCLATURE 171

PV pen Photovoltaic penetration level [%]


Q0 Exponential load reactive power under reference voltage [Mvar]
Qexp Exponential load reactive power [Mvar]
Qext External grid reactive power [Mvar]
Qlim Large induction motor load reactive power [Mvar]
QPV Photovoltaic system reactive power [Mvar]
Qsh Shunt reactive power [Mvar]
Qsim Small induction motor load reactive power [Mvar]
Qbranch Branch reactive power for plant reactive power regulation [pu]
Qext Inverter reactive power reference from power flow solution [pu]
or from plant control model
QFlag Bypass (0) or engage (1) inner voltage regulator loop [–]
Qgen Inverter reactive power [pu]
Qmax Maximum reactive power if VFlag = 1 [pu]
Qmax plant Maximum plant reactive power [pu]
Qmin Minimum reactive power if VFlag = 1 [pu]
Qmin plant Minimum plant reactive power [pu]
Qref Regulated branch initial reactive power from power flow [pu]
solution
rrpwr Maximum rate-of-change of active current [pu/s]
Rcoleq Collector system equivalent resistance [pu]
Rdntr Step-down transformer resistance [pu]
Rime Induction motor equivalent resistance [pu]
Rimr Induction motor rotor resistance [pu]
Rims Induction motor stator resistance [pu]
Ritl Interconnection transmission line resistance [pu]
Rlimr Large induction motor rotor resistance [pu]
Rlims Large induction motor stator resistance [pu]
Rsimr Small induction motor rotor resistance [pu]
Rsims Small induction motor stator resistance [pu]
Rsubtr Substation transformer resistance [pu]
Rtl1 Transmission line 1 resistance [pu]
Rtl2 Transmission line 2 resistance [pu]
Rtreq Step-up transformer equivalent resistance [pu]
Rc Line drop compensation resistance if VcmpFlag = 1 [pu]
RefFlag Plant-level reactive power (0) or voltage (1) control [–]
sim Induction motor slip [pu]
Sdntr Step-down transformer nominal apparent power [MVA]
Slim Large induction motor nominal apparent power [MVA]
Ssim Small induction motor nominal apparent power [MVA]
Ssubtr Substation transformer nominal apparent power [MVA]
Streq Step-up transformer equivalent nominal apparent power [MVA]
t Time [s]
172 NOMENCLATURE

t0 Initial time [s]


t1 Time limit 1 of Western Electricity Coordinating Council [s]
violation criterion
t2 Time limit 2 of Western Electricity Coordinating Council [s]
violation criterion
tCCT Critical clearing time [s]
tCCT1 Critical clearing time of fault 1 [s]
tCCT2 Critical clearing time of fault 2 [s]
(0)
tCCT Initial critical clearing time [s]
()
tCCT Critical clearing time at ( )-th time-domain simulation [s]
(+1)
tCCT Critical clearing time at ( + 1)-th time-domain simulation [s]
tcl Time at fault clearing instant [s]
tf Final observation time [s]
T0 Induction motor initial torque from power flow solution [pu]
Te Induction motor electrical torque [pu]
Tm Induction motor mechanical torque [pu]
Tfltr Terminal voltage filter time constant [s]
Tft Plant controller reactive power output lead time constant [s]
Tfv Plant controller reactive power output lag time constant [s]
Tgp Inverter active current regulator lag time constant [s]
Tgq Inverter reactive current regulator lag time constant [s]
Tiq Reactive current regulator lag time constant [s]
Tlag Plant controller active power output lag time constant [s]
Tp Active power filter time constant [s]
Tpord Inverter power order lag time constant [s]
Tqfltr Reactive power filter time constant [s]
Trv Terminal bus voltage filter time constant [s]
Tvfltr Voltage filter time constant [s]
va Instantaneous voltage of phase a [pu]
vb Instantaneous voltage of phase b [pu]
vc Instantaneous voltage of phase c [pu]
V0 Exponential load reference voltage [pu]
V1 Voltage limit 1 of Western Electricity Coordinating Council [pu]
violation criterion
V2 Voltage limit 2 of Western Electricity Coordinating Council [pu]
violation criterion
V3 Voltage limit 3 of Western Electricity Coordinating Council [pu]
violation criterion
Va Voltage of phase a [pu]
Vb Voltage of phase b [pu]
Vbad Bad voltage recovery [pu]
Vbus Bus voltage [pu]
Vc Voltage of phase c [pu]
NOMENCLATURE 173

