(IET Energy Engineering 2) Nand Kishor, Jesus Fraile-Ardunuy - Modeling and Dynamic Behaviour of Hydropower Plants-The Institution of Engineering and Technology (2017)

Download as pdf or txt
Download as pdf or txt
You are on page 1of 280

IET ENERGY ENGINEERING SERIES 100

Modeling and Dynamic


Behaviour of Hydropower
Plants
Other volumes in this series:
Volume 1 Power Circuit Breaker Theory and Design C.H. Flurscheim (Editor)
Volume 4 Industrial Microwave Heating A.C. Metaxas and R.J. Meredith
Volume 7 Insulators for High Voltages J.S.T. Looms
Volume 8 Variable Frequency AC Motor Drive Systems D. Finney
Volume 10 SF6 Switchgear H.M. Ryan and G.R. Jones
Volume 11 Conduction and Induction Heating E.J. Davies
Volume 13 Statistical Techniques for High Voltage Engineering W. Hauschild and W. Mosch
Volume 14 Uninterruptible Power Supplies J. Platts and J.D. St Aubyn (Editors)
Volume 15 Digital Protection for Power Systems A.T. Johns and S.K. Salman
Volume 16 Electricity Economics and Planning T.W. Berrie
Volume 18 Vacuum Switchgear A. Greenwood
Volume 19 Electrical Safety: A guide to causes and prevention of hazards J. Maxwell Adams
Volume 21 Electricity Distribution Network Design, 2nd Edition E. Lakervi and E.J. Holmes
Volume 22 Artificial Intelligence Techniques in Power Systems K. Warwick, A.O. Ekwue and R. Aggarwal (Editors)
Volume 24 Power System Commissioning and Maintenance Practice K. Harker
Volume 25 Engineers’ Handbook of Industrial Microwave Heating R.J. Meredith
Volume 26 Small Electric Motors H. Moczala et al.
Volume 27 AC–DC Power System Analysis J. Arrillaga and B.C. Smith
Volume 29 High Voltage Direct Current Transmission, 2nd Edition J. Arrillaga
Volume 30 Flexible AC Transmission Systems (FACTS) Y-H. Song (Editor)
Volume 31 Embedded Generation N. Jenkins et al.
Volume 32 High Voltage Engineering and Testing, 2nd Edition H.M. Ryan (Editor)
Volume 33 Overvoltage Protection of Low-Voltage Systems, Revised Edition P. Hasse
Volume 36 Voltage Quality in Electrical Power Systems J. Schlabbach et al.
Volume 37 Electrical Steels for Rotating Machines P. Beckley
Volume 38 The Electric Car: Development and Future of Battery, Hybrid and Fuel-Cell Cars M. Westbrook
Volume 39 Power Systems Electromagnetic Transients Simulation J. Arrillaga and N. Watson
Volume 40 Advances in High Voltage Engineering M. Haddad and D. Warne
Volume 41 Electrical Operation of Electrostatic Precipitators K. Parker
Volume 43 Thermal Power Plant Simulation and Control D. Flynn
Volume 44 Economic Evaluation of Projects in the Electricity Supply Industry H. Khatib
Volume 45 Propulsion Systems for Hybrid Vehicles J. Miller
Volume 46 Distribution Switchgear S. Stewart
Volume 47 Protection of Electricity Distribution Networks, 2nd Edition J. Gers and E. Holmes
Volume 48 Wood Pole Overhead Lines B. Wareing
Volume 49 Electric Fuses, 3rd Edition A. Wright and G. Newbery
Volume 50 Wind Power Integration: Connection and System Operational Aspects B. Fox et al.
Volume 51 Short Circuit Currents J. Schlabbach
Volume 52 Nuclear Power J. Wood
Volume 53 Condition Assessment of High Voltage Insulation in Power System Equipment R.E. James and Q. Su
Volume 55 Local Energy: Distributed Generation of Heat and Power J. Wood
Volume 56 Condition Monitoring of Rotating Electrical Machines P. Tavner, L. Ran, J. Penman and H. Sedding
Volume 57 The Control Techniques Drives and Controls Handbook, 2nd Edition B. Drury
Volume 58 Lightning Protection V. Cooray (Editor)
Volume 59 Ultracapacitor Applications J.M. Miller
Volume 62 Lightning Electromagnetics V. Cooray
Volume 63 Energy Storage for Power Systems, 2nd Edition A. Ter-Gazarian
Volume 65 Protection of Electricity Distribution Networks, 3rd Edition J. Gers
Volume 66 High Voltage Engineering Testing, 3rd Edition H. Ryan (Editor)
Volume 67 Multicore Simulation of Power System Transients F.M. Uriate
Volume 68 Distribution System Analysis and Automation J. Gers
Volume 69 The Lightening Flash, 2nd Edition V. Cooray (Editor)
Volume 70 Economic Evaluation of Projects in the Electricity Supply Industry, 3rd Edition H. Khatib
Volume 72 Control Circuits in Power Electronics: Practical Issues in Design and Implementation M. Castilla (Editor)
Volume 73 Wide Area Monitoring, Protection and Control Systems: The Enabler for Smarter Grids A. Vaccaro and
A. Zobaa (Editors)
Volume 74 Power Electronic Converters and Systems: Frontiers and Applications A.M. Trzynadlowski (Editor)
Volume 75 Power Distribution Automation B. Das (Editor)
Volume 76 Power System Stability: Modelling, Analysis and Control B. Om P. Malik
Volume 78 Numerical Analysis of Power System Transients and Dynamics A. Ametani (Editor)
Volume 79 Vehicle-to-Grid: Linking Electric Vehicles to the Smart Grid J. Lu and J. Hossain (Editors)
Volume 81 Cyber-Physical–Social Systems and Constructs in Electric Power Engineering Siddharth
Suryanarayanan, Robin Roche and Timothy M. Hansen (Editors)
Volume 82 Periodic Control of Power Electronic Converters F. Blaabjerg, K. Zhou, D. Wang and Y. Yang
Volume 86 Advances in Power System Modelling, Control and Stability Analysis F. Milano (Editor)
Volume 88 Smarter Energy: From Smart Metering to the Smart Grid H. Sun, N. Hatziargyriou, H.V. Poor, L. Carpanini
and M.A. Sánchez Fornié (Editors)
Volume 89 Hydrogen Production, Separation and Purification for Energy A. Basile, F. Dalena, J. Tong, T.N. Veziroğlu
(Editors)
Volume 93 Cogeneration and District Energy Systems: Modelling, Analysis and Optimization M.A. Rosen and S.
Koohi-Fayegh
Volume 95 Communication, Control and Security Challenges for the Smart Grid S.M. Muyeen and S. Rahman
(Editors)
Volume 97 Synchronized Phasor Measurements for Smart Grids M.J.B. Reddy and D.K. Mohanta (Editors)
Volume 101 Methane and Hydrogen for Energy Storage R. Carriveau and David S.-K. Ting
Volume 905 Power System Protection, 4 Volumes
Modeling and Dynamic
Behaviour of Hydropower
Plants
Edited by
Nand Kishor and Jesus Fraile-Ardanuy

The Institution of Engineering and Technology


Published by The Institution of Engineering and Technology, London, United Kingdom
The Institution of Engineering and Technology is registered as a Charity in England &
Wales (no. 211014) and Scotland (no. SC038698).
† The Institution of Engineering and Technology 2017
First published 2017

This publication is copyright under the Berne Convention and the Universal Copyright
Convention. All rights reserved. Apart from any fair dealing for the purposes of research
or private study, or criticism or review, as permitted under the Copyright, Designs and
Patents Act 1988, this publication may be reproduced, stored or transmitted, in any
form or by any means, only with the prior permission in writing of the publishers, or in
the case of reprographic reproduction in accordance with the terms of licences issued
by the Copyright Licensing Agency. Enquiries concerning reproduction outside those
terms should be sent to the publisher at the undermentioned address:

The Institution of Engineering and Technology


Michael Faraday House
Six Hills Way, Stevenage
Herts, SG1 2AY, United Kingdom
www.theiet.org

While the authors and publisher believe that the information and guidance given in this
work are correct, all parties must rely upon their own skill and judgement when making
use of them. Neither the authors nor publisher assumes any liability to anyone for any
loss or damage caused by any error or omission in the work, whether such an error or
omission is the result of negligence or any other cause. Any and all such liability is
disclaimed.
The moral rights of the authors to be identified as authors of this work have been
asserted by them in accordance with the Copyright, Designs and Patents Act 1988.

British Library Cataloguing in Publication Data


A catalogue record for this product is available from the British Library

ISBN 978-1-78561-195-7 (hardback)


ISBN 978-1-78561-196-4 (PDF)

Typeset in India by MPS Limited


Printed in the UK by CPI Group (UK) Ltd, Croydon
Contents

Contributors’ biographies xi

Part I Modeling and simulation of hydropower plants 1


1 Analysis and modeling of run-off-type hydropower plant 3
Roshan Chhetri and Karchung
1.1 Introduction 3
1.2 Measurements 4
1.2.1 Transducers 5
1.2.2 Signal conditioning 5
1.2.3 DAQ hardware 6
1.2.4 LabVIEW 7
1.3 Modeling of the plant 8
1.4 Governor system 11
1.5 Excitation system 11
1.6 Model validation/simulations 13
1.7 Conclusion 18
Bibliography 18

2 Time-domain modeling and a case study on regulation and


operation of hydropower plants 19
Weijia Yang, Jiandong Yang, Wencheng Guo and Per Norrlund
Nomenclature 19
2.1 Introduction 21
2.2 Numerical model of hydropower plants 22
2.2.1 Piping system 22
2.2.2 Hydropower unit with Francis turbine 27
2.2.3 Features of the model 31
2.3 Practical engineering case 31
2.4 Case study of various dynamic processes of hydropower plant 32
2.4.1 Start-up and no-load operation 33
2.4.2 Grid-connected operation 35
2.4.3 Isolated operation 40
2.4.4 Emergency stop and load rejection 43
2.5 Conclusions 46
Acknowledgments 46
References 46
vi Modeling and dynamic behaviour of hydropower plants

3 Reduced order models for grid connected hydropower plants 49


Gérard Robert and Frédéric Michaud
3.1 Introduction 49
3.2 Hydropower plant model 50
3.2.1 Penstock and tunnel models 51
3.2.2 Surge tank model 52
3.2.3 Turbine model in a water column 53
3.2.4 Hydraulic circuit model 56
3.2.5 Mechanical model of the generating unit 59
3.2.6 Hydro-mechanical model of the power plant 61
3.3 Synchronous power system models 61
3.3.1 General model 62
3.3.2 Model for an interconnected grid 64
3.3.3 Model for an isolated grid 66
3.4 Complete state-space model for a hydro plant connected to a grid 67
3.4.1 General model 67
3.4.2 Interconnected operation 68
3.4.3 Isolated operation 69
3.5 Analysis of the dynamic behaviour 69
3.5.1 Decomposition of slow and fast dynamics 70
3.5.2 Performance limitation for primary frequency control:
capability criteria 74
References 77

4 Modeling and stability analysis of turbine governing system


of hydropower plant 79
Wencheng Guo, Jiandong Yang and Weijia Yang
4.1 Introduction 79
4.2 Modeling of turbine governing system 80
4.2.1 Hydraulic submodel 82
4.2.2 Mechanic submodel 84
4.2.3 Electricity submodel 86
4.3 Stability analysis of turbine governing system 86
4.3.1 Basic knowledge of stability of dynamic system 86
4.3.2 Stability analysis of turbine governing system without
surge tank 91
4.3.3 Stability analysis of turbine governing system with
surge tank 94
4.3.4 Critical stable sectional area of surge tank 98
4.4 Conclusions 100
Acknowledgments 100
References 100
Contents vii

Part II Control of hydropower plants 103


5 Dynamic simulation issues for hydropower generation
control 105
Joël Nicolas and Gérard Robert
5.1 Introduction 105
5.2 Grid codes requirements for frequency control and balancing:
example of the European network 106
5.2.1 General overview 106
5.2.2 The European institutional context 106
5.2.3 Brief presentation of the European interconnected
network ENTSO-E 107
5.2.4 The development of European network codes 109
5.2.5 Focus on some European requirements for frequency
control 110
5.3 Application to the design and tuning of turbine governing
systems: the French EDF experience 112
5.3.1 Frequency control and turbine governing systems
specifications 114
5.3.2 Simulation numerical studies: general issues 119
5.3.3 Preliminary simulation numerical studies: principles 119
5.3.4 Preliminary simulation numerical studies: results for
some HPP cases 120
5.3.5 Application for modernised turbine governing systems
with manufacturer’s simulations and performance
field tests 123
5.4 Conclusion 127
References 127

6 Methods of signal analysis for vibration control


at hydropower plants 131
Olga Shindor and Anna Svirina
6.1 Introduction 131
6.2 Hydro units vibration control methodology: implementation
of wavelet transform 133
6.3 Hydropower plant vibration diagnostics case study 136
6.3.1 Controlling object and measurement equipment
characteristics 136
6.3.2 Hydraulic unit’s vibration condition monitoring on the
basis of diagnostics data wavelet analysis 137
6.4 Conclusions 144
References 144
viii Modeling and dynamic behaviour of hydropower plants

Part III Operation, scheduling, etc. of hydropower plants


(including pumped storage) 147
7 Island mode operation in hydropower plant 149
Roshan Chhetri and Karchung
7.1 Introduction 149
7.2 Performance in island mode 150
7.3 Measures to improve the island mode performance 157
7.4 Conclusion 158
Bibliography 158

8 Hydro generation scheduling: non-linear programming


and optimality conditions 161
Lucas S.M. Guedes, Adriano C. Lisboa, Douglas A.G. Vieira,
Pedro M. Maia and Rodney R. Saldanha
8.1 Introduction 161
8.2 Hydropower generation function 164
8.2.1 Physical properties of geometric functions 165
8.2.2 Special cases of geometric functions 167
8.2.3 Mathematical properties 170
8.3 Water conservation and discharge limits 175
8.3.1 Head sensitive discharge limits 176
8.4 Cascade D-HGS formulation 177
8.5 Global optimization approach 178
8.5.1 Computational results 180
8.6 Conclusions 183
References 184

9 A PV hydro hybrid system using residual flow of Guarita


Hydro Power Plant, in southern Brazil 187
Rafael Schultz, Alexandre Beluco, Roberto Petry Homrich
and Ricardo C. Eifler
Abstract 187
Keywords 187
9.1 Introduction 188
9.2 The Guarita hydroelectric power plant 188
9.3 The use of residual flow of Guarita 190
9.4 Components of the PV hydro hybrid system 191
9.5 Simulations with HOMER 193
9.6 Results and discussion 197
9.7 Conclusions 202
Acknowledgments 202
References 202
Contents ix

10 A PV wind hydro hybrid system with pumped storage


capacity installed in Linha Sete, Aparados da Serra,
southern Brazil 205
Alfonso Risso, Fausto A. Canales, Alexandre Beluco and
Elton G. Rossini
Abstract 205
Keywords 205
10.1 Introduction 206
10.2 The Linha Sete pumped storage power plant 207
10.3 Components of the PV wind hydro hybrid system 208
10.4 Simulations with HOMER 211
10.5 Results and discussion 212
10.6 Final remarks 219
Acknowledgments 219
References 219

Part IV Small hydropower plants 223


11 Modeling and simulation of a pico-hydropower off-grid
network 225
Sam J. Williamson, Antonio Griffo, Bernard H. Stark
and Julian D. Booker
11.1 Introduction 225
11.2 System overview 226
11.3 Component models 227
11.3.1 Turbine 228
11.3.2 Shaft assembly 228
11.3.3 Generator 230
11.3.4 Rectifier 231
11.3.5 DC–DC converter 231
11.3.6 Inverter modeling 232
11.3.7 Transmission line and load modeling 232
11.4 Control scheme design 233
11.4.1 Turbine and DC–DC converter controller design 233
11.4.2 Inverter control design 233
11.5 Simulation results 239
11.5.1 Single generator unit with varying load 239
11.5.2 Performance with non-linear load 241
11.5.3 Power sharing performance 242
11.5.4 Change in input power (drop in head) 242
11.6 Modeling of implementation in Nepal 242
11.7 Hybrid renewable off-grid network 245
11.7.1 Solar PV interface modifications 246
x Modeling and dynamic behaviour of hydropower plants

11.7.2 Wind turbine interface modifications 247


11.7.3 Hybrid grid simulation 248
11.8 Summary 249
References 250
Further reading 253

Index 255
Contributors’ biographies

Alexandre Beluco is a civil engineer and holds a PhD in engineering, working as a


professor at the Federal University of Rio Grande do Sul and as a researcher in the
field of renewable energies, specifically working with the feasibility analysis of
hybrid generation systems and hybrid energy storage systems. His research projects
deal with energy complementarity and unexplored energy potential, usually with
micro and small-scale hybrid systems.
Julian D. Booker holds a chair in Mechanical Design Engineering in the Department
of Mechanical Engineering, University of Bristol, UK. He is a fellow of the
IMechE and a chartered engineer. His teaching is associated with design and
manufacture, and his research is associated with the development of machine
design and structural integrity across all major industrial sectors. He has published
over 120 papers and has authored three books on design methods.
Fausto A. Canales is a civil engineer and holds a doctorate in water resources,
working in his native country, Nicaragua, as an engineer and as a researcher spe-
cialized in issues related to water and energy management. His doctoral thesis
explored the generation of hydroelectric power through pumped hydropower plants
and was largely based on the use of HOMER software.
Roshan Chhetri completed Diploma in Electrical Engineering from Royal Bhutan
Polytechnic, BTech from REC (NIT) Warangal, Andhra Pradesh, India, and MScE
in Electrical Engineering from University of New Brunswick, Canada. He has been
teaching engineering students since 1990. Presently he is a senior lecturer at Col-
lege of Science and Technology (CST), Royal University of Bhutan. He has pub-
lished and presented more than 13 papers in national and international journals and
conference.
Ricardo C. Eifler is a civil engineer who acts as a maintenance engineer in the
state-owned company responsible for supplying electricity in the state of Rio
Grande do Sul, the southernmost state in Brazil. He is studying the master’s degree
and in his thesis, he will evaluate the alternatives to increase installed power for the
hydroelectric plant of Ivaı́, considering the installation of photovoltaic panels on
the water surface of the small reservoir.
Antonio Griffo received his MSc degree in Electronic Engineering and the PhD
degree in Electrical Engineering from the University of Naples, Italy, in 2003 and
2007, respectively. From 2007 to 2013, he was a research associate with the
xii Modeling and dynamic behaviour of hydropower plants

University of Sheffield, UK, and the University of Bristol, UK. He is currently a


lecturer with the Department of Electronic and Electrical Engineering, University
of Sheffield. His research interests include modeling, control and condition mon-
itoring of electric power systems, power electronics converters, and electrical
motor drives, for renewable energy, automotive and aerospace applications.
Lucas S.M. Guedes received his bachelor degree in Production Engineering,
master and doctor degrees in Electrical Engineering from Federal University of
Minas Gerais (UFMG), Brazil, in 2010, 2012 and 2016, respectively. Presently, he
is a postdoctoral researcher in the graduate programme in electrical engineering,
UFMG, and a research associate in ENACOM. His recent research includes opti-
mization theory, mathematical analysis and power system planning.
Wencheng Guo received his BS and MS from the School of Water Resources and
Hydropower Engineering, Wuhan University, Wuhan, China, in 2011 and 2013,
respectively. He is currently a joint doctoral candidate at the State Key Laboratory
of Water Resources and Hydropower Engineering Science of Wuhan University
and Department of Agricultural and Biological Engineering of Purdue University.
His research interests are in the safe operation and control of hydropower plants,
transient processes for hydraulic, mechanical, and power coupling systems,
hydraulic turbine regulation, power system stability and control and vibration
damping of fluid power systems. Email: [email protected].
Roberto Petry Homrich is an electrical engineer and holds a doctorate in engi-
neering, working as a professor at the Federal University of Rio Grande do Sul and
as a researcher in the areas of electrical machinery and electromagnetic materials
and devices. He has extensive experience in projects involving superconductivity
and special systems for energy conversion.
Karchung was born in Trashiyangtse, Bhutan, on 5 March 1990. He graduated
from College of Science and Technology with BE in Electrical Engineering, under
Royal University of Bhutan, Bhutan. His special field of interest includes renew-
able energy and hydropower. He was a scholarship-supported member of the
Bhutan’s students exchange group at the University of Rostock, Germany, for
modelling and simulation of Bhutan’s Hydropower Plant, Tala, Bhutan. He is
currently an assistant lecturer at Jigme Namgyel Engineering College under Royal
University of Bhutan.
Adriano C. Lisboa received his bachelor, master and doctor degrees in Electrical
Engineering from Federal University of Minas Gerais (UFMG), Brazil, in 2001,
2003 and 2008, respectively. He is a cofounder and the technology director of
ENACOM, a company specialized in research and innovation in the areas of
optimization, computational modeling and computational intelligence. His recent
research includes game theory, analysis of analytical models and discrete event
modeling.
Pedro M. Maia received the bachelor degree in System Engineering from Federal
University of Minas Gerais (UFMG), Brazil, in 2016. He was an optimization
Contributors’ biographies xiii

research assistant in ENACOM and participated in research and development


(R&D) projects for electric utilities. His recent research includes optimization
theory and the short-term hydropower scheduling problem.
Frédéric Michaud graduated in Control Engineering from the ‘Supelec’ Advanced
School of Electricity in 2009. He then joined EDF, the main French electric utility
company. He is now Engineer at Hydro Engineering Centre of Grenoble, France.
He has been involved in several projects related to hydropower plants coordinated
control algorithms, and managed a project on ancillary services provided by EDF
hydro units.
Joël Nicolas graduated in Electrical Engineering from École Supérieure d’Électricité
(Supélec), France, in 1978. He then joined EDF Group, mainly with the General
Technical Department of EDF Hydro Generation and Engineering Division, for the
testing of turbine governing systems and voltage control systems. He is now Expert
Engineer, especially for ancillary services provided by power plants, including the topic
of the future European grid codes. He is an international expert of IEC-TC4-WG14
‘Governing systems of hydraulic turbines and automation of hydropower plants’.
Per Norrlund received his MS degree in 2000 in Engineering Physics and PhD
degree in 2005 in Numerical Analysis from Department of Information Technol-
ogy, Uppsala University, Uppsala, Sweden. At present, he is a senior research
engineer in Vattenfall AB and a researcher at Division of Electricity, Department of
Engineering Sciences, Uppsala University. His work and research interests include
hydraulic surge, torsional turbine shaft oscillations, frequency control and dis-
charge measurements in hydropower plants.
Alfonso Risso is a civil engineer and has a master’s degree in water resources. He
is a specialist in geoprocessing, with extensive experience in projects involving
environmental issues. His PhD thesis will present contributions to the under-
standing of processes related to energetic complementarity.
Gérard Robert graduated in Electrical Engineering and Automatic Control from
ENSEEIHT in 1994. He joined the R&D division of EDF in 1995 as Research
Engineer to evaluate by modeling and simulation approaches of the dynamic
behaviour of electric power systems. Since 2005, he has been working at EDF
Hydro Engineering Centre where he develops control and monitoring algorithms
for hydropower plants (off line or cascade).
Elton G. Rossini holds a BS in Physics and a doctorate in engineering, working as
a professor at the State University of Rio Grande do Sul and as a researcher in the
field of renewable energies. Acting together with the co-author Alexandre Beluco,
they lead the constitution of a new postgraduate course dedicated to renewable
resources and sustainability.
Rodney R. Saldanha received the bachelor and master degrees in Electrical
Engineering from Federal University of Minas Gerais (UFMG), Brazil, and the
doctor degree in Electrical Engineering from the Institut National Polytechnique de
xiv Modeling and dynamic behaviour of hydropower plants