Vcoleq Collector system equivalent nominal voltage [kV]


Vdntrh Step-down transformer voltage on high voltage side [kV]
Vdntrl Step-down transformer voltage on low voltage side [kV]
Vgood Good voltage recovery [pu]
Vi Voltage of i-th sub-interval [pu]
Videal Ideal voltage recovery [pu]
Vim Induction motor terminal voltage [pu]
Vitl Interconnection transmission line nominal voltage [kV]
Vmax Maximum voltage [pu]
Vmin Minimum voltage [pu]
Vosc Oscillatory voltage recovery [pu]
Vpre Pre-disturbance voltage [pu]
Vsubtrh Substation transformer voltage on high voltage side [kV]
Vsubtrl Substation transformer voltage on low voltage side [kV]
Vtl1 Transmission line 1 nominal voltage [kV]
Vtl2 Transmission line 2 nominal voltage [kV]
Vtreqh Step-up transformer equivalent voltage on high voltage side [kV]
Vtreql Step-up transformer equivalent voltage on low voltage side [kV]
VWECC Voltage of Western Electricity Coordinating Council violation [pu]
criterion
VcmpFlag Reactive droop (0) or line drop compensation (1) [–]
Vdip Low voltage condition trigger [pu]
VFlag Local reactive power (0) or voltage (1) control [–]
Vfrz Voltage for freezing volt/var regulator integrator [pu]
Vmax Maximum voltage at inverter terminal bus [pu]
Vmin Minimum voltage at inverter terminal bus [pu]
Volim Voltage limit for high voltage clamp logic [pu]
Voltage dip Low/high voltage ride-through condition [–]
Vref Regulated bus initial voltage from power flow solution [pu]
Vref0 Reference voltage for reactive current injection from power [pu]
flow solution
Vreg Regulated bus voltage from network solution [pu]
Vt Raw terminal voltage from network solution [pu]
Vt filt Filtered terminal voltage [pu]
Vup High voltage condition trigger [pu]
W Normalising factor for probability density function of [–]
reference voltage recovery
Xcoleq Collector system equivalent reactance [pu]
Xdntr Step-down transformer reactance [pu]
Xime Induction motor equivalent reactance [pu]
Ximm Induction motor magnetising reactance [pu]
Ximr Induction motor rotor leakage reactance [pu]
Xims Induction motor stator leakage reactance [pu]
174 NOMENCLATURE

Xitl Interconnection transmission line reactance [pu]


Xlimm Large induction motor magnetising reactance [pu]
Xlimr Large induction motor rotor leakage reactance [pu]
Xlims Large induction motor stator leakage reactance [pu]
Xsimm Small induction motor magnetising reactance [pu]
Xsimr Small induction motor rotor leakage reactance [pu]
Xsims Small induction motor stator leakage reactance [pu]
Xsubtr Substation transformer reactance [pu]
Xtl1 Transmission line 1 reactance [pu]
Xtl2 Transmission line 2 reactance [pu]
Xtreq Step-up transformer equivalent reactance [pu]
Xc Line drop compensation reactance if VcmpFlag = 1 [pu]
zerox Low voltage power logic zero crossing [pu]
175

Bibliography

[1] SMA Solar Technology AG, SMA Supplies System Technology for One of
Australia’s Largest PV Power Plants, Press Release, 2018.
[2] T. Ackermann, G. Andersson, and L. Söder, “Distributed Generation: A
Definition”, Electric Power Systems Research, vol. 57, no. 3, pp. 195 –204,
2001.
[3] Ecofys Germany GmbH, “Klimaschutz durch Reduzierung des Energiebe-
darfs für Gebäudekühlung”, Tech. Rep., 2011.
[4] Bundesverband Wärmepumpe (bwp) e. V., Marktzahlen 2017: Wärmepumpen-
Absatz wächst deutlich, Press Release, 2018.
[5] IEEE Standards Association, Standard for Interconnecting Distributed Re-
sources with Electric Power Systems, IEEE Std. 1547, 2014.
[6] European Network of Transmission System Operators for Electricity (ENTSO-
E), Network Code on Requirements for Grid Connection Applicable to All
Generators (NC-RfG), EU Regulation 2016/631, 2016.
[7] Verband der Elektrotechnik, Elektronik und Informationstechnik (VDE),
Guideline VDE-AR-N 4120, “Technical Requirements for the Connection
and Operation of Customer Installations to the High Voltage Network”, Tech.
Rep., 2015.
[8] E. Quitmann and E. Erdmann, “Power System Needs – How Grid Codes
Should Look Ahead”, IET Renewable Power Generation, vol. 9, no. 1, pp. 3–
9, 2015.
[9] J. C. Boemer, “On Stability of Sustainable Power Systems: Network Fault
Response of Transmission Systems with Very High Penetration of Distributed
Generation”, PhD thesis, Delft University of Technology, 2016.
176 BIBLIOGRAPHY

[10] Deutsche Energie-Agentur (dena) GmbH, “Energiewirtschaftliche Planung


für die Netzintegration von Windenergie in Deutschland an Land und Off-
shore bis zum Jahr 2020”, Tech. Rep., 2005.
[11] J. Fortmann, R. Pfeiffer, E. Haesen, F. van Hulle, F. Martin, H. Urdal, and S.
Wachtel, “Fault Ride-Through Requirements for Wind Power Plants in the
ENTSO-E Network Code on Requirements for Generators”, IET Renewable
Power Generation, vol. 9, no. 1, pp. 18–24, 2015.
[12] J. C. Boemer, K. Burges, P. Zolotarev, J. Lehner, P. Wajant, M. Fürst,
R. Brohm, and T. Kumm, “Overview of German Grid Issues and Retrofit
of Photovoltaic Power Plants in Germany for the Prevention of Frequency
Stability Problems in Abnormal System Conditions of the ENTSO-E Region
Continental Europe”, in 1st International Workshop on Integration of Solar
Power Into Power Systems, Aarhus, 2011.
[13] Verband der Elektrotechnik, Elektronik und Informationstechnik (VDE),
Forum Netztechnik/Netzbetrieb (FNN), “Weiterentwicklung des Verhaltens
von Erzeugungsanlagen am Niederspannungsnetz im Fehlerfall”, Tech. Rep.,
2014.
[14] North American Electric Reliability Corporation (NERC), “1200 MW Fault
Induced Solar Photovoltaic Resource Interruption Disturbance Report – South-
ern California 8/16/2016 Event”, Tech. Rep., 2017.
[15] Australian Energy Market Operator (AEMO), “Black System South Aus-
tralia 28 September 2016”, Tech. Rep., 2017.
[16] J. Ma, D. Zhao, M. Qian, L. Zhu, and H. Geng, “Modelling and Validating
Photovoltaic Power Inverter Model for Power System Stability Analysis”,
The Journal of Engineering, vol. 2017, no. 13, pp. 1605–1609, 2017.
[17] Bundesnetzagentur Strommarktdaten (SMARD), Electricity Generation and
Consumption in Germany, Data from May 1, 2018.
[18] R. Shah, N. Mithulananthan, R. C. Bansal, and V. K. Ramachandaramurthy,
“A Review of Key Power System Stability Challenges for Large-Scale PV
Integrationn”, Renewable and Sustainable Energy Reviews, vol. 41, pp. 1423
–1436, 2015.
[19] B. Weise, “Impact of K-Factor and Active Current Reduction During Fault
Ride-Through of Generating Units Connected via Voltage-Sourced Convert-
ers on Power System Stability”, IET Renewable Power Generation, vol. 9,
no. 1, pp. 25–36, 2015.
[20] K. Kawabe and K. Tanaka, “Impact of Dynamic Behavior of Photovoltaic
Power Generation Systems on Short-Term Voltage Stability”, IEEE Trans-
actions on Power Systems, vol. 30, no. 6, pp. 3416–3424, 2015.
BIBLIOGRAPHY 177