Grenoble, France, in 1980, 1983 and 1992, respectively. Presently, he is with the
Department of Electrical Engineering, UFMG. His recent research includes opti-
mization theory, reliability theory and power system planning.
Rafael Schultz is an electrical engineer and was at Portland State University, USA,
during the undergraduate course, with support from the Brazilian programme
‘Science without Borders’. His undergraduate work evaluated the feasibility of
increasing power installed in a micro hydroelectric plant, taking advantage of the
ecological flow.
Olga Shindor has graduated from Kazan State Technical University where she
majored in instrument making. She is a PhD candidate; her main fields of interest
are data processing with wavelet transform, systems of technical diagnostic,
renewable energy, prediction of the state of equipment bases on the wavelet
transform of the signal of this equipment. She is the author of more than 30 aca-
demic publications.
Bernard H. Stark is a reader in Electrical and Electronic Engineering at the
University of Bristol, and a member of the Electrical Energy Management
Research Group. His research interests include renewable power sources and power
electronics. He has spent time at ETH Zurich, Cambridge University, Oxford
University, and Imperial College London.
Anna Svirina has graduated from Kazan State Technical University where she
majored in industrial economics and received her PhD from the same university.
She is a doctor of economics sciences who specializes in measuring systems based
on approaches brought to social sciences from natural sciences research. She is the
author of more than 100 academic publications.
Douglas A.G. Vieira received his bachelor and doctor degrees in Electrical Engi-
neering from Federal University of Minas Gerais (UFMG), Brazil, in 2003 and
2006, respectively. He was worked as a research associate at the Imperial College
of London, UK, and as a research assistant at Oxford University, UK. He is a
cofounder and the executive director of ENACOM. His recent research includes
artificial intelligence and optimization theory.
Sam J. Williamson is a research associate in Electro-Mechanical Systems at the
University of Bristol, where he completed his PhD. Since 2009, he has been con-
ducting research into small-scale hydropower for rural electrification in developing
countries, and how microgrids can be used to support sustainable, reliable,
renewable electrification schemes.
Jiandong Yang received the PhD from Wuhan University of Hydraulic and Elec-
trical Engineering, Wuhan, China, in 1988. He is currently a professor at the State
Key Laboratory of Water Resources and Hydropower Engineering Science of
Wuhan University. His research interests are in transient process and control of
hydropower plants and pumped storage power stations, and model testing of tran-
sient processes. Email: [email protected].
Contributors’ biographies xv

Weijia Yang received his BS and MS from the School of Water Resources and
Hydropower Engineering, Wuhan University, Wuhan, China, in 2011 and 2013,
respectively. He is currently pursuing his PhD at the Division of Electricity,
Department of Engineering Sciences, Uppsala University, Uppsala, Sweden. At
present, he is also a visiting PhD student at the State Key Laboratory of Water
Resources and Hydropower Engineering Science of Wuhan University. He is a
young professional member of the International Association for Hydro-Environment
Engineering and Research (IAHR) and a student member of IEEE. His research
interests include dynamic processes and control of hydropower plants, and interaction
between hydropower plants and power systems. Email: [email protected].
Part I
Modeling and simulation of hydropower plants
Chapter 1
Analysis and modeling of run-off-type
hydropower plant
Roshan Chhetri1 and Karchung2

1.1 Introduction
A precise model representing a complete power system and its associated trans-
mission networks forms a basis for any kind of control system analyses. The
reliability aspects with regard to the operation and control of the generating units
and associated power system networks immensely depend on the competencies to
study, understand, and analyze the overall system. A real-time physical test on the
system would not always be a feasible option then, given the risks involved in
terms of operation downtime, unnecessary system disturbances, adverse effect to
the equipment tested, customers’ disconnection and the associated revenue losses.
An only realistic option is to have a precise mathematical/computer model closely
representing the actual power system that would allow for required simulation
studies concerning its overall behavior, in different operating modes such as the
grid connected mode or the islanded operation.
In this chapter, a model of run-off-type hydropower plant is presented which is
developed in MATLAB/Simulink software workspace based on the measure-
ment signals obtained from one of the power plant in Bhutan. The controller and
component parameters are initially taken out from the data sheet. Time constants
and friction constants are calculated from the given parameter or are assumed, and
then later, all pre-assumed parameters were validated by inter- and extrapolation
with measurement signals. The power plant modeling work starts with identifica-
tion of the mathematical governing differential equations of each part which are
then converted to transfer function. The block diagrams are developed using the
functional blocks mostly from the ‘‘commonly used block’’ library instead of
directly using the built in blocks from the Simulink library. Individual blocks are
connected to form whole system which represents the model. The model simulation
result should agree with the measurement signals in all kinds of tests performed

1
Department of Electrical Engineering, College of Science and Technology, Phuentsholing, Bhutan
2
Jigme Namgyel Engineering College, Bhutan
4 Modeling and dynamic behaviour of hydropower plants

(both for slow and for first change in signal); otherwise, we have to inter- or
extrapolate again each of simulation results to fit each with measurement signals. In
that way, we can find the parameters like time constants and friction factors of each
parts of hydropower plant.

1.2 Measurements
Before starting with the real modeling, it is mandatory that we have the measure-
ment data which will act as the basis on which we will confer our model step by
step. In this section, let us discuss how the measurements are taken and what most
important signals are to be measured.
Today, most scientists and engineers use personal computers (PCs) with periph-
eral component interconnect (PCI) personal computer bus, PCI extensions for
instrumentation (PXI)/Compact PCI, Personal Computer Memory Card International
Association (PCMCIA) now called PC Cards, universal serial bus (USB), IEEE1394,
instruction set architecture (ISA), or parallel or serial ports for data acquisition in
laboratory research, test and measurement, and industrial automation. Many applica-
tions use plug-in boards to acquire data and transfer it directly to computer memory.
Others use data acquisition (DAQ) hardware remote from the PC that is coupled via
parallel or serial port. Obtaining proper results from a PC-based DAQ system depends
on each of the following system elements as illustrated in Figure 1.1.
● The PC
● Transducers
● Signal conditioning
● DAQ hardware
● Laboratory Virtual Instrumentation Engineering Workbench (LabVIEW)
software

Conditio
ned
signals

SCXI chassis
PCMCIA DAQCard
or connection to
parallel port ls
na rs SCXI modules
is g nso
e
I/O d s Terminal blocks
an

Figure 1.1 The typical PC-based DAQ system


Analysis and modeling of run-off-type hydropower plant 5

1.2.1 Transducers
Transducers sense physical phenomena and provide electrical signals that the DAQ
system can measure. For example, thermocouples, resistance temperature detector
(RTDs), thermistors, and integrated circuits (IC) sensors convert temperature into
an analog signal that an analog to digital converter (ADC) can measure. In each
case, the electrical signals produced are proportional to the physical parameters
they are monitoring.

1.2.2 Signal conditioning


The electrical signals generated by the transducers must be optimized for the input
range of the DAQ board. Signal conditioning accessories can amplify low-level
signals and then isolate and filter them for more accurate measurements. In addi-
tion, some transducers require voltage or current excitation to generate a voltage
output. Figure 1.2 depicts a typical DAQ system with signal conditioning extension
for instrumentation (SCXI) signal conditioning.
Amplification – signal conditioning modules amplify input signals. The gain is
applied to the low-level signals within the SCXI chassis located close to the
transducers, sending only high-level signals to the PC and minimizing the effects of
noise on the readings.
Isolation – the system being monitored may contain high-voltage transients
that could damage the computer so the isolation is required. Another reason for the
need of isolation is to ensure that the readings from the plug-in DAQ board are not
affected by differences in ground potentials or common-mode voltages. When the

Physical
phenomena

Data acquisition system

Signal
Sensor
conditioning
Acquisition
Computer Software
hardware
Actuator

Data
analysis
Physical
phenomena

Figure 1.2 Features of DAQ system


6 Modeling and dynamic behaviour of hydropower plants

DAQ board input and the signal being acquired are each referenced to ‘‘ground,’’
problems occur if there is a potential difference in the two grounds. This difference
can lead to what is known as a ground loop, which may cause inaccurate repre-
sentation of the acquired signal, or if too large, may damage the measurement
system. Using isolated signal conditioning modules will eliminate the ground loop
and ensure that the signals are accurately acquired.
Multiplexing – a common technique for measuring several signals with a single
measuring device is multiplexing. Signal conditioning devices for analog signals
often provide multiplexing for use with slowly changing signals such as tempera-
ture. This is in addition to any built-in multiplexing on the DAQ board. The ADC
samples one channel, switches to the next channel, samples it, switches to the next
channel, and so on. As the same ADC is sampling many channels instead of one,
the effective sampling rate of each individual channel is inversely proportional to
the number of channels sampled. The SCXI modules for analog signals employ
multiplexing so that as many as 3,072 signals can be measured with one DAQ
board.
Filtering – the purpose of a filter is to remove unwanted signals from the signal
that you are trying to measure. A noise filter is used on direct current (DC)-class
signals such as temperature to attenuate higher frequency signals that can reduce
the accuracy of your measurement.
Excitation – signal conditioning also generates excitation for some transducers.
Strain gauges, thermistors, and RTDs, for example, require external voltage or
current excitation signals. Signal conditioning modules for these transducers
usually provide these signals.
Linearization – another common signal conditioning function is linearization.
Many transducers, such as thermocouples, have a nonlinear response to changes
in the phenomena being measured. The LabVIEW application software includes
linearization routines for thermocouples, strain gauges, and RTDs.
It is important to understand the nature of the signal, the configuration that is
being used to measure the signal and the effects of the surrounding environment.
Based on this information, we can easily determine whether signal conditioning
will be a necessary part of your DAQ systems.

1.2.3 DAQ hardware


The analog input specifications can give you information on both the capabilities
and the accuracy of the DAQ product. Basic specifications, which are available on
most DAQ products, tell you the number of channels, sampling rate, resolution, and
input range. The number of analog channel inputs will be specified for both single-
ended and differential inputs on boards that have both types of inputs. Single-ended
inputs are all referenced to a common ground point.
Sampling rate – this parameter determines how often conversions can take
place. A faster sampling rate acquires more points in a given time and can therefore
often form a better representation of the original signal. For example, audio signals
converted to electrical signals by a microphone commonly have frequency com-
ponents up to 20 kHz. To properly digitize this signal for analysis, the Nyquist
sampling theorem tells us that we must sample at more than twice the rate of the
Analysis and modeling of run-off-type hydropower plant 7

maximum frequency component we want to detect. So, a board with a sampling


rate greater than 40 kS/s is needed to properly acquire this signal.
Multiplexing – a common technique for measuring several signals with a single
ADC is multiplexing.
Resolution – the number of bits that the ADC uses to represent the analog signal
is the resolution. The higher the resolution, the higher the number of divisions the
range is broken into, and therefore, the smaller the detectable voltage changes.
Range – range refers to the minimum and maximum voltage levels that the
ADC can quantize. The multifunction DAQ boards offer selectable ranges so that
the board is configurable to handle a variety of different voltage levels. With this
flexibility, you can match the signal range to that of the ADC to take best advantage
of the resolution available to accurately measure the signal.
The range, resolution, and gain available on a DAQ board determine the smallest
detectable change in voltage. This change in voltage represents 1 LSB of the digital
value and is often called the code width. The ideal code width is found by dividing
the voltage range by the gain times two raised to the order of bits in the resolution.
Voltage
Code width ¼  2n
gain
where n is order of bit.
Analog output circuitry is often required to provide stimuli for a DAQ system.
Several specifications for the digital-to-analog converter (DAC) determine the
quality of the output signal produced – settling time, slew rate, and resolution.
Settling time and slew rate work together determine how fast the DAC can change
the level of the output signal. Settling time is the time required for the output to
settle to the specified accuracy. The settling time is usually specified for a full-scale
change in voltage. The slew rate is the maximum rate of change that the DAC can
produce on the output signal. Therefore, a DAC with a small settling time and a
high slew rate can generate high-frequency signals, because little time is needed to
accurately change the output to a new voltage level.
An example of an application that requires high performance in these parameters is
the generation of audio signals. The DAC requires a high slew rate and small settling
time to generate the high frequencies necessary to cover the audio range. Output reso-
lution is similar to input resolution. It is the number of bits in the digital code that
generates the analog output. A larger number of bits reduce the magnitude of each
output voltage increment, thereby making it possible to generate smoothly changing
signals. Applications requiring a wide dynamic range with small incremental voltage
changes in the analog output signal may need high-resolution voltage outputs.

1.2.4 LabVIEW
LabVIEW is a platform and development environment for a visual programing
language from National InstrumentsTM. The graphical language is named ‘‘G.’’
Originally released for the Apple Macintosh in 1986, LabVIEW is commonly used
for data acquisition, instrument control, and industrial automation on a variety of
platforms including Microsoft Windows, various flavors of UNIX, Linux, and Mac
OS X. The code files have the extension ‘‘.vi,’’ which is an abbreviation for
8 Modeling and dynamic behaviour of hydropower plants

‘‘Virtual Instrument.’’ LabVIEW plots and displays signal on the computer screen.
As soon as its execution is stopped, it automatically saves the data in per unit
system which can be called from MATLAB software for analysis.
Based on the numbers of channels available on the DAQ Hardware, that many
signals can be recorded in parallel from the field. For identification of the dynamic
behavior of the unit, the set points of active power (affecting turbine regulator) and
generator voltage (affecting voltage regulator) are changed manually in different
experiments by giving input commands from command board for increasing/
decreasing power or voltage. For instance, the load throw-off test (emergency shut-
down) and mechanical shutdown test, no load (dry) test, step-wise increase or
decrease of power test, island mode testing, etc. are most common measurement
signals recorded from the field.

1.3 Modeling of the plant


The overall structures of any power plants are same except for the type of mechanical
system and turbine. Controller structures and the generators are almost same. In
Figure 1.3, the encircled ones are the most important signals that have to be measured
so as to check whether the simulation result follows all the measured signals.
The mathematical modeling is shown in Table 1.1. The mathematical
governing equations (differential equations) of each hydraulic or electric components

yTset yT yT

Turbine regulator β β
Power generation

fN
system (PGS)

hN

hN
Hydraulic system
PT turbine
hB
ie
fN qG
UG Generator PG
wG
Voltage regulation

Ue
system (VRS)

Excitation system
qG

PG

UP UP
Voltage regulator
UGset

Figure 1.3 General representation of sub-model HPP


Analysis and modeling of run-off-type hydropower plant 9

Table 1.1 Mathematical differential equations of each hydraulic component

Section Differential equation Laplace transformed equations


Head race dqt Twc qwc  s þ hs þ Kf qs ¼ 0
tunnel Twc þ h  hs þ Kpr q ¼ 0
dt
Surge shaft dht Ts hs  s þ qs  q ¼ 0
inertia Twsc  qs ¼ 0
dt
Penstock dq Twps qwps  s þ h  hs þ Kfps qs ¼ 0
Twps t þ h  hs þ Kpr q ¼ 0
dt
Generator dn Dn DP 1
Ta  En n ¼ P m¼ ¼
dt n P0 Ta s þ En
Governor dg dn KD d2n 1 þ Ts  s
¼ Kp þ ðnref  nÞ  Kt TN y ¼ Kp þ Kst þ KD s ¼ m
dt dt Td dt bt Td  s

are transformed to Laplace equations, and as it is easy to represent Laplace equations


in block representations, we can join the blocks to get actual model.
Let us derive the differential equation of penstock as an example and the others
can be self-explored. In penstock model, an incompressible fluid and a rigid conduit
for hydroelectric power plants with short or medium penstock are to be assumed
where the traveling pressure wave effects are relatively insignificant. Then, we have
the following equation:
hðsÞ
¼ Tw s  Hf (1.1)
qðsÞ
Neglecting the hydraulic friction losses, (1.1) reduces to
hðsÞ
¼ Tw s (1.2)
qðsÞ
This is the Laplace transfer function when no frictional effect is taken into con-
sideration. When the frictional losses are taken into consideration, (1.2) can be
modified as shown in (1.3).
If hws and hedr are head available at inlet of penstock and nozzle, respectively,
then:
ð
1
q¼ ðhws  hedrÞds
Twps
where hws and hedr are available head at the inlet and outlet of penstock and
nozzle, respectively, also
hedr ¼ hws þ Dhws
where

Dhws ¼ kfps  qadr 2 (1.3)


10 Modeling and dynamic behaviour of hydropower plants

Also, water starting time in penstock will be:


Lps QN
Twps ¼ X (1.4)
gAps HN

Δhws 1 q
sTw

The starting time in penstock is defined as the time required for head HN to accel-
erate the water in penstock from standstill to velocity, UN ¼ QN =Aps . It should be
noted that Twps varies with load. Typically, Twps full load lies between 0.5 and 4.0 s.
Note that the models based on use of a water column time constant Twps as
described above may not adequately represent all of the pertinent dynamics of
plants with very long penstocks. The penstock model dynamics using Twps is valid
only if the traveling wave time is much shorter than the water starting time.
For very long penstocks, the wave travel time of the water column becomes
significant, and the reflected pressure waves in the water column cause the pre-
ceding treatment of water start time to no longer be valid. When the traveling wave
time approaches 25% of the Twps , engineers should not rely on only the classic value
of Twps , and the performance of the turbine governing system should be evaluated by
considering the effects of both the water starting time and the wave travel time.
The friction factor of the penstock can be determined by interpolating the sta-
tionary points of the measurement signals as shown in Figure 1.4. The assumption of
penstock friction loss is made initially at 5% and later corrected using identification
function Dh ¼ rf  ðqP1 Þ2 . Time constants are also corrected from measurements.
Static characteristic relationships of hydraulic turbines can be studied through
the so-called hill charts. The plots of the prototype turbine characteristics are based
in steady-state model test results. These turbine characteristics are assumed valid
during the transient state. The turbine efficiency for any operating point given by
runner speed, net head, and gate position can be extracted from the hill charts.
The hill chart gives relationship between power, efficiency, discharge, and
head. We can have a table of values derived from the hill chart by using the rela-
tionship P ¼ hrQH. From the table of values obtained, we digitize it to a 3-D
lookup table to be used in the model as shown in Figure 1.5.
An operating point of a hydraulic turbine is characterized by the specific
energy, the discharge, the rotational speed, the torque, and the gate opening posi-
tion. Therefore, the graphical representation of a turbine characteristic requires the
elimination of one of these quantities by the use of the hydraulic machines simili-
tude laws.
Individual block diagrams are connected together and the resulting overall model
is shown in Figure 1.6. In this model, we are considering a reservoir head hB which is
input to the head race tunnel. The frictional losses and inertia of water is quite
significant so it has to be taken into account. We are assuming two penstocks each
feeding equal numbers of units. Here, only one penstock model is shown but the
effect due to other penstock is taken into account for better result. The controller
signals are converted to nozzle servo movement following the conjugation function.
Analysis and modeling of run-off-type hydropower plant 11

0.05
Stationary values
Identified penstock loss function (P1)
0.04

0.03
Loss of head dhedr [p.u.]

0.02

0.01

–0.01

–0.02
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Flow qP1 [p.u.]

Figure 1.4 Iterative identification of friction factor for penstock

1.4 Governor system


The governing system model can be derived thoroughly understanding both elec-
trical controller and mechanical governor structure. The electrical signal is received
by the electro-hydraulic (EH) converter (EH transducer), and the mechanical signal
is then amplified by the pilot servo. Depending upon the direction of flow of
hydraulics from the pilot valve, the main servo is actuated in either direction (open
or close). The hydraulic pressure is high and the movement of the spear in gate
servomotor is made significant.
Hydro turbine governors are designed to have relatively large transient droop
compensation with long resetting times. This ensures stable frequency regulation
under isolated operating conditions. The response of a hydro turbine to speed
changes or to changes in speed-changer setting is relatively slow.
The simplified model of the governing system is shown in Figure 1.7. Any type
of governor will have use of two sets of servos: deflector and nozzle. Deflector
movement is controlled by the direct mechanical feedback and obtains signal from
speed signal generator (SSG), whereas the nozzle servo movement depends upon
the deflector movement, and the movement is in accordance to the correlation
function which is called ‘‘nozzle-Deflector Conjugation Function.’’