[21] K. Kawabe, Y. Ota, A. Yokoyama, and K. Tanaka, “Novel Dynamic Voltage


Support Capability of Photovoltaic Systems for Improvement of Short-Term
Voltage Stability in Power Systems”, IEEE Transactions on Power Systems,
vol. 32, no. 3, pp. 1796–1804, 2017.
[22] T. Neumann and I. Erlich, “Modelling and Control of Photovoltaic Inverter
Systems with Respect to German Grid Code Requirements”, in IEEE Power
& Energy Society General Meeting, San Diego, 2012.
[23] S. Eftekharnejad, “The Impact of Increased Penetration of Photovoltaic Gen-
eration on Smart Grids”, PhD thesis, Arizona State University, 2012.
[24] D. Premm, O. Glitza, T. Fawzy, B. Engel, and G. Bettenwort, “Grid In-
tegration of Photovoltaic Plants – A Generic Description of PV Plants for
Grid Studies”, in 21st International Conference on Electricity Distribution
(CIRED), Frankfurt, 2011.
[25] V. Sakschewski, D. Premm, H. Knopf, and B. Engel, “New System Services
of Solar Inverters Connected to the Low and Medium Voltage Grid”, in
International ETG Congress, Würzburg, 2011.
[26] S. Dasgupta, M. Paramasivam, U. Vaidya, and V. Ajjarapu, “Entropy-Based
Metric for Characterization of Delayed Voltage Recovery”, IEEE Transac-
tions on Power Systems, vol. 30, no. 5, pp. 2460–2468, 2015.
[27] M. Paramasivam, S. Dasgupta, V. Ajjarapu, and U. Vaidya, “Contingency
Analysis and Identification of Dynamic Voltage Control Areas”, IEEE Trans-
actions on Power Systems, vol. 30, no. 6, pp. 2974–2983, 2015.
[28] A. Marinopoulos, F. Papandrea, M. Reza, S. Norrga, F. Spertino, and R.
Napoli, “Grid Integration Aspects of Large Solar PV Installations: LVRT
Capability and Reactive Power/Voltage Support Requirements”, in IEEE
PowerTech, Trondheim, 2011.
[29] M. Morjaria, D. Anichkov, V. Chadliev, and S. Soni, “A Grid-Friendly Plant:
The Role of Utility-Scale Photovoltaic Plants in Grid Stability and Reliabil-
ity”, IEEE Power and Energy Magazine, vol. 12, no. 3, pp. 87–95, 2014.
[30] S. Soni, G. G. Karady, M. Morjaria, and V. Chadliev, “Comparison of Full
and Reduced Scale Solar PV Plant Models in Multi-Machine Power Sys-
tems”, in IEEE PES T&D Conference and Exposition, Chicago, 2014.
[31] R. T. Elliott, A. Ellis, P. Pourbeik, J. J. Sanchez-Gasca, J. Senthil, and
J. D. Weber, “Generic Photovoltaic System Models for WECC – A Status
Report”, in IEEE Power & Energy Society General Meeting, Denver, 2015.
[32] F. Andrén, B. Bletterie, S. Kadam, P. Kotsampopoulos, and C. Bucher, “On
the Stability of Local Voltage Control in Distribution Networks With a High
Penetration of Inverter-Based Generation”, IEEE Transactions on Industrial
Electronics, vol. 62, no. 4, pp. 2519–2529, 2015.
178 BIBLIOGRAPHY

[33] A. Yazdani, A. R. Di Fazio, H. Ghoddami, M. Russo, M. Kazerani, J. Jatske-


vich, K. Strunz, S. Leva, and J. A. Martinez, “Modeling Guidelines and
a Benchmark for Power System Simulation Studies of Three-Phase Single-
Stage Photovoltaic Systems”, IEEE Transactions on Power Delivery, vol.
26, no. 2, pp. 1247–1264, 2011.
[34] D. Fetzer, “Development of a MATLAB/Simulink Framework for Phasor-
Based Power System Simulation and Component Modeling Based on State
Machines”, PhD thesis, University of Kassel, 2018.
[35] P. Pourbeik, S. Soni, A. Gaikwad, and V. Chadliev, “Providing Primary Fre-
quency Response from Photovoltaic Power Plants”, in CIGRE Symposium,
Dublin, 2017.
[36] P. Aristidou, “Time-Domain Simulation of Large Electric Power Systems Us-
ing Domain-Decomposition and Parallel Processing Methods”, PhD thesis,
University of Liège, 2015.
[37] E. J. Coster, “Distribution Grid Operation Including Distributed Gener-
ation: Impact on Grid Protection and the Consequences of Fault Ride-
Through Behavior”, PhD thesis, Eindhoven University of Technology, 2010.
[38] B. Tamimi, C. Cañizares, and K. Bhattacharya, “System Stability Impact
of Large-Scale and Distributed Solar Photovoltaic Generation: The Case of
Ontario, Canada”, IEEE Transactions on Sustainable Energy, vol. 4, no. 3,
pp. 680–688, 2013.
[39] S. Laudahn, “Dynamisches Verhalten von wechselrichterbasierten Erzeugungs-
anlagen im Kontext eines sicheren und stabilen Netzbetriebs”, PhD thesis,
University of Braunschweig, 2017.
[40] Á. Ortega, “Dynamic Modeling, Control and Stability Analysis of VSC-based
Energy Storage Systems”, PhD thesis, University College Dublin, 2017.
[41] A. Ishchenko, “Dynamics and Stability of Distribution Networks with Dis-
persed Generation”, PhD thesis, Eindhoven University of Technology, 2008.
[42] M. Reza, “Stability Analysis of Transmission Systems with High Penetra-
tion of Distributed Generation”, PhD thesis, Delft University of Technology,
2006.
[43] E. Vittal, “The Impact of Reactive Power From Wind Generation on Power
System Stability”, PhD thesis, University College Dublin, 2011.
[44] G. Chaspierre, P. Panciatici, and T. Van Cutsem, “Dynamic Equivalent of a
Distribution Grid Hosting Dispersed Photovoltaic Units”, in IREP’ X Bulk
Power System Dynamics and Control Symposium, Espinho, 2017.
[45] V. Vittal and R. Ayyanar, Grid Integration and Dynamic Impact of Wind
Energy. Springer, 2013.
BIBLIOGRAPHY 179

[46] J. Hossain and H. R. Pota, Robust Control for Grid Voltage Stability: High
Penetration of Renewable Energy. Springer, 2014.
[47] I. Green, “CAISO Experience with Impact of High Penetration of Renew-
able Resources on Short-Term Voltage Stability”, in IEEE Power & Energy
Society General Meeting, Denver, 2015.
[48] J. N. Notholt Vergara, “Fault Ride-Through Capabilities of Inverter-Based
Distributed Generation Connected to Low and Medium Voltage Distribution
Networks”, PhD thesis, University of Kassel, 2009.
[49] P. Kundur, J. Paserba, V. Ajjarapu, G. Andersson, A. Bose, C. Cañizares, N.
Hatziargyriou, D. Hill, A. Stankovic, C. Taylor, T. Van Cutsem, and V. Vit-
tal, “Definition and Classification of Power System Stability IEEE/CIGRE
Joint Task Force on Stability Terms and Definitions”, IEEE Transactions on
Power Systems, vol. 19, no. 3, pp. 1387–1401, 2004.
[50] IEEE Task Force on Stability Definitions and Characterization of Dynamic
Behavior in Systems with High Penetration of Power Electronic Interfaced
Technologies, Ongoing Activities.
[51] C. W. Taylor, Power System Voltage Stability. McGraw Hill, 1994.
[52] T. Van Cutsem and C. Vournas, Voltage Stability of Electric Power Systems.
Springer, 2008.
[53] K. Kawabe and K. Tanaka, “Analytical Method for Short-Term Voltage Sta-
bility Using the Stability Boundary in the P–V Plane”, IEEE Transactions
on Power Systems, vol. 29, no. 6, pp. 3041–3047, 2014.
[54] S. Dasgupta, M. Paramasivam, U. Vaidya, and V. Ajjarapu, “Real-Time
Monitoring of Short-Term Voltage Stability Using PMU Data”, IEEE Trans-
actions on Power Systems, vol. 28, no. 4, pp. 3702–3711, 2013.
[55] A. A. Fouad and V. Vittal, Power System Transient Stability Analysis Using
the Transient Energy Function Method. Prentice Hall, 1992.
[56] M. Pavella, D. Ernst, and D. Ruiz-Vega, Transient Stability of Power Systems
– A Unified Approach to Assessment and Control. Springer, 2000.
[57] North American Electric Reliability Corporation (NERC), “Assessment of
Fault-Induced Delayed Voltage Recovery (FIDVR) Simulations”, Tech. Rep.,
2017.
[58] S. M. Halpin, K. A. Harley, R. A. Jones, and L. Y. Taylor, “Slope-Permissive
Under-Voltage Load Shed Relay for Delayed Voltage Recovery Mitigation”,
IEEE Transactions on Power Systems, vol. 23, no. 3, pp. 1211–1216, 2008.
[59] H. Bai and V. Ajjarapu, “A Novel Online Load Shedding Strategy for Mit-
igating Fault-Induced Delayed Voltage Recovery”, IEEE Transactions on
Power Systems, vol. 26, no. 1, pp. 294–304, 2011.
180 BIBLIOGRAPHY