1.5 Excitation system

There are different types of excitation systems like alternating current (AC), DC,
and static excitation systems (SES), but nowadays, SES are preferred due to various
Turbine power characteristic Turbine efficiency characteristic

94
250
92
200
150 90
Surface
100 88 net points
from hill chart
50

Efficiency [%]

Power [MW]
86
0 Surface
net points
from hill chart 84
–50 900
900
30 850 30
850
20 20
800 800
10 10
750 0 750 0
Pressure head [m] Discharge [m³/s] Pressure head [m] Discharge [m³/s]

Figure 1.5 3-D hill chart derived from Figure 1.5


Analysis and modeling of run-off-type hydropower plant 13

qaPS2 qU1 qU3 yT1...yT5 yD

Water channel Nozzles


opening
ΔhWC 1
kfWC q × |q| function
2
Inertia
hB – 1 qWC – aT
– sTW WC
Deflector aD
Surge chamber effect

Volume
hSC 1 qSC
sTSC –
1
Penstock 1 2

kfPS q × | q| qePS1 qU2 qT2
ΔhPS1 1
3
Inertia
1 qaPS1 – 1 hN
– sTW PS1 sTD h
Compressibility hT2

Tala

Figure 1.6 Hydraulic model of hydropower plant with two penstocks

reasons like static type do not use rotating exciters and thus have a much faster
dynamic response and a larger field forcing capability to respond to large dis-
turbances without exceeding generator field current limits. However, due to the
high initial response, they require voltage regulators with high gains that may have
an adverse impact on the damping of electromechanical oscillatory modes in power
systems. Power system stabilizers are often used as supplementary controls to add
positive damping to the affected oscillatory modes through the excitation system by
adding an electric torque in phase with the generator rotor speed. Typical
arrangement of elements in excitation system is illustrated in Figure 1.8 and SES
simulation model in Figure 1.9.

1.6 Model validation/simulations

The last step is the unification of the sub-models to a complete model, which enables
the simulation of the system in closed loop. Measurement delays had to be imple-
mented owing to real time scenarios, where the signals and feedbacks are to be
inevitably transformed into the controllers’ signal format, where small delays are
unavoidable. The sub-models according to Figure 1.3 is developed and simulated.
For the proof of accuracy of the model, comparison of measurement and
simulation in interconnected operation mode and load shedding operation is carried
out and shown in Figure 1.10. Simulation results in Figure 1.10 shows the results of
14 Modeling and dynamic behaviour of hydropower plants

yPD max Deflector servos 1

yDctr 1 1 yD
sTDP sTDM

yPD min Valve opening 0
characteristic
Pilot servo Main servo

PD T1
yDlim
klim MIN
– –
yD → yT yD → yT
Table

Stop
MIN kBA Digital nozzle
controller

yPT max 1
yTctr Nozzle servos
1 1 yT1...5
sTTP sTTM

yPT min Valve opening
0
characteristic
Pilot servo Main servo
Tala

Figure 1.7 Sub-model of oil-hydraulic part of governor as well as digital nozzle


controller, exemplary for one nozzle

Limiters and
protective circuits

3
Terminal voltage
transducer and
load compensator

2 1
To power
Ref Regulator Exciter Generator system

Power system
stabilizer

Figure 1.8. Elements of the excitation system (source: power system analysis by
P.B. Kundur)
Vt0
1
Vref VA0/Ka
VA0/Ka
VA0=Efd0+KLR*(Ifd-ILR)
VA0/Ka VAmax
+
VA0/Ka Vlmax VA0/Ka VA0/Ka VA0 VtVRmax-KcIfd
Vt0 + Efd0
1 Vt0 48s2+14s+1 1
2 – + Efd Efd0
Vt 0.02s+1 Vt Vl 35s2+12s+1 2s+1 VA – Efd
Vt + 1
Low pass filter – Vlmin Transient gain
Vt
Efd
1/(Tr.s+1) Limits Vlmin, Vlmax reduction VAmin Ifd Efd
(Tc.s+1).(Tc1.s+1) Main regulator VtVRmin
4 Ka/(Ta.s+1)
(Tb.s+1).(Tb1.s+1)
Vstab

0 Efd0
3s
Vf 4s+1
Damping filter
Kfs/(Tf.s+1)
KLR*(Ifd-ILR)
KLR –+ 3
Initial values are shown in blue
Ifd
KLR
ILR

ILR

Figure 1.9 ST1A static excitation system model (source: DigSILENT PowerFactory library)
16 Modeling and dynamic behaviour of hydropower plants
200 1
Measurement Measurement
Setpoint Simulation
150 Simulation

Position [p.u.]
pG [MW]

0.8
100
Change to 5-nozzle
operation 0.6
50
Change to 2-nozzle
operation
0 0.4
0 500 1,000 1,500 2,000 0 500 1,000 1,500 2,000
Time [s] Time [s]
(a) Active power (b) Deflector position

0.8
yT1 Measurement 0.06
Measurement
yT2 Measurement
0.6 0.04 Simulation
Position [p.u.]

yT1 Simulation

Signal [p.u.]
yT2 Simulation
0.4 0.02

0.2 0
–0.02
0
–0.04
0 500 1,000 1,500 2,000 0 500 1,000 1,500 2,000
Time [s] Time [s]
(c) Nozzle position 1 (d) Deflector control signal

Figure 1.10 Comparison of measurement and simulation for change of power


set points

0.2
defl. Measurement Turbine flow simulation
0.6 yT1 Measurement 0.15 Nozzles flow simulation
Position [p.u.]

defl. Simulation
Flow [p.u.]

0.4 yT1 Simulation 0.1

0.2 0.05

0
0
–0.05
0 50 100 150 200 30 40 50 60 70
Time [s] Time [s]
(a) Defector position, nozzle 1 (b) Turbine flow/nozzle

0.1 1.04
Measurement
Turbine speed [p.u.]

0 Simulation
1.02
Signal [p.u.]

–0.1
1
–0.2
0.98
–0.3 Measurement
Simulation
–0.4 0.96
0 50 100 150 200 0 50 100 150 200
Time [s] Time [s]
(c) Deflector control signal (d) Turbine speed

Figure 1.11 Comparison of measurement and simulation for load shedding


of 30 MW

a change of power set-points within the whole operational range. Even the transi-
tions from five to two nozzle operations and back are included. A sequence of three
tests in a time range of more than half an hour is done. This wide range is needed to
get a proof of the surge shaft oscillations with very high period lengths, which
occur after a fast shut down of power plant.
The state of the plant is detected by the control structure and the model acts
like the real plant under the boundary conditions of test ambience.
Analysis and modeling of run-off-type hydropower plant 17

Surge chamber oscillations after shut down


90

Pressure head [kp/cm²]


Simulation
Measurement 8
88 Measurement 9
Measurement 10

86

84

0 1,000 2,000 3,000 4,000


Time [s]
90
Pressure head [kp/cm²]

88

86

84

250 260 270 280 290 300


Time [s]

Figure 1.12 Variation in surge tank oscillations following the shut down

1.1
Measurement
Turbine speed [p.u.]

1.05 Simulation

0.95

0.9

0.85
100 200 300 400 500
Time [s]

Figure 1.13 Speed simulation of a load shedding test of full load (170 MW)

In Figure 1.11, a load of about 30 MW is shed by opening the generator main


circuit-breaker/gas circuit breaker (GCB). Also, the simulated water flow through
the nozzle assembly and the diverting effect of the deflector are shown in
Figure 1.11(b). For both simulations, it can clearly be seen, that the model is able to
fit the real behavior with a high accuracy. Slow as well as fast changes and different
controller paths can be modeled with a comparable high quality.
Figure 1.12 shows the variation is pressure in surge tank oscillation after
the shutdown. In the zoomed-in picture below, the fast slightly damped penstock
18 Modeling and dynamic behaviour of hydropower plants

traveling waves can be recognized. The different oscillations of very different time
ranges are reproduced very close to the measurement signal. From this comparison,
it can be derived that the fundamental wave model of water hydraulics is sufficient
for these purposes.
The speed variation in Figure 1.13 suggests a good fit of ramp rates for
speeding up and slowing down of the rotating unit after a load shedding of full load
(170 MW). The factors responsible are the inertia of the rotating system with an
identified time constant of 10 s and the no load losses (mechanical losses) of about
2.3 MW at nominal speed. The reaction of the system under cooperation of the
deflector are thus accurately be simulated.

1.7 Conclusion
A model run-of-type hydropower plant is developed in MATLAB/Simulink soft-
ware workspace based on the measurement signals obtained from power plant.
Time constants and friction constants are calculated from the given parameter or
are assumed and then later all pre-assumed parameters were validated by inter-
and extrapolation with measurement signals. The power plant model starts with
identification of the mathematical governing differential equations of each part
which are then converted to transfer function. The block diagrams were developed
using the functional blocks in the Simulink library. Individual blocks are connected
to form whole system which represents the model.

Bibliography
[1] IEEE Committee. 1973. Dynamic models for steam and hydro turbines in
power system studies. IEEE Transactions on Power Apparatus and Systems;
92:1904–1915.
[2] Qijuan C. and Zhihuai X. 2000. Dynamic modeling of hydroturbine gen-
erating set. In: IEEE International Conference on Systems, Man and
Cybernetics, IEEE, 8–11 Oct. 2000, pp. 3427–3430.
[3] Acakpovi A., Hagan E. B., and Fifatin F. X. 2014. Review of hydropower
plant models. International Journal of Computer Applications (0975–8887);
108(18), 33–38.
[4] Bosona T. G. and Gebresenbet G. 2010. Modeling Hydropower Plant System
to Improve Its Reservoir Operation. Department of Energy and Technology,
Swedish University of Agricultural Sciences, Box 7032, 750 07 Uppsala,
Sweden.
[5] Kozdras K. 2015. Modeling and Analysis of a Small Hydropower Plant and
Battery Energy Storage System Connected as a Microgrid. University of
Washington, Seattle, WA, USA.
[6] Machowski J., Bialek J., and Bumby J. 2008. Power System Dynamics, 2nd
ed. West Sussex: Wiley.
[7] Yang W., Yang J., Guo W., et al. 2015. A mathematical model and its
application for hydro power units under different operating conditions.
Energies; 8:10260–10275; doi:10.3390/en80910260.
Chapter 2
Time-domain modeling and a case study on
regulation and operation of hydropower plants
Weijia Yang1,2, Jiandong Yang2, Wencheng Guo2,3
and Per Norrlund1,4

Nomenclature
a velocity of pressure wave
A cross-sectional area of pipeline
Ao cross-sectional area of the orifice in surge tank
Ast cross-sectional area of the surge tank
AP cross-sectional area at the point P (the same for other points, e.g., S, L, R,
P1, P2, P3, etc.)
ATh critical cross-sectional area of the surge tank based on Thoma criterion
D inner diameter of the pipe
D1 diameter of runner
eg coefficient of load damping
F Darcy–Weisbach coefficient of friction resistance
fc given frequency
fg generator frequency
G gravitational constant
H piezometric water head in the pipeline
H0 gross water head (in the formula of Thoma criterion)
H1 net water head (in the formula of Thoma criterion)
HP piezometric water head of the point P (the same for other points, e.g., S, L,
R, P1, P2, P3, TP)

1
Department of Engineering Sciences, Uppsala University, Uppsala SE-751 21, Sweden
2
The State Key Laboratory of Water Resources and Hydropower Engineering Science, Wuhan Uni-
versity, Wuhan 430072, China
3
Maha Fluid Power Research Center, Department of Agricultural and Biological Engineering, Purdue
University, West Lafayette, IN 47907, USA
4
Vattenfall R&D, Älvkarleby SE-814 26, Sweden
20 Modeling and dynamic behaviour of hydropower plants

Hu piezometric water head of the upstream reservoir


Hd piezometric water head of the downstream reservoir
hw0 head loss in the draw water tunnel (in the formula of Thoma criterion)
hwm head loss in the penstock (in the formula of Thoma criterion)
J moment of inertia
L length of the pipeline (in the formula of Thoma criterion)
M11 unit moment
Mg resistance moment of generator
Mt mechanical moment of turbine
n11 unit rotation speed
N rotation speed
nc given rotation speed
nr rated rotation speed
pc given power
pr rated power output
pg generator power
Q discharge
QP discharge of the point P (the same for other points, e.g., S, L, R, P1, P2,
P3, TP)
Q11 unit discharge
Sp wetted perimeter at the point P
Ss wetted perimeter at the point S
T time
Tr response time
V flow velocity in the pipeline
X position
xf relative value of speed (frequency) deviation, xf ¼ ( fg  fc)/fc
Y guide vane opening(servomotor stroke)
yc given opening
yPID opening deviation after PID terms
yservo opening deviation after servo block
Z Piezometric water head in the surge tank
Zd water level of the downstream reservoir
a coefficient of head loss (in the formula of Thoma criterion): a ¼ hw0/v2
aP correlation coefficient of kinetic energy at the point P
aS correlation coefficient of kinetic energy at the point S
q angle between axis of pipeline and horizontal plane
zsj coefficient of head loss for the series junction
Time-domain modeling and a case study 21

zu coefficient of head loss for the upstream reservoir


zd coefficient of head loss for the downstream reservoir
z1–2 coefficient of head loss for the branch junction
z1–3 coefficient of head loss for the branch junction
zst coefficient of head loss for the surge tank
Dt time step in simulation
Dn speed deviation
   
DH DH ¼ aP =2gAP 2  aS =2gAS 2 QP 2
GV guide vane
GVO guide vane opening
HPP hydropower plant
PI proportional–integral
PID proportional–integral–derivative
PFC primary frequency control
OF opening feedback
PF power feedback

Other symbols in governor equations are illustrated in Figure 2.8.

2.1 Introduction
Hydropower units undertake the frequency control, peak load modulation, and
emergency reserve in electric power systems because of the great rapidity and
amplitude of their power regulation. A hydropower system is a complex nonlinear
system that contains hydraulic, mechanical, electrical, and magnetic subsystems.
For the sake of ensuring safe, stable, and efficient operation of hydropower plants
(HPPs), the numerical modeling of HPPs and the research on dynamic processes in
the regulation and operation of HPPs is of great importance.
Much research has been concentrated on the modeling and dynamic processes
of HPPs. A hydraulic system modeling method is proposed, and the interactions
between power system oscillation and the dynamic characteristics of the hydraulic–
mechanical system are further discussed in [1,2]. Nonlinear models for the transient
processes of the HPPs, with a focus on the influence of the surge tank, are
constructed in [3–5]. An integrated system analysis model, with respect to the
rotational speed and active power control during HPP operation, is proposed [6].
A high-order model of HPPs in islanded power networks is built and unsteady
operation of hydroelectric systems is studied in [7,8]. A refined model for pumped
storage power plants is established, and the nonlinear, multivariable and time-
variant system characteristics are investigated based on a real case in Great Britain
[9]. An operating model for grid-connected pumped storage power plants, for
studying hydraulic short-circuit characteristics, is presented in [10]. In a
22 Modeling and dynamic behaviour of hydropower plants

comprehensive review [11], research results in modeling, control strategies, as well


as regulation and operation performance for HPPs are introduced extensively. In
two books [12,13], the modeling and various operational control strategies of HPPs
are presented in detail.
In this chapter, the main scope lies in the time-domain analysis of regulation
and operation of HPPs. The primary goal is to briefly present the methods and
procedures of the analysis. First, a numerical model in the software TOPSYS [14],
developed for scientific studies and consultant analyses of the transient processes of
HPPs, is briefly presented. Second, information of a practical engineering case, i.e.,
areal Chinese HPP with a surge tank and Francis turbines, is introduced. In the third
place, a case study of various dynamic processes is conduced based on a real
consultant project of the Chinese HPP; key requirements and main influencing
factors of the diverse conditions of regulation and operation are analyzed.

2.2 Numerical model of hydropower plants

In this section, a numerical model of hydropower plants implemented in the soft-


ware TOPSYS [14] is introduced, including the model of piping system and gen-
erating units.

2.2.1 Piping system


In this subsection, modeling of the piping systems of HPPs and the solving method
are introduced. Various boundary conditions, i.e., upstream reservoir, downstream
reservoir, series junction, branch junction, and surge tank, are included.
Considering the elasticity of water and pipe wall, equations for one-dimensional
compressible flow in draw water tunnel and penstock are described by the continuity
equation and the momentum equation, as shown in (2.1) and (2.2), respectively:

Q @H @H a2 @Q a2 Q @A Q
Continuity equation: þ þ þ  sin q  ¼ 0 (2.1)
A @x @t gA @t gA @x A

@H @Q @Q fQjQj
Momentum equation: gA2 þQ þA þ ¼0 (2.2)
@x @x @t 2D

The details of all the symbols in this chapter are given in the Nomenclature.
This set of hyperbolic partial differential equations can be solved by a standard
and widely used approach, the method of characteristics [15]. The common char-
acteristic line and the characteristic grid are demonstrated in Figure 2.1.
For the computation point P shown in Figure 2.1, the equation set may be
transferred to a simple form, as shown in the following equation:

C þ : QP ¼ QCP  CQP  HP (2.3)



C : QP ¼ QCM þ CQM  HP : (2.4)
Time-domain modeling and a case study 23

P
(n + 1) . Δt
C+ C–

n . Δt
A L R B

(i – 1) . Δx i . Δx (i + 1) . Δx x

Figure 2.1 Characteristic lines and the characteristic grid in x–t plane, with
interpolation points L and R

where the symbols are explained as follows:

1
CQP ¼ ;
ðC  C3 Þ=AP þ C ðC1 þ C2 Þ
1
CQM ¼ ;
ðC þ C3 Þ=AP þ C ðC4 þ C5 Þ
   
C þ C3
QCP ¼ CQP QL  C  C1 þ HL ;
AL
   
C  C3 a
QCM ¼ CQM QR  C  C4  HR ; C¼ ;
AR g
aðAP  AL Þ DtSP jQL j 1
C1 ¼ ; C2 ¼ f; C3 ¼ Dt sin q;
2AP ðaAL þ QL Þ 8AL A2P 2
aðAP  AR Þ DtSP jQR j
C4 ¼ ; C5 ¼ f: (2.4a)
2AP ðaAR  QR Þ 8AR A2P

The details for the interpolation and the implementation for solving the equations
can be found in [15]. For simulating the whole piping system, suitable boundary
conditions are crucial, and they are described in the following five subsections.

2.2.1.1 Upstream reservoir with constant water level


For an upstream reservoir with constant water level (head), as shown in Figure 2.2,
there are two unknown variables, i.e., HP and QP. The boundary condition can be
represented by the following equation:

QP 2 QP jQP j
HP ¼ Hu   zu : (2.5)
2gAP 2 2gAP 2
24 Modeling and dynamic behaviour of hydropower plants

Upstream reservoir

Hu
C–

HP

Figure 2.2 Model of upstream reservoir

Here QP greater than zero indicates that the flow is in the forward direction as
demonstrated in Figure 2.2. If QP is smaller than zero, the flow will be in the
reversed direction. Hence, the boundary condition becomes:

QP 2
HP ¼ Hu  ð1  zu Þ : (2.6)
2gAP 2

By combining (2.5) with the above equation:


C  : QP ¼ QCM þ CQM  HP (2.7)

the two unknowns can be solved analytically.

2.2.1.2 Downstream reservoir with constant water level


For a downstream reservoir with constant water level (head), as shown in
Figure 2.3, there are two unknown variables, i.e., HP and QP. The boundary con-
dition can be represented by the following equation:

QP 2 QP jQP j
HP ¼ Hd  þ zd (2.8)
2gAP 2 2gAP 2

As QP is greater than zero, it indicates that the flow is in the forward direction that
is shown in Figure 2.3. If QP is smaller than zero, the flow will be in reversed
direction; therefore, the boundary condition becomes:

QP 2
HP ¼ Hd  ð1 þ zd Þ : (2.9)
2gAP 2

By combining (2.8) or (2.9) with (2.10), values of the two unknowns can be
obtained analytically:

C þ : QP ¼ QCP  CQP  HP ; (2.10)


Time-domain modeling and a case study 25

Downstream reservoir

Hd
C+

HP

Figure 2.3 Model of downstream reservoir

C+ C–

1 2

Figure 2.4 Model of series junction

2.2.1.3 Series junction


For a series junction, as shown in Figure 2.4, there are four unknown variables, i.e.,
HP1, QP1, HP2, and QP2. The number in the subscript corresponds to the number in
the figure. The boundary condition can be represented by the following equations:
QP1 ¼ QP2 (2.11)
QP1 2 QP2 2 QP1 jQP1 j
HP1 þ 2
¼ HP2 þ þ zsj : (2.12)
2gA1 2gA2 2 2gA1 2
The four unknowns can be solved through the above continuity equation and
momentum equation with the following characteristic equations:

C þ : QP1 ¼ QCP1  CQP1  HP1 (2.13)



C : QP2 ¼ QCM2 þ CQM2  HP2 : (2.14)

2.2.1.4 Branch junction (pipeline junction)


The branch junction is also called ‘‘pipeline junction’’ [15]. Taking a type of branch
junction, as shown in Figure 2.5, as the example, there are six unknown variables,
26 Modeling and dynamic behaviour of hydropower plants

C2–

C+
2
1
3 C3–

Figure 2.5 Model of branch junction

i.e., HP1, QP1, HP2, QP2, HP3, and QP3. The number in the subscript corresponds to
the number in the figure. The boundary condition can be presented in the following
equations:
QP1 ¼ QP2 þ QP3 (2.15)
QP1 2
QP2 2
QP1 jQP1 j
HP1 þ 2
¼ HP2 þ 2
þ z12 (2.16)
2gAP1 2gAP2 2gAP1 2
QP1 2 QP3 2 QP1 jQP1 j
HP1 þ 2
¼ HP3 þ þ z13 : (2.17)
2gAP1 2gAP3 2 2gAP1 2

By combing three characteristic equations:

C þ : QP1 ¼ QCP1  CQP1  HP1 (2.18)


C2 : QP2 ¼ QCM2 þ CQM2  HP2 (2.19)
C3 : QP3 ¼ QCM3 þ CQM3  HP3 ; (2.20)

values of the six unknown variables can be obtained.

2.2.1.5 Surge tank


A surge tank comes in a variety of different forms. In this section, only the widely
used orifice surge tank [15] is taken as an example to present the modeling method.
As shown in Figure 2.6, the model of the orifice surge tank contains seven
unknowns, namely, HP1, QP1, HP2, QP2, HTP, QTP, and Z.
The seven unknowns can be solved through the following seven equations for
the surge tank model:

QTP jQTP j
Momentum equation: Z ¼ HTP þ Zd  zst : (2.21)
2gAo 2
HTP ¼ HP1 (2.22)
Time-domain modeling and a case study 27

QTP

HTP
Qp1,Hp1 Qp2,Hp2

Figure 2.6 Model of orifice surge tank

HTP ¼ HP2 (2.23)


Continuity equation: QP1 ¼ QP2 þ QTP (2.24)
QTP þ QTPDt Dt
Equation of water level: Z ¼ ZDt þ (2.25)
2 Ast

C þ : QP1 ¼ QCP1  CQP1  HP1 (2.26)

C  : QP2 ¼ QCM2 þ CQM2  HP2 (2.27)

Here the water level of the downstream reservoir, Zd, is selected as the datum
elevation for the piezometric water head. The inertia of the water inside the surge
tank and the frictional head loss are ignored. The subscript ‘‘Dt’’ represents that
the value is for the last time step. Equations (2.21) and (2.22) are the momentum
equation and continuity equation of the water flow of the surge tank.

2.2.2 Hydropower unit with Francis turbine


For simulation of the dynamic processes of the hydraulic–mechanical–electrical
coupling system in HPPs, the key is the appropriate model of the generating unit. In
[14], the model of the hydropower unit, including the governor system model for
different operating conditions, is presented in detail. Therefore, only the main
equations are shown in this subsection, and the explanation and discussion can be
found in [14].

2.2.2.1 Turbine and generator


In this chapter, only Francis turbine, the common form of reaction turbine, is dis-
cussed. The turbine model is illustrated in Figure 2.7, and the equations of the unit
are shown in Table 2.1.
28 Modeling and dynamic behaviour of hydropower plants

C+

P C–

Figure 2.7 Model of Francis turbine

Table 2.1 Equations of the model of hydropower units [14]

Francis Continuity equation: QS ¼ QP (2.28)


turbine Characteristic þ
C : QP ¼ QCP  CQP  HP (2.29)
equations: C  : QS ¼ QCM þ CQM  HS (2.30)
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Flowing equation of QP ¼ Q11 D21 ðHP  HS Þ þ DH (2.31)
turbine:
Equations of unitary nD1
n11 ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (2.32)
parameters: ðHP  HS Þ þ DH
Mt ¼ M11 D31 ðHP  HS þ DH Þ (2.33)
Characteristic curve Q11 ¼ f1 ðn11 ; yÞ (2.34)
equations of turbine: M11 ¼ f2 ðn11 ; yÞ (2.35)
Generator Isolated operation p dn 30eg pr
J ¼ Mt  Mg  2 Dn (2.36)
(single-machine) 30 dt nr p
 
Single-machine to n ¼ nc fg ¼ fc (2.37)
infinite bus
Off-grid operation p dn
J ¼ Mt (2.38)
30 dt

Here functions f1 and f2 indicate the interpolation of the characteristic curves of


the turbine. The transform from torque to power output is given as
Mt  pn
pg ¼ : (2.39)
30
2.2.2.2 Governor system
The governor system model contains three main control modes, i.e., frequency con-
trol, opening control, and power control. These three modes are based on different
inputs to the control system, i.e., frequency deviation, opening deviation, and power
deviation, respectively. The block diagram of the governor system is shown in
Figure 2.8, and the equations for different control modes are presented in Table 2.2.
Time-domain modeling and a case study 29

Dead Kp
Servo Rate
zone S1 (lag) Saturation limiting Backlash
Given + –xf + + +
fc Edz 1 Kd s +
yPID 1 yservo
frequency –
O2 + + + Ty s + 1
0
S2 Ki y
1 2 3 s
Turbine
(water way
Droop bp 0 ep Droop system)
yc Given opening + –
(set point) Generator power pg
S3
1 Feed-forward
0 2
3
Given power + – Generator
pc
(set point)
Generator frequency fg

Figure 2.8 Block diagram of the governor system [14]

Table 2.2 Equations for different control modes [14]

Frequency control dyPID 2   dyPID


under opening b p Kd 2
þ 1 þ bp Kp þ bp Ki ð yPID  yc Þ
d t dt
feedback:   (2.40)
dxf 2 dxf
¼  Kd 2 þ Kp þ Ki xf
d t dt

Frequency control dpg 2 dpg   dyPID


under power e p Kd þ ep Kp þ ep Ki pg  pc þ
d2t dt dt
feedback:   (2.41)
dxf 2 dxf
¼  Kd 2 þ Kp þ Ki xf
d t dt

Opening control: d ðyPID  yc Þ2   d ð yPID  yc Þ


b p Kd þ 1 þ bp Kp
d t2 dt (2.42)
þ bp Ki ð yPID  yc Þ ¼ 0
yPID ¼ yc ðsimpler formÞ (2.43)
 2  
Power control: d pg  pc d pg  pc
e p Kd þ e K
p p
d2t dt (2.44)
  dpc dyPID
þ ep Ki pg  pc  þ ¼0
dt dt
  dpc dyPID
ep Ki pg  pc  þ ¼ 0ðsimpler formÞ (2.45)
dt dt
30 Modeling and dynamic behaviour of hydropower plants

1
0.9
Guide vane opening [pu] 0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0 20 40 60 80 100 120
Servomotor stroke [mm]

Figure 2.9 Nonlinear relation between servomotor stroke and angular opening of
guide vanes

Table 2.3 Status of selectors in different control modes [14]

Control mode Equation S1 status S2 status S3 status


Frequency control (2.40) 1 1 1
(2.41) 1 3 3
Opening control (2.42) 2 1 1
(2.43) 2 2 1
Power control (2.44) 2 3 3
(2.45) 2 2 3

For the servo part, the output opening ( yservo) is found by solving:
dyservo
yPID ¼ Ty þ yservo : (2.46)
dt
Besides, in the practical governor system, usually there is a nonlinear relation
between the servomotor stroke ( y) and angular opening of guide vane (GV) (agv),
which is described as
agv ¼ fy!gv ðyÞ; (2.47)
and a real case is demonstrated in Figure 2.9. However, in this chapter, the non-
linear relation is ignored for the simplicity of the case study.
In the governor system, the selectors (S1, S2, and S3) are related to each other,
and various status (or states) of selectors in different control modes are shown in
Table 2.3. The status in Figure 2.8 demonstrates the frequency control under
opening feedback (OF), which is described by (2.40). Table 2.4 collects the equa-
tion sets of hydropower units under diverse operating conditions.
Time-domain modeling and a case study 31

Table 2.4 Equation set of the hydropower unit under different operating
conditions [14]

Operating condition Equation set

Governor Generator Turbine


Normal operation Frequency control (2.40) or (2.41) (2.36) or (2.37) (2.28)–(2.35)
Opening control (2.42) or (2.43) (2.36) or (2.37)
Power control (2.44) or (2.45) (2.36) or (2.37)
Start-up Open-loop (2.43) (2.38)
Closed-loop (2.40) (2.38)
No-load operation (2.40) (2.38)
Emergency stop (2.43) (2.38)
Load rejection (2.40) (2.38)

2.2.3 Features of the model


The fundamental features of the model in TOPSYS are listed as follows:
● Compressible flow is included in the draw water tunnel and penstock, con-
sidering the elasticity of water and pipe wall.
● Diverse types of surge tanks and tunnels are utilized. In this chapter, only basic
boundary conditions of the piping system in the TOPSYS model are presented.
In other complex conditions, which are included in TOPSYS, are not shown
here, e.g., free-surface flow in tailrace tunnel, a more complicated reservoir
with an extra forebay, ball valve, and various forms of surge tanks consider
ventilation, overflow and long-connection at the bottom, etc.
● Characteristic curves of turbines are implemented, instead of applying six
simplified transmission coefficients.
● The governor model contains different control modes, enabling the simulation
of regulation and operation in various conditions. The essential nonlinear
factors in the governor system, i.e., dead-zone, saturation, rate limiting, and
backlash are also considered.
Other discussions and features of the software TOPSYS, as well as the comparison
between simulations and on-site measurements, can be found in [14].

2.3 Practical engineering case


In this chapter, a real Chinese HPP is selected for the case study in the next section.
The HPP contains long-draw water tunnels, four surge tanks, and eight generating
units with Francis turbines. The whole power plant is divided into four individual
and equivalent ‘‘groups,’’ each ‘‘group’’ consists of a surge tank and two units, as
shown in Figure 2.10. Two identical generating units (U1 and U2 corresponding to
the J16 and the J17 in the figure) share the common upstream pipeline, and this
32 Modeling and dynamic behaviour of hydropower plants

Figure 2.10 Interface of the software TOPSYS and the model of the study case in
this chapter. The large Chinese hydropower plant has surge tank, a
long draw water tunnel and Francis turbines. Two generating units
share the common upstream pipeline.

Table 2.5 Basic information of each hydropower unit in the HPP (J16 and J17)

Rated power Rated water Rated discharge Rated rotation Inertia time
[MW] head [m] [m3/s] speed [r/min] constant Ta [s]
610 288 228.6 166.7 9.46

could cause the hydraulic disturbances. This ‘‘group’’ is chosen as the main study
objective. The basic information of the generating units is presented in Table 2.5;
detailed data of the pipelines and featured values of the piping system in the HPP
are shown in Tables 2.6 and 2.7, respectively. The default parameter settings of the
governor actuator are listed in Table 2.8.

2.4 Case study of various dynamic processes of


hydropower plant
In this section, case studies are conducted under diverse dynamic processes of the
HPP. It is worth emphasizing that the numerical model has already been examined
by comparison between the simulations and on-site measurements in [14]. The
Case 2 in [14] is the case in this chapter, whereas in [14], only relatively simple
conditions are analyzed. The case study here is based on a real consultant work; all
the regulation and operation cases are from the practical tasks that are highly
concerned by the owner and design company of the Chinese HPP.
Time-domain modeling and a case study 33

Table 2.6 Detailed data of the pipelines in the HPP. The pipeline numbers in the
first column correspond to the number near the pipeline model in
Figure 2.10.

Pipeline Length [m] Cross-sectional area [m2] Description


L1 100.00 115.35 \
L2 1,697.26 113.12 \
L3 4,202.73 113.12 \
L4 10,300.00 113.44 \
L5 323.18 103.56 \
L6 39.00 103.56 \
L7, L8 17.35 44.18 \
L9, L10 36.28 42.58 \
L11, L12 359.06 33.18 \
L13, L14 118.00 33.18 \
L15, L16 26.64 24.45 Volute
L17, L18 29.73 25.99 Draft tube-1
L19, L20 32.20 77.38 Draft tube-2
L21, L22 133.25 114.55 \
L23, L24 121.31 114.55 \

Table 2.7 Featured values of the piping system in the HPP. Tw means the water
starting time constant.

Upstream reservoir Surge tank Turbine to Surge tank Surge fluctuation


to surge tank, to turbine, downstream to reservoir, period [s]
Tw1 [s] Tw2 [s] reservoir, Tw3 [s] Tw2 þ Tw3 [s]
23.88 1.35 0.30 1.65 496

Table 2.8 Default parameter settings of the governor actuator in this chapter

Parameter Servo (Ty) Saturation Rate limiting Backlash Dead zone


Value 0.2 [0.1 pu] 0.1 pu/s 0 0

2.4.1 Start-up and no-load operation


A rapid, stable, and safe start-up process of hydropower unit has been a critical
factor of power system stability and power quality. In [14], different start-up modes
(open-loop, closed-loop and ‘‘open-loop þ closed-loop’’ start-up) are introduced.
The most common mode is the ‘‘open-loop þ closed-loop’’ method, which is
described below. First, the rotational speed increases to a certain set value under
34 Modeling and dynamic behaviour of hydropower plants

opening control that is in open-loop mode; then, the turbine governor automatically
switches to frequency control, that is, closed-loop mode, to stabilize the speed at
the rated value. The set value determining the point of automatic mode switch is
usually set to larger than 80% of the rated speed.
In order to ensure a rapid, stable, and safe start-up process, the regulation
strategy can be improved in the following aspects:
1. Given guide vane opening (GVO) in opening control (open-loop) stage
2. Switch point between the two control modes
3. Governor parameters in frequency control (closed-loop) stage
4. Given frequency (rotational speed) in frequency control (closed-loop) stage
In this subsection, based on the consultant project for the case HPP, three different
start-up strategies (slow, medium, and fast start-up) for the unit U1 are compared.
The operation requirements and detailed settings are presented below.
1. Requirements:
(i) Achieve a short start-up time;
(ii) Avoid extremely rapid changes of the volute pressure;
(iii) Keep low hydraulic disturbances on the other unit (U2), which share the
same piping system.
2. Simulation settings:
(i) Three different settings of GVO in opening control (open-loop) stage are
suggested by the turbine producer, as shown in Figures 2.11–2.13. Only
the unit U1 is in start-up process, the U2 is in the normal operation on
rated condition.
(ii) The switch point between the opening control and frequency control is
95% of the rated frequency, i.e., 47.5 Hz.
(iii) Governor parameters are Kp ¼ 2.0, Ki ¼ 0.25, Kd ¼ 1.2, and bp ¼ 0.0.
(iv) Given frequency: normally the given frequency is the rated value of the
local power system, e.g., 50 Hz; while in this start-up process, the given
value is set to increase from 95% to 100% of the rated value in 25 s and
then keep constant. This setting aims to guide the real rotational speed
reach the rated value smoothly.

The simulation results are shown in Figures 2.11–2.13. The start-up time is
mainly determined by the settings of GVO in the open-loop stage. In the fast strat-
egy, the rapid opening of the GVO leads to the fastest increase process of rotational
speed. The overshoot of the speed in all three strategies are controlled well, due to
the rational settings of governor parameters and the given frequency. Meanwhile,
the fast increase of GVO results in larger changes in volute pressure and the power
output of the other unit sharing the same piping system (U2): The volute pressure is
directly influenced by the rapid movements of GV due to the water hammer [16].
In the slow strategy, a sharp pressure drop occurs in the initial stage caused by the
very quick change of GVO in the first 2 s.
Time-domain modeling and a case study 35

0.4 1.2
0.35
1

Guide vane opening [pu]

Rotational speed [pu]


0.3
GV opening 0.8
0.25
Rotational speed
0.2 0.6
0.15
0.4
0.1
0.2
0.05
0 0
0 50 100 150 200
(a) Time [s]

560 360
Active power of U2 [MW]

555 340

Volute pressure [m]


550
320
545
300
540

Power-U2 280
535
Volute pressure
530 260
0 50 100 150 200
(b) Time [s]

Figure 2.11 Simulation results under the slow start-up strategy: (a) GVO and
rotational speed ( frequency); (b) volute pressure of the start-up
unit and active power of the other unit sharing the same piping
system (U2)

2.4.2 Grid-connected operation


2.4.2.1 Primary frequency control
Primary frequency control (PFC) means that the process in which generating units
change their power output automatically according to the grid frequency fluctua-
tion, to make the active power balanced in the power system again. In [17], more
relevant information, e.g., different technical standards in different countries, and
the research background of the PFC are introduced; the implementation of the
mathematical model for PFC is discussed in [14].
In this subsection, based on the consultant work for the study case, the PFC
performance of the unit is examined by step change tests under OF and power
36 Modeling and dynamic behaviour of hydropower plants

0.4 1.2
0.35
Guide vane opening [pu]
1
0.3

Rotational speed [pu]


GV opening 0.8
0.25
Rotational speed
0.2 0.6
0.15
0.4
0.1
0.2
0.05
0 0
0 50 100 150 200
(a) Time [s]
560 360
Active power of U2 [MW]

555 340

Volute pressure [m]


550
320
545
300
540

535 Power-U2 280


Volute pressure
530 260
0 50 100 150 200
(b) Time [s]

Figure 2.12 Simulation results under the medium start-up strategy: (a) GVO and
rotational speed (frequency); (b) volute pressure of the start-up unit
and active power of the other unit sharing the same piping system (U2)

feedback (PF). The operation requirements and detailed settings are presented
below.
1. Requirements: according to the specifications of China Electricity Council
[18], the most crucial requirements are:
(i) The power adjustment quantity should reach 90% of the static char-
acteristic value within 15 s.
(ii) If the rated head of the unit is larger than 50 m, the power delay time [17]
should be less than 4 s.
2. Simulation settings:
(i) The frequency step change is 0.1 Hz (0.002 pu), and it occurs at 10 s.
Only the unit U1 is in the PFC, the GV of the unit U2 keeps constant.
(ii) Governor parameters under both feedback modes are Kp ¼ 1.0, Kd ¼ 0.0,
bp ¼ 0.04 and two values of Ki are selected, i.e., Ki ¼ 5.0 or 0.417. The
parameter set (Kp ¼ 1.0, Ki ¼ 0.417, Kd ¼ 0.0, bp ¼ 0.04) is a standard
setting [19] of Swedish HPPs owned by Vattenfall, the largest hydro-
power owner and operator in Sweden.
Time-domain modeling and a case study 37

0.4 1.2

Guide vane opening [pu]


0.35 1

Rotational speed [pu]


0.3
GV opening 0.8
0.25
Rotational speed
0.2 0.6
0.15 0.4
0.1
0.05 0.2
0 0
0 50 100 150 200
(a) Time [s]
560 360
Active power of U2 [MW]

555 340

Volute pressure [m]


550
320
545
300
540
Power-U2 280
535
Volute pressure
530 260
0 50 100 150 200
(b) Time [s]

Figure 2.13 Simulation results under the fast start-up strategy: (a) GVO and
rotational speed (frequency); (b) volute pressure of the start-up
unit and active power of the other unit sharing the same piping
system (U2)

The simulation results are shown in Figures 2.14, 2.15 and Table 2.9. The
results support the conclusion in [17] that the response time highly depends on
the governor parameters, especially the value of Ki. Another crucial point is the
performance difference between the OF and PF. For the static characteristics, under
the 0.002 pu frequency change and the value of bp which is 0.04, the OF and the
PF lead to the regulation of 5% GVO increase and 5% power increase, respectively,
therefore the regulation targets of these two modes are different. What is more
important is the dynamic process of the two modes. Under the OF, the GVO keeps
constant after reaching the target value, resulting in the power fluctuation due to the
water level change in the surge tank, and the overshoot of the power output is more
obvious. While under the PF, the power fluctuation is controlled well and only the
power oscillation with small amplitude occurs, due to the surge in the downstream
gate shaft.

2.4.2.2 Automatic generation control (secondary frequency control)


The main objective of the automatic generation control (secondary frequency
control) is adjusting the power output of the generating units to regulate grid
38 Modeling and dynamic behaviour of hydropower plants

530 0.65
525
0.64

Guide vane opening [pu]


520
Active power [MW]

515 0.63
510 Power-OF
0.62
505 Power-PF
500 GVO-OF 0.61
495 GVO-PF
0.6
490
485 0.59
0 50 100 150 200
Time [s]

Figure 2.14 Simulation results of GVO and active power under frequency control
with opening feedback and power feedback (Ki ¼ 0.417)

530 0.65
525

Guide vane opening [pu]


0.64
Active power [MW]

520
515 0.63
510 Power-OF
0.62
505 Power-PF
500 GVO-OF 0.61
495 GVO-PF
0.6
490
485 0.59
0 50 100 150 200
(a) Time [s]

1,639 1,334
Surge in surge tank-OF
Surge in surge tank [m]

Surge in surge tank-PF


Surge in gate shaft [m]

1,638 Surge in gate shaft-OF 1,333


Surge in gate shaft-PF
1,637 1,332

1,636 1,331

1,635 1,330
0 50 100 150 200
(b) Time [s]

Figure 2.15 Simulation results of frequency control with opening feedback and
power feedback (Ki ¼ 5.0): (a) GVO and active power; (b) water
level fluctuation in the surge tank and the downstream gate shaft
Time-domain modeling and a case study 39

Table 2.9 Response time (Tr) of different PFC processes. The


response time [17] means the time when the active power
output reaches the target value, which corresponds to 90%
of the regulation amount, after the frequency step change.