[60] S. M. Halpin, R. A. Jones, and L. Y. Taylor, “The MVA-Volt Index: A


Screening Tool for Predicting Fault-Induced Low Voltage Problems on Bulk
Transmission Systems”, IEEE Transactions on Power Systems, vol. 23, no.
3, pp. 1205–1210, 2008.
[61] S. Wildenhues, J. L. Rueda, and I. Erlich, “Optimal Allocation and Sizing
of Dynamic Var Sources Using Heuristic Optimization”, IEEE Transactions
on Power Systems, vol. 30, no. 5, pp. 2538–2546, 2015.
[62] IEEE Committee, “Voltage Stability of Power Systems: Concepts, Analytical
Tools, and Industry Experience”, Tech. Rep., 1990.
[63] J. Machowski, J. W. Bialek, and J. R. Bumby, Power System Dynamics:
Stability and Control. John Wiley & Sons, 2012.
[64] P. Kundur, Power System Stability and Control. McGraw Hill, 1994.
[65] F. Milano, Power System Modelling and Scripting. Springer, 2010.
[66] O. Ruhle and F. Balasin, “Simulations of Power System Dynamic Phenom-
ena”, in IEEE PowerTech, Bucharest, 2009.
[67] K. K. Kaberere, K. A. Folly, and A. I. Petroianu, “Assessment of Commer-
cially Available Software Tools for Transient Stability: Experience Gained
in an Academic Environment”, in IEEE 7th Africon Conference in Africa,
2004.
[68] K. K. Kaberere, K. A. Folly, M. Ntombela, and A. I. Petroianu, “Compar-
ative Analysis and Numerical Validation of Industrial-Grade Power System
Simulation Tools: Application to Small-Signal Stability”, in 15th Power Sys-
tems Computation Conference (PSCC), Liège, 2005.
[69] DIgSILENT GmbH, DIgSILENT PowerFactory, Version 2018.
[70] H. W. Dommel, “Digital Computer Solution of Electromagnetic Transients in
Single- and Multiphase Networks”, IEEE Transactions on Power Apparatus
and Systems, vol. PAS-88, no. 4, pp. 388–399, 1969.
[71] F. Plumier, P. Aristidou, C. Geuzaine, and T. Van Cutsem, “Co-Simulation
of Electromagnetic Transients and Phasor Models: A Relaxation Approach”,
IEEE Transactions on Power Delivery, vol. 31, no. 5, pp. 2360–2369, 2016.
[72] Q. Huang and V. Vittal, “Application of Electromagnetic Transient–Transient
Stability Hybrid Simulation to FIDVR Study”, IEEE Transactions on Power
Systems, vol. 31, no. 4, pp. 2634–2646, 2016.
[73] B. Stott, “Power System Dynamic Response Calculations”, Proceedings of
the IEEE, vol. 67, no. 2, pp. 219–241, 1979.
[74] G. W. Stagg and A. H. El-Abiad, Computer Methods in Power System Anal-
ysis. McGraw Hill, 1968.
BIBLIOGRAPHY 181