Opening feedback Power feedback


Ki 5.0 0.417 5.0 0.417
Tr 13.2 s 272.2 s 5.8 s 120.6 s

frequency to the nominal value. This function is also referred to as load-frequency


control (LFC) [20].
In this subsection, based on the consultant work for the HPP, the transient
processes in secondary frequency control are analyzed with different strategies
under opening control and power control. The operation case is that the power
output increase from 0% to 90% of the rated value, i.e., to 549 MW. The operation
requirements and detailed settings are listed below.
1. Requirements:
(i) Ensure the regulation process is rapid and stable;
(ii) Avoid large power fluctuations of the other unit, which shares the same
piping system, due to the hydraulic disturbance.
2. Simulation settings:
(i) For each of two feedback modes, three strategies for power output
increase are adopted:
(a) the given value (GVO or active power) ascend linearly with a rate
that is 0.033 pu/s (1/30 pu/s);
(b) the given value ascend linearly with a rate that is 0.0083 pu/s (1/120
pu/s);
(c) the given value ascend in steps.
The three strategies correspond to Figures 2.16, 2.17, and 2.18, respectively.
(ii) The simulations of opening control and the power control adopt the
simpler form described as (2.43) and (2.45), respectively. Governor
parameters under the power control are Ki ¼ 0.25 and bp ¼ 0.04. Only the
unit U1 is in the secondary frequency control, the unit U2 is in the power
control mode with a constant given power.
The simulation results are shown in Figures 2.16–2.18. The power control
leads to a faster response of the power output; however, the overshoot also occurs,
especially under the strategy of fast linear increase. In terms of hydraulic dis-
turbance on the other unit (U2), the strategy of step increase outperforms the linear
strategies: the power fluctuation of the U2 is relatively small under the strategy of
step increase.
40 Modeling and dynamic behaviour of hydropower plants

700 1

600

Guide vane opening [pu]


0.8
Active power [MW] 500
400 0.6

300 0.4
Power
200
Power-U2 0.2
100 GVO
0 0
0 500 1,000 1,500
(a) Time [s]

700 1
600

Guide vane opening [pu]


0.8
Active power [MW]

500

400 0.6

300 0.4
200 Power
Power-U2 0.2
100
GVO
0 0
0 500 1,000 1,500
(b) Time [s]

Figure 2.16 Simulation results of GVO and active power of the unit in AGC and
active power of the other unit which shares the same piping system
(U2), under the strategy of fast linear increase: (a) opening control;
(b) power control

2.4.3 Isolated operation


The isolated operation means that a generating unit operates without being inter-
connected with other generating units [21]. Simulation and analysis are conducted
in isolated operation to examine the stability of the HPP system. In this subsection,
the aim is to analyze the stability issue caused by the surge tank. More information
and the research background of this topic can be found in [22].
According to the formula of critical cross-sectional area of the surge tank [23]
based on Thoma criterion [24,25]:

LA LA
ATh ¼ ¼ ; (2.48)
2g ða þ 1=2g ÞðH0  hw0  3hwm Þ 2gða þ 1=2g ÞðH1  2hwm Þ
Time-domain modeling and a case study 41

700 1

600

Guide vane opening [pu]


0.8

Active power [MW]


500
400 0.6

300 0.4
Power
200
Power-U2 0.2
100 GVO
0 0
0 500 1,000 1,500
(a) Time [s]

700 1
600

Guide vane opening [pu]


0.8
Active power [MW]

500
400 0.6

300 0.4
200 Power
Power-U2 0.2
100 GVO
0 0
0 500 1,000 1,500
(b) Time [s]

Figure 2.17 Simulation results of GVO and active power of the unit in AGC and
active power of the other unit which shares the same piping system
(U2), under the strategy of slow linear increase: (a) opening control;
(b) power control

the calculation results for the case HPP are shown in Table 2.10. The critical
area of the surge tank of this HPP is 400.13 m2 and the safety factor (n ¼ 1.04) is
rather small.
Based on the consultant work for the study case, the dynamic response after a
load step change is simulated to analyze the stability of the HPP. The operation
requirements and detailed settings are presented below.
1. Requirements:
(i) Ensure the stability of the whole system in the HPP
(ii) Achieve an acceptable settling time of the rotational speed (frequency)
after the disturbance
42 Modeling and dynamic behaviour of hydropower plants

700 1

600

Guide vane opening [pu]


0.8
Active power [MW]
500

400 0.6

300 0.4
Power
200
Power-U2 0.2
100 GVO
0 0
0 500 1,000 1,500
(a) Time [s]

700 1

600

Guide vane opening [pu]


0.8
Active power [MW]

500

400 0.6

300 0.4
200 Power
Power-U2 0.2
100
GVO
0 0
0 500 1,000 1,500
(b) Time [s]

Figure 2.18 Simulation results of GVO and active power of the unit in AGC
and active power of the other unit which shares the same piping
system (U2), under the strategy of step increase: (a) opening control;
(b) power control

Table 2.10 Calculation of the critical cross-sectional area of the surge tank in the
case HPP

Upstream Downstream Average hw0 hwm Coefficient Net head Thoma Real Safety
water water velocity, [m] [m] of head [m] critical area factor,
level [m] level [m] v [m/s] loss, area [m2] n [pu]
a [pu] [m2]

1,640 1,333.74 4.007 12.642 2.512 0.787 290.886 400.13 416.2 1.04
Time-domain modeling and a case study 43

168
n = 1.04
n = 1.20
Rotational speed [r/min]

167.5 ±0.2% Bandwidth n = 1.40

167

166.5

166
0 2,000 4,000 6,000 8,000 10,000
Time [s]

Figure 2.19 Simulated rotational speed ( frequency) after the load disturbance
under different values of the cross-sectional area of the
surge tank

2. Simulation settings:
(i) The control mode is the frequency control under OF, as described
in (2.40). Governor parameters are Kp ¼ 2.0, Ki ¼ 0.25, and Kd ¼ 1.2,
bp ¼ 0.01.
(ii) A simple sensitivity analysis on the cross-sectional area of the surge tank
is conducted. Three values of the area corresponding to three safety
factors, n, are selected, i.e., n ¼ 1.04 (the original value), n ¼ 1.20 and
n ¼ 1.40.
(iii) The operation case is the suggested case in [22]: load (or power output)
ascends from 90% to 100% of rated power under the lowest water head.
This case is very unfavorable to the system stability.
As shown in Figure 2.19, the cross-sectional area of the surge tank highly
impact the stability of the system, and a cross-sectional area close to the critical
value (when n ¼ 1.04) leads to an un-damped oscillation. The aim of this subsec-
tion is only to present the time-domain analysis method and emphasize the
importance of examining the stability issue caused by surge tank, more detailed
studies on the hydropower system stability under diverse conditions can be found in
[22,26–28].

2.4.4 Emergency stop and load rejection


The large disturbance (severe transient disturbance) is the most dangerous and
important condition, and highly concerns the safety of HPPs. It can be separated
into two conditions, emergency stop and load rejection, as introduced in Tables 2.4
and 2.11. In [14], related introduction and analysis are presented.
In this subsection, based on the consultant work for the case, the safety of the
HPP under the two large disturbance conditions is discussed. The main objective of
44 Modeling and dynamic behaviour of hydropower plants

Table 2.11 Main information and technique requirements for the emergency stop
and load rejection. The values of the requirements are only for the
case HPP in this chapter.

Condition: Emergency stop Load rejection


Control mode: Opening control Frequency control
Main requirements: Ensure the following important values in the safe range:
(1) Volute pressure <410 m.w.c.
(2) Pressure at draft tube > 6.5 m.w.c.
(3) Rotational speed of the unit <150% pu
(4) Water lever in the surge tank: 1,576.2 m < the surge < 1,696.5 m
(5) Other pressure values, e.g., pressure along the pipelines, pressure
inside the surge tank
Main optimization GV closure law Governor parameters and the rate limit
object: of the GV movement

Table 2.12 Control strategies of the emergency stop and the load rejection

Condition Strategy Description


Emergency stop (opening control) OC1 Linear closing law: GVO decrease from
rated value to 0 in 10 s
OC2 Linear closing law: GVO decrease from
rated value to 0 in 15 s
OC3 Broken-line closing law: GVO decrease
from rated value to 0 in 15 s
Load rejection (frequency control) FC1 The rate limit is 12.5%/s
FC2 The rate limit is 6.67%/s

analysis is the control strategies, e.g., the GV closure law and the rate limit.
The detailed settings are presented below.
1. Simulation settings:
(i) The control strategies of the two conditions are shown in Table 2.12.
(ii) Under the frequency control, the governor parameters are Kp ¼ 2.0, Ki ¼
0.25, Kd ¼ 0, and bp ¼ 0.01.
(iii) The emergency stop or load rejection happens simultaneously on two
units when they are in the rated operating point (rated power output
under rated water head).
The simulation results are demonstrated in Figure 2.20 and Table 2.13. The
results show the direct influence of the control strategies of GVO on the safety of
the whole power plant. Therefore, the optimization of the strategies through the
Time-domain modeling and a case study 45

0.9
0.8 OC1

Guide vane opening [pu]


0.7 OC2
0.6 OC3
0.5
FC1
0.4
0.3 FC2
0.2
0.1
0
0 10 20 30 40 50
(a) Time [s]

1.6
OC1 OC2 OC3
FC2
Rotational speed [pu]

1.4 FC1

1.2

0.8

0.6
0 20 40 60 80 100
(b) Time [s]

Figure 2.20 Simulation results of the emergency stop and the load rejection under
different strategies: (a) GVO and (b) rotational speed

Table 2.13 Simulation results of important indicators of safety

Strategy Max. volute Min. pressure at Max. rotational Highest Lowest


pressure [m] draft tube [m] speed [pu] (%) surge [m] surge [m]
OC1 370.05 4.08 142.84 1,679.3 1,604.11
OC2 363.98 6.12 150.21 1,679.32 1,604.06
OC3 365.23 5.58 147.37 1,679.31 1,604.08
FC1 392.29 0.23 135.48 1,678.11 1,605.67
FC2 374.08 5.54 147.21 1,678.13 1,605.62

trial simulations is exceedingly (or extremely) necessary. The goal of this section is
only a brief introduction of the analysis on the large disturbance condition, hence
just one operating case is discussed; while in the real consultant study, normally
more extreme cases, e.g., different units stop one after another, need to be
considered.
46 Modeling and dynamic behaviour of hydropower plants

2.5 Conclusions
In this chapter, a numerical model of HPPs, including the model of piping system
and generating units, is introduced, and its features are presented. Based on the
consultant project of areal Chinese HPP, the case study is conducted for various
operating conditions (e.g., start-up process and no-load operation, grid-connected
operation, isolated operation, emergency stop, and load rejection). Diverse opera-
tion requirements and different simulation settings are presented and discussed.
The main methods and procedures of the time-domain analysis are introduced, and
the necessity of prediction of the dynamic behavior of HPPs is demonstrated. The
results show the influence of various factors (from both control system and piping
system) on different transient processes, supplying the guidance for understanding
and optimizing the quality and stability of regulation and operation of HPPs.

Acknowledgments
The authors thank the China Scholarship Council (CSC) and Stand Up for Energy.
The research presented was also carried out as a part of ‘‘Swedish Hydropower
Centre – SVC.’’ SVC has been established by the Swedish Energy Agency, Elforsk
and Svenska Kraftnät together with Luleå University of Technology, KTH Royal
Institute of Technology, Chalmers University of Technology and Uppsala Uni-
versity (www.svc.nu). The authors also acknowledge the support from the National
Natural Science Foundation of China under Grant No. 51379158.

References
[1] F. Demello, R. Koessler, J. Agee et al. ‘‘Hydraulic-turbine and turbine
control-models for system dynamic studies,’’ IEEE Transactions on Power
Systems, vol. 7, pp. 167–179, 1992.
[2] E. De Jaeger, N. Janssens, B. Malfliet, and F. Van De Meulebroeke, ‘‘Hydro
turbine model for system dynamic studies,’’ IEEE Transactions on Power
Systems, vol. 9, pp. 1709–1715, 1994.
[3] O. Souza Jr, N. Barbieri, and A. Santos, ‘‘Study of hydraulic transients in
hydropower plants through simulation of nonlinear model of penstock
and hydraulic turbine model,’’ IEEE Transactions on Power Systems,
vol. 14, pp. 1269–1272, 1999.
[4] H. Fang, L. Chen, N. Dlakavu, and Z. Shen, ‘‘Basic modeling and simulation
tool for analysis of hydraulic transients in hydroelectric power plants,’’ IEEE
Transactions on Energy Conversion, vol. 23, pp. 834–841, 2008.
[5] Y. Zeng, Y. Guo, L. Zhang, T. Xu, and H. Dong, ‘‘Nonlinear hydro turbine
model having a surge tank,’’ Mathematical and Computer Modelling of
Dynamical Systems, vol. 19, pp. 12–28, 2013.
[6] B. Strah, O. Kuljaca, and Z. Vukic, ‘‘Speed and active power control
of hydro turbine unit,’’ IEEE Transactions on Energy Conversion, vol. 20,
pp. 424–434, 2005.
Time-domain modeling and a case study 47

[7] C. Nicolet, B. Greiveldinger, J.-J. Hérou, et al., ‘‘High-order modeling of


hydraulic power plant in islanded power network,’’ IEEE Transactions on
Power Systems, vol. 22, pp. 1870–1880, 2007.
[8] C. Nicolet, ‘‘Hydroacoustic modelling and numerical simulation of unsteady
operation of hydroelectric systems,’’ École Polytechnique Fédérale de
Lausanne, 2007.
[9] S. Mansoor, D. Jones, D. A. Bradley, F. Aris, and G. Jones, ‘‘Reproducing
oscillatory behaviour of a hydroelectric power station by computer simula-
tion,’’ Control Engineering Practice, vol. 8, pp. 1261–1272, 2000.
[10] J. I. Pérez-Dı́az, J. I. Sarasúa, and J. R. Wilhelmi, ‘‘Contribution of a
hydraulic short-circuit pumped-storage power plant to the load–frequency
regulation of an isolated power system,’’ International Journal of Electrical
Power & Energy Systems, vol. 62, pp. 199–211, 2014.
[11] N. Kishor, R. Saini, and S. Singh, ‘‘A review on hydropower plant
models and control,’’ Renewable and Sustainable Energy Reviews, vol. 11,
pp. 776–796, 2007.
[12] G. A. Munoz-Hernandez, S. P. Mansoor, and D. I. Jones, Modelling and
Controlling Hydropower Plants. Springer, London, UK, 2013.
[13] S. Wei, Hydraulic Turbine Regulation (in Chinese). Huazhong University of
Science and Technology Press, Wuhan, 2009.
[14] W. Yang, J. Yang, W. Guo, et al., ‘‘A mathematical model and its applica-
tion for hydro power units under different operating conditions,’’ Energies,
vol. 8, pp. 10260–10275, 2015.
[15] V. L. Streeter and E. B. Wylie, Fluid Transients. McGraw-Hill, New York,
1978.
[16] W. Yang and J. Yang ‘‘Research on the volute pressure in start-up process of
hydroelectric generating units,’’ in IOP Conference Series: Earth and
Environmental Science, 2012, p. 042018.
[17] W. Yang, J. Yang, W. Guo, and P. Norrlund, ‘‘Response time for primary
frequency control of hydroelectric generating unit,’’ International Journal of
Electrical Power & Energy Systems, vol. 74, pp. 16–24, 2016.
[18] China Electricity Council, DL/T 1040-2007: The grid operation code
(in Chinese), 2007.
[19] L. Saarien, ‘‘A hydropower perspective on flexibility demand and grid fre-
quency control,’’ Licentiate, Department of Engineering Sciences, Uppsala
University, Uppsala, 2014.
[20] P. Kundur, N. J. Balu, and M. G. Lauby, Power System Stability and
Control, vol. 7. McGraw-Hill, New York, 1994.
[21] Energy Development and Power Generation Committee of the IEEE Power
Engineering Society, ‘‘IEEE guide for the application of turbine governing
systems for hydroelectric generating units,’’ IEEE Std 1207-2011 (Revision
to IEEE Std 1207-2004), pp. 1–131, 2011.
[22] W. Yang, J. Yang, W. Guo, and P. Norrlund, ‘‘Frequency stability of iso-
lated hydropower plant with surge tank under different turbine control
modes,’’ Electric Power Components and Systems, vol. 43, pp. 1707–1716,
2015.
48 Modeling and dynamic behaviour of hydropower plants

[23] China Electricity Council, DL/T5058-1996: Specifications for design of


surge chamber of hydropower stations (in Chinese), 1997.
[24] C. Jaeger, Fluid Transients in Hydro-Electric Engineering Practice. Blackie,
London, UK, 1977.
[25] G. I. Krivchenko, ‘‘Admissibility of deviating from the Thoma criterion
when designating the cross-sectional area of surge tanks,’’ Hydrotechnical
Construction, vol. 22, pp. 403–409, 1988/07/01 1988.
[26] W. Guo, J. Yang, W. Yang, J. Chen, and Y. Teng, ‘‘Regulation quality for
frequency response of turbine regulating system of isolated hydroelectric
power plant with surge tank,’’ International Journal of Electrical Power &
Energy Systems, vol. 73, pp. 528–538, 2015.
[27] W. Guo, J. Yang, J. Chen, W. Yang, Y. Teng, and W. Zeng, ‘‘Time response
of the frequency of hydroelectric generator unit with surge tank under iso-
lated operation based on turbine regulating modes,’’ Electric Power Com-
ponents and Systems, vol. 43, pp. 2341–2355, 2015.
[28] W. Guo, J. Yang, J. Chen, and M. Wang, ‘‘Nonlinear modeling and
dynamic control of hydro-turbine governing system with upstream surge
tank and sloping ceiling tailrace tunnel,’’ Nonlinear Dynamics, vol. 84(3),
pp. 1383–1397, 2016.
Chapter 3
Reduced order models for grid connected
hydropower plants
Gérard Robert 1 and Frédéric Michaud1

3.1 Introduction
This chapter aims at modeling a hydropower plant connected to a power system.
A practical engineering approach based on physical models (hydraulic, mechanical
and electrical) is proposed. With necessarily a low order, these models are dedi-
cated for generation control design and more particularly for power-frequency
controller design with issues presented in Chapter 5.
Hydroelectric power plant (HPP) is a non-linear system that can be divided
into two infinite dimensional subsystems, which remain complex to model the
hydropower plant and the electric power system wherein the HPP is connected.
The former is a specific generating unit facility which behaves as a non-minimum
phase with water oscillations in the hydraulic circuit described by partial differ-
ential equations with uncertain friction coefficients. The latter is an interconnected
multi-machine system with a huge number of uncertain parameters (synchronous
machine parameters, controllers’ parameters, line impedances, etc.) and a variable
structure (load changes, outage of transmission lines or generating units, etc.).
The complexity of the model and the corresponding accuracy of the results
depend on its use. For simulation studies, a fine model is expected because the need
is to match simulations results with field test results. In this case, the model needs a
good accuracy both for the hydraulic process and for the control system, taking
into account all nonlinearities (characteristic curves of the hydraulic turbine, head
losses, water hammer effect, dead band, saturations, anti-windup, computational
delays, logic functions, filters, etc.). On the contrary, detailed non-linear models
with many parameters are not suitable for control design because they give com-
plex calculations which are time consuming for controller tuning and yield a design
which is not always robust in respect to parametric uncertainties. A better approach
is to choose a simple model yet representative enough to give the dominant time
constants and the main dynamics.

1
EDF Hydropower Generation and Engineering, CIH, Savoie Technolac, 73373 Le Bourget du Lac,
France
50 Modeling and dynamic behaviour of hydropower plants

Usually, the procedure to develop a process model for controller design is to


take into account all the dynamic modes leading to a high-order detailed model and
then to simplify the model by using reduction techniques. In this chapter, an
opposite approach is proposed by exhibiting dominant dynamics immediately from
the physical equations in order to obtain directly a low-order model with no need to
use reduction techniques.
The steps involved in reduction from infinite dimensional to low-order HPP
and power system models is given in a comprehensive manner in justifying all
simplification stages in accordance with [1]. A non-linear and a linearized model of
HPP equipped with a surge tank are given in Section 3.2 by detailing all the main
components of the HPP. An equivalent electric circuit is also proposed from a
hydraulic–electric analogy in order to allow an easy modeling of complex
hydraulic circuits. In Section 3.3, a general power system model uncovered by the
literature is developed in transfer function form and then is focused on two parti-
cular operation modes: interconnected and isolated modes. Section 3.4 presents a
complete state-space model of a HPP connected to any type of synchronous power
system, and finally, Section 3.5 gives a deep analysis of the dynamic behaviour
with the help of perturbation method.

3.2 Hydropower plant model


This section is concerned with obtaining a detailed model of a hydropower plant
consisting of a reservoir, a tunnel, a surge tank, a single penstock, several turbines
and an output canal (see Figures 3.1 and 3.7). The purpose is to develop a repre-
sentative low-order model defined by known parameters (geometry, characteristic
curve of the hydraulic turbine, etc.) and applicable in industrial world for controller
design or simulation studies.

A
C

Qa B Qc

Hb Z

E
D

Figure 3.1 Hydraulic–electric analogy for an impulse turbine


Reduced order models for grid connected hydropower plants 51

A methodology based on hydraulic–electric analogy is proposed in [1,2] to


quickly construct a hydraulic model especially if the topology of the HPP is com-
plex (e.g. two water supplying reservoirs, two surge tanks or many penstocks).
Non-linear equations and linear state space model are given for a HPP including
any type of turbines (impulse or reaction). This model is validated by comparison
with real-world field tests carried out on large hydropower plants (see Chapter 5).

3.2.1 Penstock and tunnel models


The way to transform an infinite model order of penstock to a low-order model is
considered in this section.
The penstock is made of steel, and small signal transients due to primary
power-frequency control are likely to excite elastic modes of long penstocks. The
frequency domain of the control input is typically inferior to 0.1 rad/s. In these
conditions, elastic modes are often not too much excited (natural frequencies are
usually higher than 0.1 rad/s) and justify the use of an inelastic model. Nevertheless,
for some particular geometry of penstocks, the first natural frequency of the pen-
stock can be lower than 0.1 rad/s, and thus necessitates the need of an elastic model.
The partial differential equations system linking the flow Q and the head H in a
pipe are similar to the ‘telegraph equations’ and are used for electrical transmission
line modeling. Following the assumptions of compressible water and elastic pipe, a
penstock with a length L and a cross-section S is described by two partial differ-
ential equations [3]:

8
> @Q @H g  S  K
>
< @t þ g  S @x þ QjQj¼0
L
(3.1)
> @H
> @Q
: gS þ a2 ¼0
@t @x

with L and S the length (m) and the section (m2), K the head losses coefficient
(m/(m3/s)2), and g the standard acceleration of gravity (m/s2).
For the need of developing a numerical simulator, a centred Euler-discretization
can be used for the ith pipe element of length dx:
 
@H  Hiþ1  Hi @Q  Qiþ1  Qi Qi þ Qiþ1
¼ ; ¼ ; Qiþð1=2Þ ¼
@x iþð1=2Þ dx @x iþð1=2Þ dx 2

Hence, (3.1) becomes:


8
> dx dQiþ1 KjQjdx dx dQi KjQjdx
>
> Hiþ1  Hi þ 2gS dt þ 2L Qiþ1 þ 2gS dt þ 2L Qi ¼ 0
<
(3.2)
>
> dH a2
>
: iþð1=2Þ þ ðQiþ1  Qi Þ ¼ 0
dt gSdx
52 Modeling and dynamic behaviour of hydropower plants

Qi Qi+1

dRh · dx/2 dLh · dx/2 dRh · dx/2 dLh · dx/2

dCh · dx
Hi Hi+1

Figure 3.2 Penstock discretisation

Let us consider:
● dLh ¼ 1=gS the linear hydraulic inductance
● dRh ¼ KjQj=L the linear hydraulic resistance
● dCh ¼ gS=a2 the linear hydraulic capacity
From (3.2), Figure 3.2 illustrates the equivalent circuit for the pipe element [1,4].
Then, to obtain a complete hydraulic model, the equations in (3.2) are replaced
by a set of 2  N equations, where N is the number of elements in the penstock
(N can be chosen from the Courant–Friederichs–Lewy criterion for numerical sta-
bility). Therefore, by this simple method, the infinite dimensional system was
reduced to a finite one.
The elastic model described by (3.1) can be simplified by considering a uni-
form flow ð@Q=@x ¼ 0Þ in a rigid conduit with an incompressible water. This leads
to neglect the water hammer effect for obtaining an inelastic model featured with a
low order and hence suited for controller design. So, if Hi and Ho is the hydraulic
charge at the input and the output respectively, we find with the approximation
@H=@x ¼ ðHo  Hi Þ=L the hydraulic model (3.3) of an inelastic conduit used in
this chapter to represent the main dynamics associated to penstock and tunnel:

L dQ
Hi  Ho ¼  þ K  jQj  Q (3.3)
g  S dt

3.2.2 Surge tank model


Hydraulic transients can produce high pressure in the conduit system. To prevent
excessive pressure, a surge tank is constructed providing a storage volume via
which the flow can pass and a flow damping to the turbine. Different types of surge
tank exist, but we will consider only the simplest case illustrated in Figure 3.3.
For a surge tank with a constant cross-section Sz, the conservation of the mass
gives:

dZ
Sz ¼ Qi  Qo (3.4)
dt
Note that in the Section 3.2.4, the head losses in the surge tank will be neglected.
Reduced order models for grid connected hydropower plants 53

Qi Qo

Figure 3.3 Surge tank

3.2.3 Turbine model in a water column


There are three technologies of hydraulic machines which are commonly used in
hydropower stations: Pelton, Francis and Kaplan turbine. With respect to their
efficiency, we find Pelton turbine for high head work, Francis turbine for medium
head and Kaplan turbine for low head. The former is an impulse turbine which
works at atmospheric pressure (the wheel is not submerged). This means that the
mechanical energy is only converted from kinetic energy. On contrary, the two later
turbines are called reaction turbine because the wheel is submerged and converts
both potential and kinetic energy in mechanical energy.
This will lead to two distinct hydraulic equations, one for impulse turbine (3.5)
and one for reaction turbine (3.8), which will be unified to give a generic turbine
model (3.7), in which it will be shown that all kinds of turbines can be represented
by a pressure source equal to the net head.
As flow development in hydraulic machines is very complex, no analytical
model is available to represent the dynamics of a given turbine. Characteristic
curves of a hydraulic turbine (also called hill curves) are key data which determine
the behaviour and the performance of a turbine under different working conditions.
These curves are plotted from the result of test performed in static conditions.
Six variables are measured: W, H, Q, Pm, h and u. Both the discharge Q and the
mechanical power Pm depend on three independent variables: the net head H,
the gate opening u and the rotational speed W. These curves are non-linear (see the
qualitative form given in Figure 3.4) and must be taken into account to have a good
representative model. They are obtained either by reduced scale model (often when
the turbine is not built yet) or by field test in real scale.

3.2.3.1 Impulse turbine


The speed and the power provided by impulse turbine or the Pelton wheel are
controlled by adjusting the flow of water through a needle valve that increases or
decreases the nozzle opening as shown in Figure 3.5.
Without loss of generality, consider Figure 3.1, wherein we eliminate the surge
tank in order to simplify the demonstration. In addition, a uniform conduit will be
assumed between the reservoir and the turbine (length L and cross-section S).
54 Modeling and dynamic behaviour of hydropower plants

P H 2 > H1 Q H2 > H1

H1 H1

u u

Figure 3.4 Qualitative P/Q-u turbine characteristic curves for Wn

Buckets

Needle

Nozzle

Figure 3.5 Pelton turbine [Internet source]

Notice that the mass reference is located to the atmospheric pressure at point E
(Figures 3.1 and 3.7) which is the reference of the water levels Hb and Z.
Using the momentum equation [5,6] derived from the Newton’s second
law [7], assuming an incompressible fluid and a rigid pipe, the transient flow
between points A and E (Figure 3.1) can be described by the following differential
equation:

PA VA2 PE VE2 L dQ
Hb þ þ ¼ þ þ KAE jQjQ þ  (3.5)
rg 2g rg 2g Sg dt

where PA ¼ PE ¼ atmospheric pressure, KAE the head loss coefficient in conduits


and L=Sg  dQ=dt the inertial term of the flow.
Admitting a large surface of the reservoir, the velocity VA can be neglected:

VE2 L dQ
Hb ¼ þ KAE jQjQ þ  (3.6)
2g Sg dt

For a Pelton turbine, it is known thatpffiffiffiffiffiffiffiffiffi


the velocity yielded by the nozzle is
given by the quasi-static model VE ¼ 2gH with H is the net head such that
Reduced order models for grid connected hydropower plants 55

Figure 3.6 Francis turbine (on the left) and Kaplan turbine (on the right)
[Internet source]

H ¼ Hb  K jQjQ. Therefore, by noting K ¼ KAE the head loss coefficient in con-


duits (tunnel þ penstock), the following equation can be expressed1:

L dQ
Hb ¼ H þ K jQjQ þ  (3.7)
Sg dt

3.2.3.2 Reaction turbine


Figure 3.6 illustrates the scheme of a Francis turbine (on the left) and a Kaplan
turbine (on the right).
Again cancelling the surge tank in Figure 3.7, we consider the conduit system
between the points A and E. Thus, applying the momentum equation, the flow
equation for transients is:

PA VA2 PE VE2 L dQ
Hb þ þ ¼ þ þ H þ KAD jQjQ þ  (3.8)
rg 2g rg 2g Sg dt

As previously, PA ¼ PE and velocities are neglected in reservoirs (VA ¼ VE ¼ 0).


Thus, (3.8) related to a reaction turbine is reduced to (3.7), which is obtained for a
Pelton turbine. Note that for a reaction turbine, the gross head Hb is the deviation of
the level between free surfaces of the reservoir and the output canal while for a
Pelton turbine, Hb is the deviation between the free surface of the reservoir and the
axis of the turbine.
In conclusion, it was demonstrated that for any kind of turbine (impulse or
reaction), the turbine can be represented by a net head H in conduit systems leading
to a unified model.

1
Equation (3.8) is similar to the energy equation of Bernoulli (valid only for steady flow) corrected with
an inertial term.
56 Modeling and dynamic behaviour of hydropower plants

A
C

Qa B Qc

Hb Z

D E

Figure 3.7 Hydraulic–electric analogy for a reaction turbine

3.2.4 Hydraulic circuit model


A usual hydraulic circuit met in hydroelectric project is considered and illustrated
by Figures 3.1 and 3.7. It consists of a single tunnel, a single penstock, a single
surge tank, and n turbines assumed to be identical.
With hydraulic–electric analogy [1,4], it is possible to establish an equivalent
electric circuit by replacing the flow by a current, head by voltage, surge tank by a
capacitor (whose capacity is equal to the cross-section of the surge tank), head loss
by a non-linear resistance K jQj, water inertia by an inductance (whose value is
equal to L/Sg). Following demonstration (3.5)–(3.8), the turbine is represented like
[1,4] by a variable voltage source H(u,W) controlled by the gate opening and
influenced (in small proportion) by the rotational speed. Since the gate opening, u
can vary from 0% to 100%, the voltage source can vary, respectively, from
Hmax ¼ Hb to Hmin ¼ Hb  KQ2max leading to a discharge variation from 0 to Qmax .
This macroscopic turbine model integrates the environment of the turbine, i.e. its
water column with the conduit system. From a physical point of view, we recover
that the discharge is well generated both by a head (potential difference) and by a
gate opening.
Therefore, it comes the equivalent electric circuit drawn in Figure 3.8 which is
convenient to model any kind of hydraulic circuits. It enables to use all electric’s
theorem (Thevenin, Norton and Millman) to easily simplify any complex HPP
(multi-reservoir, multi-penstock and multi-tank) For instance, a multi-penstock
HPP can be modeled by a single equivalent penstock thanks to electric’s theorem.
Another way for obtaining a hydraulic model of a HPP consists in writing
Kirchhoff’s laws from the equivalent electric circuit and in replacing electric
variables-parameters set by hydraulic variables-parameters set.
Reduced order models for grid connected hydropower plants 57

Qa Qc
A B D
Q

C
Hb H(u,Ω)
Z

Figure 3.8 Generic electric equivalent circuit

By neglecting the surge tank head losses and by considering an equal flow
distribution among turbines, thanks to the equivalent electric circuit (Figure 3.8),
we can write unsteady flow equations (3.9) related to inelastic model:
8
> dQa g  Sa g  Sa
> dt ¼ L ðHb  Z Þ  Ka L jQa jQa
>
>
>
> a a
>
< dQ g  Sc g  Sc
¼ ðZ  H Þ  Kc n jQjQ (3.9)
>
> dt n  Lc Lc
>
>
>
>
>
: dZ ¼ 1 ðQa  n  QÞ
dt St
with:
● Qa the tunnel flow, in m3/s
● Qc the penstock flow, in m3/s
● Q the turbine discharge, in m3/s (Q ¼ Qc/n)
● Z the surge tank water level, in m
● n the number of generating units in the plant
● Sa, Sc and St the cross-section of the tunnel, penstock and surge tank, in m2
● La and Lc the length of the tunnel and the penstock, in m
The turbine discharge Q depends on the net head H, the gate opening or guide
vane opening u (control input) and in a least amount on the rotational speed W:
pffiffiffiffi
Q ¼ Gðu; WÞ H (3.10)

where G(u, W) is the gate opening function2 which can be obtained from the
turbine characteristic curve Q(H, u, W). It is to notice that for Pelton turbines, this
function does not depend on the rotational speed W since the wheel is not
submerged.

2
A second-order polynomial is often sufficient to approximate G(u).
58 Modeling and dynamic behaviour of hydropower plants

3.2.4.1 Non-linear model in per unit


In order to have a per unit model based on the operating point (Q0, H0, W0) which is
the initial equilibrium point, we set:
● q ¼ QQ0 ; qa ¼ Q
Q0 ; pm ¼ Pm0 ; z ¼ H0 ; hb ¼ H0
a Pm Z Hb

● w ¼ ðW=W0 Þ in per unit with the initial speed W0 often be to the rated speed Wn
● Twa ¼ ðLa  Q0 =g  Sa  H0 Þ the tunnel water start time, in s
● Twc ¼ nðLc  Q0 =g  Sc  H0 Þ the penstock water start time, in s
● Tz ¼ ðSz  H0 Þ=Q0 the surge tank drain time, in s
● la ¼ Ka ðg  Sa  Q0 =La Þ coefficient taking into account the tunnel head
losses, in s1
● lc ¼ n  Kc ðg  Sc  Q0 =Lc Þ coefficient taking into account the penstock head
losses, in s1
With these notations, equations in p.u. are:

8
> dqa 1
>
> ¼ ðhb  zÞ  la jqa jqa
>
> dt T
>
>
wa
< dq 1
¼ ðz  hÞ  lc jqjq (3.11)
>
> dt Twc
>
>
>
>
>
:dz ¼ 1 ðqa  n  qÞ
dt Tz

In steady state, note that


pffiffiffiffiffiwith
ffi our p.u. base, the tunnel flow is qa0 ¼ n. Moreover,
since Q0 ¼ Gðu0 ; W0 Þ H0 , we can write with the auxiliary variable v ¼ Gðu; WÞ=
Gðu0 ; W0 Þ and the turbine efficiency h0 ¼ Pm0 =Ph0 associated with the considered
operating point (Q0, H0, W0):

pffiffiffi q2
q¼v h or h ¼ (3.11a)
v

3.2.4.2 Linearized model


As we are interested in small changes around the operating point, previous equa-
tions can be linearized to yield:
1
Dq ¼ Dv þ Dh; Dv ¼ a0 Du þ s0 Dw
2
and

Dh ¼ 2ðDq  a0 Du  s0 DwÞ (3.12)

with Du ¼ u  u0 ; Dq ¼ q  1; Dpm ¼ pm  1; Dv ¼ v  1
Reduced order models for grid connected hydropower plants 59

and
   
1 @G W0 @G
a0 ¼ ; s0 ¼
Gðu0 ; W0 Þ @u u0 Gðu0 ; W0 Þ @W W0

Therefore, the linearized hydraulic model is:


8
> dDqa 1
>
> dt ¼ T Dz  2nla Dqa
>
>
>
>
wa
>
< dDq  
1 1 2s0 2a0
¼ 2 þ lc Dq þ Dz þ Dw þ Du (3.13)
>
> dt Twc Twc Twc Twc
>
>
>
>
>
> dDz 1
: ¼ ðDqa  nDqÞ
dt Tz

The model (3.13) can be used for controller design like in [8] or to set up perfor-
mance indicators in order to evaluate the dynamic capability of a HPP as detailed
in [9,10].

3.2.5 Mechanical model of the generating unit


The rotation dynamics of the synchronous generator shaft is given by the motion
equation, the so-called swing equation obtained applying the Work-Energy
theorem [11–13] and by neglecting the friction on the shaft:

dE
¼ Pm  Pe (3.14)
dt

where E ¼ ð1=2ÞJ W2 is the kinetic energy (in J), J is the combined moment of
inertia of the rotating masses associated with the rotor and the turbine (kg/m2), W is
the rotor angular velocity (rad/s) and Pe is the electrical power (W) demanded by
the power system (see Section 3.3).
Notice that the equilibrium point is defined by Pm0 ¼ Pe0 ¼ P0 and that the
rotational speed W is proportional (number of pair of poles) with the power system
frequency (F).
The mechanical power provided by a turbine can be written as [14] with the
hydraulic power Ph ¼ rgQH and mechanical losses LðQÞ  W3 which depends on
the rotational speed W:

Pm ¼ Ph  LðQÞ  W3 (3.15)

with r the water density (kg/m3) and L(Q) a function which can usually be
approximated by a second-order polynomial.
60 Modeling and dynamic behaviour of hydropower plants

3.2.5.1 Non-linear model in per unit


Writing (3.9) in per unit notation yields:

de
Tm ¼ pm  pe (3.16)
dt
with:
● e the kinetic energy in per unit such that e ¼ w2
qh W30
pm ¼  LðQÞw3 (3.17)
h0 P0

● Tm the mechanical time constant3 (in s) related to the inertia of the whole
turbine þ generator and defined by Tm ¼ ðE0 =P0 Þ with the steady state kinetic
energy E0 ¼ ð1=2ÞJ W20 .
It is interesting to emphasize that most of the electrical engineering literature
[11,12] simplifies the non-linear equation (3. 13) by writing: de/dt ¼ 2w(dw/dt) 
2(dw/dt). This approximation is acceptable since even in severe conditions, varia-
tions of w are lower than 4%. Moreover, we can notice that the linearization of
(3.13) gives the same equation (3.17) than the one obtained by the literature sim-
plification [11,12].

3.2.5.2 Linearized model


The linearization of (3.17) and (3.18) gives, respectively:
first,

dDw
2Tm ¼ Dpm  Dpe (3.18)
dt
second,

Dpm ¼ h1 1
0  b0 Dq þ h0 Dh  g0 Dw

Hence, by using (3.14), the latter equation becomes:

Dpm ¼ c1 Dq þ c2 Dw þ d1 Du (3.19)
  
Q0 W0 3
with c1 ¼ 3h1 1 1
0  b0 ; c2 ¼  2s0 h0 þ g0 ; d1 ¼ 2a0 h0 ; b0 ¼ P0
@L
@Q Q and
0
3W30
g0 ¼ P0 LðQ0 Þ
It is important to emphasise that the influence of the speed w on hydraulics
variables (Pm, Q, H and h) is low due to the fact that coefficients ðs0 ; c2 Þ are
usually very small compared to other coefficients. That is why speed is not

3
In the literature [12], Tm is also called the mechanical starting time of the generating unit.
Reduced order models for grid connected hydropower plants 61

taken into account in hydraulic models applied to interconnected power systems


(see Section 3.4.2). Note also the negative sign of the coefficient d1 which illustrates
the non-minimum phase behaviour of a HPP.

3.2.6 Hydro-mechanical model of the power plant


Finally, we can gather the hydraulic model (3.13) with the mechanical model (3.19)
to yield a linear hydro-mechanical state model (3.20) of the HPP with a state vector
x ¼ ½ qa q z :
2 3 2 3
a1 0 a2 0 0
d Du
Dx ¼ 4 0 a3 a4 5Dx þ 4 b1 b2 5
dt Dw
a5 a6 0 0 0
(3.20)
Du
Dpm ¼ ½ 0 c1 0  Dx þ ½ d1 c2 
Dw

1
 1 1
with a1 ¼ 2nla ; a2 ¼ Twa ; a3 ¼ 2 Twc þ lc ; a4 ¼ Twc ; a5 ¼ Tz1 ;

a6 ¼  Tz ; b1 ¼ 2a0 =Twc ; b2 ¼ 2s0 =Twc ; c1 ¼ 3h0  b0 ; c2 ¼  2s0 h1
n 1
0 þ g0 ;
1
and d1 ¼ 2a0 h0 .
This model will be used in Section 3.4 to construct the complete state model of
a HPP connected to a power system.
Similarly, it is possible to construct a non-linear model of a HPP by gathering
(3.11) and (3.17).

3.3 Synchronous power system models


Power system consists of producers with different energy sources (thermal, nuclear,
renewable) and consumers with different load nature (AC or DC motors, heating,
lightning, power electronics, etc.). The stable operation is ensured when there is
equilibrium between production and consumption thanks to frequency control offer
by dedicated power plants. As the power system frequency is common to the whole
grid, a load variation at one point will be reflected on the grid as a frequency
variation. The relation between load and frequency variations is developed in this
section using transfer function representation of the system.
A power system is an infinite dimensional non-linear system which remains
complex to model with a huge number of uncertain parameters (synchronous
machine parameters, controllers’ parameters and line impedances) and a variable
structure (load changes, outage of transmission lines or generating units). To
manage this complexity, three simplified models are developed; the first one is a
general model dedicated for any type of synchronous multi-machine system and not
met in literature (Section 3.3.1); the second presents the particular case of a large
interconnected power system (Section 3.3.2); and the third gives an extreme
operating mode in separated network (Section 3.3.3).
62 Modeling and dynamic behaviour of hydropower plants

nG G1 G2 Gns

npe pe1 pens

pL

Load

Figure 3.9 Simplified multi-machine system

3.3.1 General model


A general model is described in this section to hold for any type of synchronous
power system. It is based on a simplified multi-machine system shown in Figure 3.9
where the HPP consists of n hydraulic machines G connected to any type of syn-
chronous power system characterized by ns generating units G1  Gns and a variable
load [1]. We assume a synchronous operation, i.e. a perfect synchronous rotation
among the ns þ n generators (electromechanical oscillations neglected between
generating units), and we consider an invariant voltage in the electrical network.
The aim of the proposed simplified model is to be representative of the power-
frequency dynamic behaviour of any kind of synchronous system featured with any
size and any number of machines connected to a load. This concerns the following
different power grids:
● Strongly or weakly interconnected system (wide or small synchronous area)
● Isolated grid
● Black start4 operation
For frequency control studies applied to a small power system where the grid is
weakly interconnected (few machine), the following general model, not met in
literature, is well suited because it takes into account not only the inertia of the
machines but also the regulation time constant of the governing systems.
The modeling procedure consists in writing mechanical equations of the multi-
machine system, in defining the mechanical and electrical powers and then to
deduce transfer functions related to frequency variations.
Normalising with the base (P0, Q0, H0, W0) associated with the hydraulic unit
and using the synchronous operating hypothesis w ¼ W=W0 ¼ Wi =W0i ¼ F=F0 ¼ f
each generator i in the power system is described by the swing equation:

dw2
2Tmi ¼ pmi  pei with i  ½1; ns  (3.20a)
dt

4
Black start is a service proposed by some HPP to restore a part of an electric grid to operation without
relying on the external transmission network.
Reduced order models for grid connected hydropower plants 63

or its linearized form:


dDw
2Tmi ¼ Dpmi  Dpei with i  ½1; ns  (3.21)
dt
where Tmi is the mechanical time constant (in s) of the generating unit i defined by
Tmi ¼ ðE0i =P0 Þ with E0i ¼ ð1=2ÞJ W20i
Concerning the load, as the voltage is assumed to be constant, the load depends
on two terms: a resistive term Dp and a frequency-sensitive term [11]:

pL ¼ p þ mDðw  w0 Þ or DpL ¼ Dp þ mDDw (3.22)

where m ¼ ðPLn =P0 Þ is a normalisation factor. The parameter D denotes the self-
regulation of the load in the synchronous area and is typically defined with respect
to PLn the rated power of the load: D ¼ ððDPL =PLn Þ=DwÞ. It is usually assumed [7]
to be 1%/Hz; that means a load decrease of 1% occurs for a frequency drop of 1 Hz.
So, for a 50 Hz power system frequency, it gives D ¼ 0.5.
From Figure 3.9 the Kirchhoff’s law yields:
X
n
DpL ¼ nDpe þ Dpei (3.23)
i¼1

By adding n equations (3.19) with ns equations (3.21) and by injecting (3.22) and
(3.23), we find:
dDw
2ðnTm þ Tms Þ þ mDDw ¼ nDpm þ Dpms  Dp
dt
with:

X
ns X
ns
pms ¼ pmi ; Tms ¼ Tmi (3.24)
i¼1 i¼1

Let’s suppose that the ns turbines participate to primary frequency control


thanks to a governing system. In these conditions, the mechanical power pmi gen-
erated by each turbine i is affected by the governor droop defined in steady state by
Ri ¼ ðDw=ðDPmi =Pni ÞÞ1 with Pni the rated power of the generating unit i [15].
Assuming that the power-frequency dynamic response of each machine is
characterized by an individual time constant ts (governing system dynamics), we
can express the mechanical power due to the grid with respect to the base P0
through the normalisation factor li ¼ ðPni =P0 Þ:

dDpmi li
ts þ Dpmi ¼  Dw with i  ½1; ns 
dt Ri
(3.25)
dDpms X
ns
li
) ts þ Dpms ¼ Ks Dw with Ks ¼
dt i¼1
Ri
64 Modeling and dynamic behaviour of hydropower plants

∆p


+ 1 + τ ss
n∆pm ∆ω
δ2s2 + δ1s + δ0

Figure 3.10 Block diagram of a synchronous power system

Thus, the power grid can be seen as a single generating unit with a large mechanical
time constant Tms (sum of Tmi) and a small regulation constant ts (linked to one
generating unit).
Equations (3.24) and (3.25) give a linear state model of a synchronous power
system with two states ðDw; Dpms Þ, one input ðnDpm  DpÞ and one output Dw.
By using the Laplace operator, a transfer function model can also be given:
1 þ ts s
DwðsÞ ¼ ðnDpm ðsÞ  DpðsÞÞ
d2 s2 þ d1 s þ d0
with
d2 ¼ 2ðnTm þ Tms Þts ; d1 ¼ 2ðnTm þ Tms Þ þ mDts ; d0 ¼ Ks þ mD (3.26)

The model (3.26) illustrated by Figure 3.10 is a general model suitable for any
kind of network with any number of machines supplying an aggregated load. For
instance, it can be used for sensitivity studies regarding the frequency dynamics or
for primary frequency control design in a small power grid consisting of a few
number of generating units [1,8]. This general model holds in synchronous opera-
tion, so for an electrical network split in distinct synchronous zones (like for multi-
zone secondary frequency control studies or for interconnection studies), the model
could be used for each zone and completed by a tie line model giving the power
exchange between zones.
Furthermore, lower order models can be found in literature [7,12] where in the
regulation time constant of the grid (ts ) is not taken into account leading to a less
accurate and less representative model (see details in Section 3.3.2).