[75] IEEE Task Force on Terms and Definitions, “Proposed Terms and Definitions
for Power System Stability”, IEEE Transactions on Power Apparatus and
Systems, vol. PAS-101, no. 7, pp. 1894–1898, 1982.
[76] Z. Wang, X. Wang, and C. Y. Chung, “An Analytical Method for Calculating
Critical Voltage Sag Clearance Time of Induction Motors”, IEEE Transac-
tions on Power Delivery, vol. 27, no. 4, pp. 2412–2414, 2012.
[77] S. Kullback and R. A. Leibler, “On Information and Sufficiency”, Annals of
Mathematical Statistics, vol. 22, no. 1, pp. 79–86, 1951.
[78] L. K. Nash, Elements of Classical and Statistical Thermodynamics. Addison-
Wesley, 1968.
[79] T. M. Cover and J. A. Thomas, Elements of Information Theory. Springer,
1991.
[80] D. J. Shoup, J. J. Paserba, and C. W. Taylor, “A Survey of Current Practices
for Transient Voltage Dip/Sag Criteria Related to Power System Stability”,
in IEEE PES Power Systems Conference and Exposition, New York, 2004.
[81] M. Paramasivam, A. Salloum, V. Ajjarapu, V. Vittal, N. B. Bhatt, and S.
Liu, “Dynamic Optimization Based Reactive Power Planning to Mitigate
Slow Voltage Recovery and Short-Term Voltage Instability”, IEEE Transac-
tions on Power Systems, vol. 28, no. 4, pp. 3865–3873, 2013.
[82] Western Electricity Coordinating Council (WECC) Renewable Energy Mod-
eling Task Force, “WECC Solar PV Dynamic Model Specification”, Tech.
Rep., 2012.
[83] Western Electricity Coordinating Council (WECC) Renewable Energy Mod-
eling Task Force, “WECC Solar Plant Dynamic Modeling Guidelines”, Tech.
Rep., 2014.
[84] Western Electricity Coordinating Council (WECC) Renewable Energy Mod-
eling Task Force, “WECC Second Generation Wind Turbine Models”, Tech.
Rep., 2014.
[85] National Renewable Energy Laboratory (NREL), “User Guide for PV Dy-
namic Model Simulation Written on PSCAD Platform”, Tech. Rep., 2014.
[86] Western Electricity Coordinating Council (WECC) Renewable Energy Mod-
eling Task Force, “WECC Central Station Photovoltaic Power Plant Model
Validation Guideline”, Tech. Rep., 2015.
[87] Vereinigung Deutscher Elektrizitätswerke (VDEW), “Eigenerzeugungsanla-
gen am Mittelspannungsnetz – Richtlinie für Anschluss und Parallelbetrieb
von Eigenerzeugungsanlagen am Mittelspannungsnetz”, Tech. Rep., 1998.
[88] Vereinigung Deutscher Elektrizitätswerke (VDEW), “Erzeugungsanlagen am
Niederspannungsnetz – Richtlinie für Anschluss und Parallelbetrieb von Eigen-
erzeugungsanlagen am Niederspannungsnetz”, Tech. Rep., 2001.
182 BIBLIOGRAPHY

[89] RWE Germany AG, “Anschlusspraxis von Erzeugungsanlagen hinsichtlich


einer dynamischen Netzstützung”, Tech. Rep., 2014.
[90] National Grid, “Guidance Notes – Power Park Modules”, Tech. Rep., 2012.
[91] Western Electricity Coordinating Council (WECC) Renewable Energy Mod-
eling Task Force, “WECC PV Plant Power Flow Modeling Guidelines”, Tech.
Rep., 2010.
[92] M. H. J. Bollen, Understanding Power Quality Problems. John Wiley & Sons,
2000.
[93] Electric Power Research Institute (EPRI), Renewable Energy Model Valida-
tion (REMV), Version 2.0.
[94] Electric Power Research Institute (EPRI), “Generic Models and Model Val-
idation for Wind and Solar PV Generation: Technical Update”, Tech. Rep.,
2011.
[95] P. Pourbeik, J. J. Sanchez-Gasca, J. Senthil, J. D. Weber, P. S. Zadehkhost,
Y. Kazachkov, S. Tacke, J. Wen, and A. Ellis, “Generic Dynamic Models for
Modeling Wind Power Plants and Other Renewable Technologies in Large-
Scale Power System Studies”, IEEE Transactions on Energy Conversion,
vol. 32, no. 3, pp. 1108–1116, 2017.
[96] J. H. Chow, Power System Coherency and Model Reduction. Springer, 2013.
[97] R. Leelaruji and L. Vanfretti, “Detailed Modelling, Implementation and Sim-
ulation of an All-in-One Stability Test System Including Power System Pro-
tective Devices”, Simulation Modelling Practice and Theory, vol. 23, pp. 36–
59, 2012.
[98] IEEE Task Force on Test Systems for Voltage Stability and Security As-
sessment, “Test Systems for Voltage Stability Analysis and Security Assess-
ment”, Tech. Rep., 2015.
[99] K. Morison, H. Hamadani, and L. Wang, “Practical Issues in Load Modeling
for Voltage Stability Studies”, in IEEE Power & Energy Society General
Meeting, Toronto, 2003.
[100] IEEE Task Force on Load Representation for Dynamic Performance, “Load
Representation for Dynamic Performance Analysis (of Power Systems)”,
IEEE Transactions on Power Systems, vol. 8, no. 2, pp. 472–482, 1993.
[101] IEEE Task Force on Load Representation for Dynamic Performance, “Stan-
dard Load Models for Power Flow and Dynamic Performance Simulation”,
IEEE Transactions on Power Systems, vol. 10, no. 3, pp. 1302–1313, 1995.
[102] A. Arif, Z. Wang, J. Wang, B. Mather, H. Bashualdo, and D. Zhao, “Load
Modeling – A Review”, IEEE Transactions on Smart Grid, vol. 9, no. 6,
pp. 5986–5999, 2018.
BIBLIOGRAPHY 183