3.3.2 Model for an interconnected grid


The general model described in Section 3.3.1 can be reduced if we consider a large
interconnected power system also called wide area synchronous grid, i.e. a bulk
electrical network with a large number of power plants. For this case, the power
provided by the HPP is very small compared with the power generated by the grid.
It means that npm can be neglected compared with pms in (3.24). In addition, as the
number ns is very large (ns  100), we have Tms  nT m . In these conditions, it
yields a second-order filter between load and frequency variations:

DpðsÞ
Df ðsÞ ¼ DwðsÞ  (3.27)
2Tms ts s2 þ ð2Tms þ mDts Þs þ Ks þ mD
Reduced order models for grid connected hydropower plants 65

60

40

20
DF (mHz)

–20

–40

–60
0 0.5 1 1.5 2 2.5 3 3.5 4
Time (h)

Figure 3.11 Power frequency variations measured in France around


50 Hz

Equation (3.27) gives numerous information concerning the behaviour of the fre-
quency in a large synchronous interconnected system:

● In first approximation, it is admitted that power system frequency does not


depend on the power generation of the studied HPP so that it is allowed to
consider the frequency as an independent input for power-frequency control
studies in interconnected operation.
● Frequency variations (see Figure 3.11) are the image of load variations passing
through a filter.
● The transient response of the frequency is influenced by parameters
Tms ; ts ; m; D which determine the dynamics of the signal in terms of settling
time, overshoot, damping of oscillations.
● A frequency residual static deviation exists with the primary frequency control
and is equal to Df1 ¼ Dp1 =ðKs þ mDÞ. That’s why the secondary frequency
control is needed to cancel this static deviation.

The model (3.27) was validated on several scenarios of grid incidents [1].
Figure 3.12 presents a validation for a 1,332 MW production loss in the French
grid. Note that this model gives a good match between measurement and simula-
tions. Common simpler models [7,12] neglecting the time constant, ts are not able
to reproduce the overshoot appearing in Figure 3.12 because they are based on a
first-order filter.
66 Modeling and dynamic behaviour of hydropower plants

49.995
Simulation
49.99 Real data

49.985

49.98
Frequency (Hz)

49.975

49.97

49.965

49.96

49.955

49.95

49.945
15 20 25 30 35 40 45 50 55 60
Time (s)

Figure 3.12 Comparison between frequency measurement (solid curve) and


simulation (dashed curve) for a 1,332 MW production loss in the
French grid

3.3.3 Model for an isolated grid


In contrast to an interconnected grid, an isolated system is defined as a very small
grid with few or without power plants and a consumption area. The isolated
operation mode occurs when there is some failure in tie lines. Two kinds of isolated
grid can be distinguished:
1. A small synchronous area consisting of few number of generating units and
some loads. This is the typical case of an island or an area disconnected from
the main power supply zone following a blackout.
2. A passive network without any producer except the HPP connected to a load.
This situation is met for instance after a black out in a black start operation (as
defined in Section 3.3.1) when the HPP supplies its auxiliaries and then con-
tributes to system restoration by line charging.
Case (1) can be solved by using the general model (3.26) developed in
Section 3.3.1.
We will focus on the extreme situation (2) where the passive isolated grid is
featured with ns ¼ 0 and n hydraulic units connected to a sensitive frequency load.
So, by setting pms ¼ 0, the general model (3.26) is reduced to:

nDpm ðsÞ  DpðsÞ


Df ðsÞ ¼ DwðsÞ  (3.28)
2nTm  s þ mD
Reduced order models for grid connected hydropower plants 67

For this severe operating condition, large variations of frequency are expected and
necessitate to take into account the speed in characteristic curves of the hydraulic
turbine. We will notice that the worst case is met for a null self-regulation of the
load (D ¼ 0).

3.4 Complete state-space model for a hydro plant


connected to a grid
The aim of this section is to develop a linear state-space model of a hydropower
plant connected to a power grid. A distinction will be done as defined in Section 3.3
between a large interconnected grid and an isolated grid.

3.4.1 General model


The proposed general model is based on Figure 3.13 giving a block diagram of a
HPP connected to a synchronous power system. This is a functional representation
of the whole multi-variable system which is split in two interconnected blocks: a
hydro-mechanical subsystem containing the state model (3.20) of the studied hydro
generating unit and an electro-mechanical subsystem related to the following state
model (3.29) of the whole generators (n þ ns).
3.4.1.1 Hydro-mechanical subsystem
This subsystem is described by (3.20) given in Section 3.2.5. Its input Dw is provided
by the electro-mechanical subsystem and its output Dpm supplied this latter block.
3.4.1.2 Electro-mechanical subsystem
Regrouping (3.24) and (3.25), the linear state model linked to the electro-
mechanical behaviour of the n þ ns generators is given by (3.29) with two states
ðDw; Dpms Þ, two inputs ðDpm ; DpÞ and two outputs ðDw; Dpe Þ which are useful for
the governing system of the HPP.
d Dw r1 a10 Dw na10 a10 Dpm
¼ þ
dt Dpms a9 a11 Dpms 0 0 Dp
(3.29)
Dw 1 0 Dw 0 0 Dpm
¼ þ
Dpe c3 c4 Dpms d2 d3 Dp

∆p ∆pe
Electro-
Hydro- mechanical
∆u mechanical subsystem
∆ω
subsystem ∆pm

Figure 3.13 Block diagram of a HPP connected to a synchronous power system


and perturbed by a load variation D
68 Modeling and dynamic behaviour of hydropower plants

with
mD 1 Ks 1
r1 ¼  ; a10 ¼ ; a9 ¼  ; a11 ¼ 
2ðnT m þ Tms Þ 2ðnT m þ Tms Þ ts ts
c3 ¼ 2r1 Tm ; c4 ¼ 2a10 Tm ; d2 ¼ 1  2na10 Tm ; d3 ¼ 2a10 Tm
(3.29a)
A transfer function form can be derived as well. The dynamics of the speed/
frequency was already presented by (3.26). On the other hand, we can extract the
electric power provided by the hydraulic generating unit from (3.18) yielding
Dpe ¼ Dpm  2Tm ðdDw=dtÞ, and by substituting (3.26), we obtain:

ðd2  2nTm ts Þs2 þ ðd1  2nTm Þs þ d0 2Tm ð1 þ ts sÞs


Dpe ðsÞ ¼ Dpm ðsÞ þ DpðsÞ
d2 s2 þ d1 s þ d0 d2 s2 þ d1 s þ d0
(3.30)
Equation (3.30) shows clearly a fast dynamics to respond quickly to the demand
Dp (derivative action Tm  s) and a slower action linked to the mechanical dynamics
of Dpm.
3.4.1.3 Concatenation
The two state models (3.30) can be gathered to yield a fifth-order MIMO5 system
with a larger state vector x ¼ ½ qa q z wpms :
2 3 2 3
a1 0 a2 0 0 0 0
6 0 a3 a4 b2 0 7 6 b1 0 7
d 6 7 6 7 Du
Dx ¼ 6 6 a a 0 0 0 7
7 Dx þ 6
6 0 0 7
7
4 b3 a10 5 Dp
5 6
dt 4 0 a7 0 a8 a10 5
0 0 0 a11 a12 0 0

Dw 0 0 0 1 0 0 0 Du
¼ Dx þ (3.31)
Dpe 0 c1 d2 0 c5 c4 d1 d2 d3 Dp

with coefficients which depend on the equilibrium point (P0, Q0, H0, W0) such as:
1
 1 1
a1 ¼ 2nla ; a2 ¼ Twa ; a3 ¼ 2 Twc þ lc ; a4 ¼ Twc ; a5 ¼ Tz1 ; a6 ¼ n=Tz ;
Ks 1 1
a7 ¼ nc1 a10 ; a8 ¼ r1 þ nc2 a10 ; a9 ¼  ; a10 ¼ ; a11 ¼  ;
ts 2ðnT m þ Tms Þ ts
b1 ¼ 2a0 =Twc ; b2 ¼ 2s0 =Twc ; b3 ¼ nd1 a10 ; c1 ¼ 3h1 0  b0 ; c4 ¼ 2a10 Tm ;
c5 ¼ c2 d2 þ c3 ; d1 ¼ 2a0 h1
0 ; d2 ¼ 1  2na 10 Tm and d3 ¼ 2a10 Tm
(3.31a)

3.4.2 Interconnected operation


In normal operating conditions, the frequency never change more than 2%, so it is
allowable to neglect its effect on the hydraulic model of the HPP (Section 3.2.4).

5
Multiple Input Multiple Output.
Reduced order models for grid connected hydropower plants 69

Thus, we will consider in this section that the discharge Q and the mechanical
power Pm are only function of H and u.
Now, if we observe the transfer function (3.30), a remarkable asymptotic
behaviour can be seen. For this, let’s tend ns to infinity in order to represent a large
interconnected power system. In this condition, only d0 ; d1 and d2 tend to infinity.
As a consequence, we have:

lim Dpe ðsÞ ¼ Dpm ðsÞ (3.32)


ns !1

That means in interconnected operation, the dynamics of electric and mechanical


power of the hydro unit can be matched (not true in separated operation
Section 3.4.3). Moreover, as demonstrated in Section 3.3.2, the frequency ( f ¼ w)
can be considered as an independent variable which does not depend on the HPP
and whose dynamics is given by (3.28).
Therefore, the general model (3.31) can be simplified by a third-order state
model with x ¼ ½ qa q z  and a single input Du:
2 3 2 3
a1 0 a2 0
d
Dx ¼ 4 0 a3 a4 5Dx þ 4 b1 5Du
dt (3.33)
a5 a6 0 0

Dpe ¼ Dpm ¼ ½ 0 c1 0 Dx þ d1 Du

3.4.3 Isolated operation


With the same condition as Section 3.3.3 of a passive isolated grid (ns ¼ 0, isolated
load), the dynamics of the frequency is directly influenced by the HPP and can no
longer be neglected in the hydraulic model. Therefore, by injecting pms ¼ Tms ¼ 0
in the general model (3.31), we obtain for x ¼ ½ qa q z w :
2 3 2 3
a1 0 a2 0 0 0
60 b2 7 6 b1 0 7
d
Dx ¼ 6
a3 a4 7Dx þ 6 7 Du
dt 4 a5 a6 0 05 40 0 5 Dp
0 a7 0 a8 b3 a10
Dw 0 0 0 1 0 0 Du
¼ Dx þ (3.34)
Dpe 0 c1 d2 0 c5 d1 d2 d3 Dp

with the same parameters as in (3.31) except for a10 ¼ ð1=2nT m Þ and
r1 ¼ ðmD=2nT m Þ.

3.5 Analysis of the dynamic behaviour


This section concerns in one hand the dynamic decoupling and on other, the
dynamic performance limitations of a HPP operating in interconnected mode.
70 Modeling and dynamic behaviour of hydropower plants

Bode diagram
–20
Magnitude (dB)

–30

–40

–50
720
Phase (deg)

360

0
10–2 10–1 100 101 102
Frequency (rad/s)

Step response
0.06

0.04
Amplitude

0.02

Low head/high flow conditions


–0.02
High head/low flow conditions
–0.04
0 20 40 60 80 100 120 140 160 180 200
Time (s)

Figure 3.14 Visualisation of slow and fast dynamic (open-loop time and
frequency responses)

3.5.1 Decomposition of slow and fast dynamics


As shown in Figure 3.14, the time domain and frequency responses of the hydraulic
model (open loop) can be split up into two kinds of dynamics:
● a quick behaviour, linked to the penstock flow inertia (water start time)
● slower oscillating dynamics, related to mass water hammer phenomenon
between the tunnel and the surge tank
Such behaviours can also be observed in the field. Figure 3.15 shows a real
step power response applied to a 12 MW generating unit (Francis turbine) in closed
loop. The fast transient dynamics is framed with a vertical rectangle while the slow
dynamics is framed with a horizontal rectangle.
According to previous statements, it seems possible to write the transfer
function between the actuator position and the unit active power as the product of
two sub-systems:

Dpm ðsÞ
GðsÞ ¼ ¼ G1 ðsÞG2 ðsÞ (3.35)
DuðsÞ
Reduced order models for grid connected hydropower plants 71

10.5

10
Power (MW)

9.5

9
0 ~ tm 100 200 300 400 500 600 700 800 900
Time (s)

Figure 3.15 Visualisation of slow and fast dynamics (closed-loop field tests)

with G1 ðsÞ the fast first-order-like dynamics and G2 ðsÞ the slow second-order-like
dynamics.
In the previous state model (3.33), as we have Twc < Twa and Twc < Tz , the fast
state is Dq, and the slow states are Dqa and Dz. We can then rewrite the system,
highlighting these different dynamics:
   
d Dx1 A11 A12 Dx1 B1
¼ þ Du (3.36)
dt Dx 2 A 21 A 22 Dx 2 B 2
 
Dx1
Dpm ¼ ½ C1 C2  þ DDu
Dx2
qa
with x1 ¼ q; x2 ¼ and:
zce

0 a a2
A11 ¼ a3 ; A12 ¼ ½ 0 a4 ; A21 ¼ ; A22 ¼ 1
a6 a5 0
B 1 ¼ b 1 ; B2 ¼ ½ 0 0 T ; C1 ¼ c; C2 ¼ ½ 0 0 ; D ¼ d1 (3.37)
Based on this representation, our goal here is to decompose GðsÞ in two sim-
plified sub-systems:
● the faster one, G1 ðsÞ, defined such as the step responses of GðsÞ and G1 ðsÞ are
equivalent for t < tm
● the slower one, G2 ðsÞ, defined such as the step responses of GðsÞ and G2 ðsÞ are
equivalent for t > tm
with tm the rising time of the power response.
The transfer functions can be calculated by using the singular perturbations
method [16] (model reduction method based on quasi-static approximation).
72 Modeling and dynamic behaviour of hydropower plants

3.5.1.1 Fast dynamics


For t < tm , since Dx2 is featured with a slow dynamic, we can consider that
ðd=dtÞDx2 ¼ 0. So, from (3.36), it is obtained:

0 ¼ A21 Dx1 þ A22 Dx2 þ B2 Du ) Dx2 ¼ A1


22 A21 Dx1 (3.38)
leading to the state-space representation of G1 ðsÞ:
d 
Dx1 ¼ A11  A12 A22 1 A21 Dx1 þ ðB1 þ 0ÞDu (3.39)
dt

Dpm ¼ C1  C2 A22 2 A21 Dx1 þ ðD  0ÞDu (3.40)

or
   
d 2  2a0
Dq ¼  1 þ n2 Twa la  2lc Dq þ Du (3.41)
dt Twc Twc
   
3 a0
Dpm ¼  b0 Dq  2 Du (3.42)
h0 h0

By neglecting the head losses as in [14], (3.41) is simplified:


 
d 2 2a0
Dq ¼  Dq þ Du (3.43)
dt Twc Twc
Then, the transfer function G1 ðsÞ is equivalent to the classical model IEEE [14]
used for turbine speed control studies:

a0 h1
0 ð1  b0 h0  Twc sÞ 1  Twc s
G1 ðsÞ ¼  K1 (3.44)
1 þ ðTwc =2Þs 1 þ ðTwc =2Þs
with K1 ¼ a0 h10
Therefore, the fast dynamics of hydro unit power response are mainly related
to penstock and turbine characteristics. Let’s also notice the presence of a negative
zero: G1 ðsÞ is known as a ‘non minimum phase’ system.

3.5.1.2 Slow dynamics


For t > tm , the fast dynamics Dx1 vanishes, so we can write ðd=dtÞDx1 ¼ 0. Hence,
from (3.36), we have:

0 ¼ A11 Dx1 þ A12 Dx2 þ B1 Du (3.45)


leading to the state-space representation of G2 ðsÞ:
d  
Dx2 ¼ A22  A22 A11 1 A12 Dx2 þ B2  A21 A11 1 B1 Du (3.46)
dt
 
Dpm ¼ C2  C1 A11 1 A12 Dx2 þ D  C1 A11 1 B1 Du (3.47)
Reduced order models for grid connected hydropower plants 73

or
d 1
Dqa ¼ 2n  la  Dqa  Dz (3.48)
dt Twa
d 1 n n  a0
Dz ¼ Dqa  Dzce  Du (3.49)
dt Tz 2Tz ð1 þ Twc  lc Þ Tz ð1 þ Twc  lc Þ
    
3h1
0  b0 a0 a0 3h1 0  b0
Dpm ¼ Dz þ 2  Du (3.50)
2ð1 þ Twc  lc Þ h0 1 þ Twc  lc

Neglecting the head losses, (3.48)–(3.50) are simplified:


d 1
Dqa ¼  Dz (3.51)
dt Twa
d 1 n n  a0
Dz ¼ Dqa  Dz  Du (3.52)
dt Tz 2Tz Tz
 1   
3h0  b0 a0 
Dpm ¼ Dz þ 2  a0 3h1  b Du (3.53)
2 h0 0 0

Then, we can obtain the transfer function G2 ðsÞ


1 þ 2ðx1 =w1 Þs þ ð1=w1 2 Þs2
G2 ðsÞ ¼ K2
1 þ 2ðx2 =w2 Þs þ ð1=w2 2 Þs2

With K2 the static gain of G2 ðsÞ and


rffiffiffiffiffiffiffi rffiffiffiffiffiffiffi
1 n n Twa n n Twa
w1 ¼ w2
2 2
¼ ; x1 ¼ w1 Twa ¼  ; x2 ¼ w2 Twa ¼ ;
Tz Twa 2 2 Tz 4 4 Tz
(3.54)
the period T matching with the angular frequency w1 ¼ w2 is:
sffiffiffiffiffiffiffiffiffi
pffiffiffiffiffiffiffiffiffiffiffi La Sz
T ¼ 2p Tz Twa ¼ 2p (3.55)
gSa

known as the natural period of the tunnel flow and the surge tank water level
oscillations, that is invariant regardless of the operating point, as Q0 and H0 do not
appear in (3.55). On the contrary, the damping of the couple of poles and zeros of
G2 ðsÞ depends on the considered operating point on flow and head.
If we now take into account head losses in (3.48)–(3.50), it is possible to refine
the expression of w1 ; w2 ; x1 and x2 :

1  2ðn2  T wa  la þ Twc lc Þ


w1 2 ¼ (3.56)
Tz  Twa ð1  2  Twc  lc Þ
74 Modeling and dynamic behaviour of hydropower plants

2n  Tz  Twa  la ð1  2  Twc  lc Þ  n  Twa


x1 ¼ w1 (3.57)
2ð1  2ðn2  Twa  la þ Twc  lc ÞÞ
1 þ n2  Twa  la þ Twc  lc
w2 2 ¼ (3.58)
Tz  Twa ð1 þ Twc  lc Þ
4n  Tz  Twa  la ð1 þ Twc  lc Þ þ n  Twa
x2 ¼ w2 (3.59)
4ð1 þ n2  Twa  la þ Twc  lc Þ
We can notice that x2 > 0, whereas x1 can be positive or negative following the
cross-section of the surge tank and the amount of head losses. If x1 < 0, G2 become
non-minimum phase and oscillations are expected limiting control performance.
This will lead us to develop performance limitation criteria Section 3.5.2.
The present approach was tested on the real case of ‘Malgovert’ hydropower
plant (a 4  80-MW HPP in the French Alps equipped with Pelton turbines).
Figure 3.16 illustrates the results of singular perturbations method to find the fast
and slow subsystems. The two distinct dynamics are well reproduced: the sub-
systems and the complete system are matched.