[103] J. V. Milanović, K. Yamashita, S. Martı́nez Villanueva, S. Ž. Djokić, and


L. M. Korunović, “International Industry Practice on Power System Load
Modeling”, IEEE Transactions on Power Systems, vol. 28, no. 3, pp. 3038–
3046, 2013.
[104] P. Pourbeik, D. Wang, and K. Hoang, “Load Modeling in Voltage Stability
Studies”, in IEEE Power & Energy Society General Meeting, San Francisco,
2005.
[105] N. Jelani, M. Molinas, and S. Bolognani, “Reactive Power Ancillary Service
by Constant Power Loads in Distributed AC Systems”, IEEE Transactions
on Power Delivery, vol. 28, no. 2, pp. 920–927, 2013.
[106] J. Brochu, C. Larose, and R. Gagnon, “Generic Equivalent Collector System
Parameters for Large Wind Power Plants”, IEEE Transactions on Energy
Conversion, vol. 26, no. 2, pp. 542–549, 2011.
[107] O. Samuelsson, L. Lindgren, and B. Eliasson, “Simulated Power System
Restoration”, in 43rd International Universities Power Engineering Confer-
ence, Padova, 2008.
[108] T. Van Cutsem, “The Nordic Test System for Voltage Stability Assessment”,
in IEEE Power & Energy Society General Meeting, Presentation at panel
session, Washington D.C., 2014.
[109] L. Robitzky, D. Mayorga Gonzalez, C. Kittl, C. Strunck, J. Zwartscholten,
S. C. Müller, U. Häger, J. Myrzik, and C. Rehtanz, “Impact of Active Distri-
bution Networks on Voltage Stability of Electric Power Systems”, in IREP’
X Bulk Power System Dynamics and Control Symposium, Espinho, 2017.
[110] P. Aristidou, G. Valverde, and T. Van Cutsem, “Contribution of Distribution
Network Control to Voltage Stability: A Case Study”, IEEE Transactions
on Smart Grid, vol. 8, no. 1, pp. 106–116, 2017.
[111] D. Fabozzi, S. Weigel, B. Weise, and F. Villella, “Semi-Implicit Formulation
of Proportional-Integral Controller Block with Non-Windup Limiter Accord-
ing to IEEE Standard 421.5-2016”, in IREP’ X Bulk Power System Dynamics
and Control Symposium, Espinho, 2017.
ISBN 978-3-7376-0716-2

Vol 9
ISBN 978-3-7376-0716-2

Energy Management and Power System Operation Vol 9


9 783737 607162

9 783737 607162

Photovoltaic Systems in Power System Dynamic Studies


Gustav Lammert

Modelling, Control and Stability Analysis of


This thesis investigates the impact of: i) the low voltage ride-through Modelling, Control and
and dynamic voltage support capability; ii) the active current recov-
ery rate; iii) the local voltage control; and iv) the plant-level voltage Stability Analysis of
Photovoltaic Systems in
control of large-scale photovoltaic systems on short-term volt-
age stability and fault-induced delayed voltage recovery as well
as transient and frequency stability. The power system dynamic
performance is analysed using state-of-the-art methods, such as
Power System Dynamic Studies
phasor mode time-domain simulations and the calculation of the
critical clearing time that determines the stability margin. Moreo-
ver, the recently developed Kullback-Leibler divergence measure is
applied to assess the quality of the voltage recovery. Drawbacks of
this metric are outlined and a novel metric, the so-called voltage re-
covery index, is defined that quantifies the delayed voltage recovery
more systematically. The studies are performed with a generic pho-
tovoltaic system model and typical model parameters are used that

Gustav Lammert
were determined in collaboration with a manufacturer. The stability
analysis is performed in DIgSILENT PowerFactory using: i) a one-load
infinite-bus system; and ii) an IEEE multi-machine voltage stability
test system, namely the Nordic test system. The results show that
with the adequate control of photovoltaic systems, power system
dynamic performance can be significantly improved.

You might also like