3.5.2 Performance limitation for primary frequency


control: capability criteria
Hydropower plants are well known for their great responsiveness (more than
thermal power plants), which make them useful for the electricity grid, particularly
in order to reach a perfect balance between production and consumption (load–
frequency control). To increase the supply of such kind of services with hydro
units, the so-called ancillary services, it is necessary to evaluate the performance
limitations of a HPP to offer primary frequency control. The issues are to specify
realistic performance requirements to governing system manufacturers, to quantify

Step response Step response


1 0.99

0.9 Fast subsystem G1(s) Slow subsystem G2(s)


0.985 Complete system G(s)
Complete system G(s)
0.8
0.98
0.7

0.6 0.975
Amplitude

Amplitude

0.5 0.97

0.4
0.965
0.3
0.96
0.2

0.1 0.955

0 0.95
0 1 2 3 4 5 0 200 400 600 800 1000
Time (s) Time (s)

Figure 3.16 Validation of the dynamics decoupling for the linear model
(4  80 MW Malgovert HPP)
Reduced order models for grid connected hydropower plants 75

the number of HPP which can be contracted with the transmission system operator
and to give technical arguments for investors (for maintenance of existing units, or
development of new ones). Computer simulation could be good tools for that, but
for producers with a huge number of units, it is unrealistic and too much time
consuming to make a bunch of simulations with non-linear model for each one.
EDF has answered to these issues by developing capability criteria related to
primary frequency control performance [9,10]. The ability for power and speed
control are obviously linked to the unit itself (including it servo-positioner limita-
tion with its slew rate) and to the hydraulic circuit featured with a non-minimum
phase which can be problematic for controller design, as the unstable zero can
cause the appearance of an unstable pole in closed-loop:
● G1 ðsÞ has an unstable first-order zero, which is inherent to the process
● G2 ðsÞ can be non-minimum phasis, if x1 < 0: see (3.57)
Thanks to the transfer functions G1, G2 calculated in Section 3.5.1, we develop
hereafter two limitation criteria due to the presence of a non-minimum phase
affecting G1, G2 and valid for any control loop [13].

3.5.2.1 Penstock water start time criterion


The fast dynamics during a power step response described in G1 ðsÞ must be suffi-
ciently fast to respect the rising time tm . The rapidity performance of the control
loop is limited by the non-minimum phase in G1 whatever the controller tuning is.
This will lead to our first limitation criterion.
Here, we will use a result from De Larminat [13] about a stability/rapidity
compromise for non-minimum phase system (as G1 ).
Consider LðsÞ ¼ K ðsÞGðsÞ ¼ K ðsÞGnmp ðsÞGmp ðsÞ a controlled open-loop, i.e.
the so-called Loop Transfer, which can be written as the product of a minimum
phase and a non-minimum phase system:

LðsÞ ¼ Lmp ðsÞLnmp ðsÞ (3.60)


with

Lnmp ðsÞ ¼ Gnmp ðsÞ and Lmp ðsÞ ¼ Gmp ðsÞ  K ðsÞ (3.61)

We also define the minimum phase and non-minimum phase subsystems of G as


 ðsÞ ¼ G
G  mp ðsÞGnmp ðs Þ. It is possible to find theses subsystems with the property:
Lnmp ðsÞ ¼ Gnmp ðsÞ ¼ 1.
Bode established in 1945 a relation [13] between the phase j and the slope S of
jF ðjwÞj
the magnitude jF ðjwÞj so that S ¼ dlndlnw for w > 0 and a transfer function F( jw)
which is minimum phase i.e. with poles and zeros on the left plan of Nyquist
diagram and without any delay:

j ¼ arg½F ðjwÞ  90  S (3.62)


with j expressed in degree.
76 Modeling and dynamic behaviour of hydropower plants

In the Bode diagram with logarithmic scale, a slope S = 1 corresponds to 20


dB/decade. It is known that to avoid the roll-off phenomenon, it is recommended to
use the roll-off constraint associated to the slope of the magnitude of L around the
cut-off frequency6 wc [13]: 30 dB < roll  off < 10 dB , 1:5 < S < 0:5.
Thus, for the open loop read at w ¼ wc , we can apply the roll-off constraint
knowing that jLðjwÞj ¼ Lmp ðjwÞ:
3 1
 < S ½Lðjwc Þ ¼ S Lmp ðjwc Þ < 
2 2
Using the relation (3.62), we get: j½Lðjwc Þ < j Lnmp ðjwc Þ  45
We want L to be stable, so j½Lðjwc Þ > 135 to assure a good phase margin
(>45 ). Hence the robust stability condition is:
j Lnmp ðjwc Þ > 90 or j Gnmp ðjwc Þ > 90 (3.63)
Considering the approximated transfer functions G1, G2 and assuming x1 > 0 in
compliance with the surge tank criterion (see below), we set:
1  Twc s
Gnmp ðsÞ ¼ and Gmp ðsÞ ¼ K1 1 þ 2ðx1 =w1 Þs
1 þ ðTwc =2Þs
 
þ 1=w1 2 s2 1 þ 2ðx2 =w2 Þs þ 1=w2 2 s2
where the gain K1 is affected to the transfer function Gmp.
For w ¼ 1=Twc we have j Gnmp ðjð1=Twc ÞÞ ¼ 90 , so by applying the con-
dition (3.63) with a phase decreasing, the following criterion on the cut-off fre-
quency related to the Loop Transfer is yielded:
1
wc < (3.64)
Twc
Besides, a classic relation between the cut-frequency and the rising time (tm) of a
dynamic system is given by wc  1; 7=tm . Finally, we obtain the following condi-
tion for the penstock dynamics:
tm
Twc < (3.65)
1:7

3.5.2.2 Surge tank cross-section criterion


To limit oscillations of the surge tank water level (and the output power response),
it is necessary to compel x1 > 0. From (3.57), it is possible to obtain:
ð1  2  Twc  lc Þ
Tz  la 
ð1  2ðn2  Twa  la þ Twc  lc ÞÞ
(3.65a)
1
>
2ð1  2ðn  Twa  la þ Twc  lc ÞÞ
2

6
The cut-off frequency wc is the cross frequency corresponding to jLðjwÞj ¼ 1.
Reduced order models for grid connected hydropower plants 77

Then, we deduce:

1
Tz > (3.65b)
2la ð1  2  Twc  lc Þ

This inequality can be reformulated thanks to notations given in Section 3.2.4 to


find a capability criterion for the surge tank cross-section:

La
Sz >  (3.66)
2gKa Sa H0  2n2 Kc Q0 2

References
[1] Robert G., Michaud F., ‘Reduced models for grid connected hydro
power plant – application to generation control’, IEEE-CCCA, Hammamet,
2011.
[2] Robert G., Michaud F., ‘Hydro power plant modeling for generation control
applications’, ACC, Montréal, 2012.
[3] Chaudhry M. H., Applied Hydraulic Transients, 3rd ed., Springer-Verlag
New York Inc., New York, 2014.
[4] Nicolet C., ‘Hydroacoustic modelling and numerical simulation of unsteady
operation of hydroelectric systems’, Thesis, EPFL, 2007.
[5] Roberson J. A., Hydraulic Engineering, Wiley, New York, 1998.
[6] Wylie E. B., Streeter V. L., Fluid Transients in Systems, Prentice-Hall,
New York, 1993.
[7] Munoz-Hernandez G.A., Mansoor S.P., Jones D.L., Modelling and Con-
trolling Hydropower Plants, Springer, London, 2014.
[8] Robert G., Michaud F., ‘Flatness based control of a hydro power plant’,
IEEE-MELECON, Malta, 2010.
[9] Robert G., Michaud F., ‘Dynamic capability of hydro power plants for pri-
mary load-frequency control’, IFAC Power Plant and Power System Control
Symposium, Toulouse, 2012.
[10] Koehl A., Michaud F., Gubert S., Nicolas J., Libaux A., ‘A generic method
for the capability evaluation of hydraulic power plant to participate to
the load-frequency control (LFC)’, SHF ‘Enhancing Hydropower Plants’
Conference, Grenoble, April 2014.
[11] Kundur P., Power System Stability and Control, McGraw-Hill, New York,
1994.
[12] Eremia M., Handbook of Electrical Power System Dynamics: Modeling,
Stability and Control, Wiley, Hoboken (New Jersey), 2013.
[13] De Larminat Ph., Automatique Appliquée, 2nd ed., Hermes Lavoisier, Paris,
2009.
78 Modeling and dynamic behaviour of hydropower plants

[14] deMello, F. P., IEEE Working Group on Prime Mover and Energy Supply
Models for System Dynamics Performance Studies ‘Hydraulic turbine and
turbine control models for system dynamic studies’, IEEE Transactions on
Power Systems, vol. 7, pp. 167–179, 1992.
[15] IEEE Std 1207-2004, ‘Guide for the application of turbine governing
systems for hydroelectric generating units’.
[16] Kokotovic P., Singular Perturbations Methods in Control: Analysis and
Design, SIAM, USA, 1986.
Chapter 4
Modeling and stability analysis of turbine
governing system of hydropower plant
Wencheng Guo1,2, Jiandong Yang1
and Weijia Yang1,3

4.1 Introduction
Turbine governing system is the core component of load frequency control (LFC) of
hydropower plant [1–5]. During the transient process of LFC, the stability is the most
basic and important requirement of the turbine governing system [6,7]. Aiming at this
topic, this chapter first establishes the complete mathematical model for the turbine
governing system of hydropower plant without and with surge tank. Then, the
stability of the system without and with surge tank is analyzed, respectively.
For the modeling of turbine governing system: under the assumptions of isolated
operation and rigid water hammer, the linearized complete mathematical model for
the hydroturbine governing system of hydropower plant without and with surge tank,
which is used for analyzing the transient process and dynamic performance of the
turbine governing system under load disturbance, is established by combining the
submodels of pipelines, surge tank, turbine, generator, and governor.
For the stability analysis of turbine governing system: first, the stability of
turbine governing system without surge tank is analyzed. Based on the linearized
complete mathematical model for the turbine governing system, the comprehensive
transfer function and linear homogeneous differential equation that characterize the
dynamic characteristics of system are derived. Then, the stability domain that
characterizes the good or bad of stability quantitatively is drawn by using the sta-
bility conditions. First, the effects of influence factors, such as fluid inertia and
generator characteristics on the stability, are analyzed through stability domain.
Then, by proceeding in a similar manner, the basic stability analysis of turbine
governing system with surge tank is carried out. The effects of surge tank are

1
State Key Laboratory of Water Resources and Hydropower Engineering Science, Wuhan University,
Wuhan 430072, China
2
Maha Fluid Power Research Center, Department of Agricultural and Biological Engineering, Purdue
University, West Lafayette, IN 47907, USA
3
Department of Engineering Sciences, Uppsala University, Uppsala SE-751 21, Sweden
80 Modeling and dynamic behaviour of hydropower plants

investigated. The method for the enhancement of stability of turbine governing


system is proposed based on the analysis results. Finally, the critical stable sectional
area of surge tank is studied. According to the homogeneous differential equation
of turbine governing system, the analytical formula of critical stable sectional area
is deduced.

4.2 Modeling of turbine governing system


The pipeline and power generating system of isolated hydropower plant is shown in
Figure 4.1. The hydropower plant is composed of penstock, turbine, generator, and
governor. For the isolated hydropower plant with surge tank, it also includes
headrace tunnel and surge tank. During the operation of hydropower plant, the LFC
is actualized by the turbine governing system. The task of LFC contains two
aspects: (1) according to the arrangement of the load diagram, the output power of
the generating unit is rapidly adjusted by the turbine governing system with the
variation of the load and (2) the unpredictable load fluctuation in short period is

Upstream Generating unit


reservoir
Penstock

Downstream
reservoir

(a) Draft tube

Surge tank

Upstream Headrace tunnel Generating unit


reservoir
Penstock

Downstream
reservoir

(b) Draft tube

Figure 4.1 Pipeline and power generating system of hydropower plant:


(a) without surge tank and (b) with surge tank
Modeling and stability analysis of turbine governing system 81

undertaken by the turbine governing system through adjusting the frequency of


power system. The turbine governing system of hydropower plant is illustrated
in Figure 4.2.
The turbine governing system is a closed-loop system which is composed of
the turbine control system and the controlled system. The turbine control system is
constituted by some equipment which can detect the deviations between the actual
values and given values of controlled parameters (such as frequency, power output,
guide vane opening, pressure, and discharge) and then convert the deviations
of controlled parameters into the deviations of the displacement of the main
servomotor. The core component of the turbine control system is governor. The
controlled system is the system that controlled by the turbine control system. It
contains pipelines, turbine, generator, grid, and load. The controlled system is also
called the controlled plant.
The working process of the turbine governing system can be described as
follows. First, the values of controlled parameters (such as frequency, power out-
put, and guide vane opening) are measured by the measuring devices/sensors of
the turbine control system. These values are used as the feedback signals. Then, the
feedback signals are synthesized and compared with their given values to determine

Headrace tunnel
Turbine Grid
(surge tank)
generator load
Penstock

Controlled system
Measurement element

Amplification element – +
Actuator Point element
Correction element –

Feedback element

Turbine control system

(a)

Headrace tunnel
qt (surge tank)
r u Following y Hydro- Penstock h
Governor Guide
+ mechanism vane turbine mt
– Generator x
+ – load
mg
(b)

Figure 4.2 Turbine governing system of hydropower plant: (a) composite


structure and (b) control flow
82 Modeling and dynamic behaviour of hydropower plants

the deviations. Finally, the deviations are handled by the amplification element and
correction element, and then, they are used by the servomotor to control the
movements of the guide vane of turbine. As a result, the frequency and power
output of generator are regulated.
The output parameters of turbine governing system (such as frequency and
power output) have a direct influence on the control of hydropower plant, and this
influence is usually called the feedback effect. Both the turbine governing system
and the turbine control system are closed-loop systems. The difference between the
input signal and the feedback signal is called the error. The feedback is used by the
closed-loop system to reduce the error and then make the output parameters stable.
For a closed-loop governing system, the stability is the primary problem. The
dynamic process and dynamic quality of closed-loop governing system are much
more complicated than those of open-loop system. Even if the closed-loop gov-
erning system can keep a stable state, it also come up the phenomena of overshoot
and damped oscillation during the dynamic process.
The transient process of the pipeline and power generating system of hydro-
power plant is a complicated dynamic process that coupled by hydraulic, mechanic,
and electricity subsystems. Hence, the turbine governing system includes three
submodels: hydraulic, mechanic, and electricity. An accurate and complete math-
ematical model is significantly important for the study of the stability control
of turbine governing system. In addition, the research aims at the small load
disturbance condition.

4.2.1 Hydraulic submodel


Hydraulic submodel contains the mathematical model of pipeline (i.e., headrace
tunnel and penstock) and the mathematical model of surge tank.

4.2.1.1 Mathematical model of pipeline


When the elasticity of the water flow and the pipe wall are concerned, [8,9] the
basic equations (i.e., momentum equation and continuity equation) of unsteady
flow of pressurized pipe are as follows:

Q @H @H a2 @Q a2 Q @A Q
þ þ þ  sin q  ¼ 0 (4.1)
A @x @t gA @t gA @x A

@H @Q @Q fQjQj
gA2 þQ þA þ ¼0 (4.2)
@x @x @t 2D

where x is the position along the axis of the pipeline; q is the included angle
between the connecting line of centroid of the pipeline section and the horizontal
plane; A is the cross-sectional area; a is the wave speed of the water hammer; H is
the net head; Q is the discharge; f is the Darcy–Weisbach coefficient of friction
resistance; and D is the section diameter. For the nonprismatic pipe, we have
@A=@x 6¼ 0; and for the prismatic pipe, we have @A=@x ¼ 0.
Modeling and stability analysis of turbine governing system 83

In this chapter, we consider the condition of prismatic pipe. Then, we have


@A=@x ¼ 0. For (4.1), the values of ðQ=AÞð@H=@xÞ and sin q  ðQ=AÞ are much
less than the other terms, so these two terms can be neglected. For (4.2), the term of
@Q=@x is under the similar condition. Hence, (4.1) and (4.2) can be simplified to
the following two equations:

@H a2 @Q
þ ¼0 (4.3)
@t gA @t

@H @Q fQjQj
gA þ þ ¼0 (4.4)
@x @t 2DA

By changing the variables into the form of relative deviation, i.e., q ¼ ðQ  Q0 Þ=Q0 ¼
DQ=Q0 and h ¼ ðH  H0 Þ=H0 ¼ DH=H0 , and by conducting the Laplace transform
of (4.3) and (4.4), we can obtain (Note: the subscript ‘‘0’’ refers to the initial value. The
positive direction is set from the upstream to the downstream.):
   
Tr Tr
hðl; sÞ ¼ ch s þ a hð0; sÞ  2bsh s þ a qð0; sÞ (4.5)
2 2
   
1 Tr Tr
qðl; sÞ ¼  sh s þ a hð0; sÞ þ ch s þ a qð0; sÞ (4.6)
2b 2 2
 
1 Tr
hð0; sÞ ¼ hðl; sÞ þ 2bth s þ a qð0; sÞ (4.7)
chððTr =2Þs þ aÞ 2
 
1 Tr 1
qðl; sÞ ¼  th s þ a hðl; sÞ þ qð0; sÞ (4.8)
2b 2 chððTr =2Þs þ aÞ

where Tr ¼ 2L=a is the phase period of the water hammer wave; L is the length of
pipeline; b ¼ ðTw s þ f1 Þ=Tr s; Tw ¼ LQ=gHA is the water inertia time constant of
pipeline; f1 ¼ Dh=H0 ; Dh is the head loss of pipeline; a ¼ f1 Tr =2Tw ; s is the
complex variable; and l ¼ Dx=x0 .
Based on the boundary conditions, we can get the overall transfer function of
the pipeline in the condition of elastic water hammer. For example, for the headrace
tunnel in Figure 4.1(b), the transfer function of the head and discharge in the tail
end of headrace tunnel can be obtained as follows if the boundary condition of
upstream reservoir is hð0; sÞ ¼ 0:
 
hðl; sÞ Tr
¼ 2bth sþa (4.9)
qðl; sÞ 2

where hðl; sÞ and qðl; sÞ are the Laplace transforms of the relative deviations of the
head and discharge in the tail end of headrace tunnel, respectively.
84 Modeling and dynamic behaviour of hydropower plants

In this chapter, we consider the condition of rigid water hammer, i.e., the
elasticity of the water flow and the pipe wall are neglected. Based on the Newton’s
Second Law of Motion and the Laplace transform, we can get the momentum
equation of pipeline. For the penstock shown in Figure 4.1(a), its momentum
equation is presented as follows:

dqt 2ht0
h ¼ Twt  qt (4.10)
dt H0

For the headrace tunnel and penstock shown in Figure 4.1(b), their momentum
equations are respectively presented as follows:

dqy 2hy0
z ¼ Twy þ qy (4.11)
dt H0

dqt 2ht0
h ¼ Twt  qt  z (4.12)
dt H0

where Qy is the headrace tunnel discharge; Qt is the penstock discharge; z is the surge
tank water level (positive direction is downward) and Dz is the change of surge tank
water level; hy0 is the head loss of headrace tunnel; ht0 is the head loss of headrace
tunnel; Twy is the water inertia time constant of headrace tunnel; Twt is the water
inertia time constant of penstock; z ¼ Dz/H0, h ¼ (H  H0)/H0, qy ¼ (Qy  Q0)/Q0,
qt ¼ (Qt  Q0)/Q0 are the relative deviations of corresponding variables.

4.2.1.2 Mathematical model of surge tank


For the hydropower plant with surge tank (shown in Figure 4.1(b)), the continuity
equation of surge tank is as follows:
dz
qy ¼ qt  TF (4.13)
dt
where TF is the time constant of surge tank, TF ¼ FH0/Qy0; F is the sectional area of
surge tank.

4.2.2 Mechanic submodel


4.2.2.1 Mathematical model of turbine
The kinetic moment Mt and discharge Qt are the functions of guide vane opening Y,
head H, and unit frequency x. By using the forms of relative deviations of corre-
sponding variables, we can get [10–13]:

mt ¼ mt ðh; x; yÞ (4.14)
qt ¼ qt ðh; x; yÞ (4.15)
Modeling and stability analysis of turbine governing system 85

For the condition of small load disturbance, (4.14) and (4.15) can be dealt
with Taylor series expansion. If we neglect the second order and above traces,
we have:

@mt @mt @mt


dmt ¼ dh þ dx þ dy (4.16)
@h @x @y
@qt @qt @qt
dqt ¼ dh þ dx þ dy (4.17)
@h @x @y

Because of the small load disturbance condition, we have DMt  dMt . Then, based
on mt ¼ (Mt  Mt0)/Mt0, we have mt  dmt . In the same way, we can get qt  dqt ,
h  dh, x  dx and y  dy. Substitution of these relationships into (4.16) and (4.17)
yields:

@mt @mt @mt


mt ¼ hþ xþ y (4.18)
@h @x @y
@qt @qt @qt
qt ¼ hþ xþ y (4.19)
@h @x @y

Define the transfer coefficients of turbine as follows:


● Moment transfer coefficients of turbine:

@mt @mt @mt


eh ¼ ; ex ¼ ; ey ¼
@h @x @y

● Discharge transfer coefficients of turbine:

@qt @qt @qt


eqh ¼ ; eqx ¼ ; eqy ¼
@h @x @y
Hence, (4.18) and (4.19) can be changed into the following forms:
mt ¼ eh h þ ex x þ ey y (4.20)
qt ¼ eqh h þ eqx x þ eqy y (4.21)
It should be noted that (4.20) and (4.21) can be set up only in the case of small load
disturbance. With the change of operating points, the six transfer coefficients of
turbine changes too. The six transfer coefficients of turbine can be determined
from the comprehensive character curves of turbine by finite difference method.
For the ideal turbine, the values of the six transfer coefficients of turbine are
as follows:
eh ¼ 1:5; ex ¼ 1; ey ¼ 1; eqh ¼ 0:5; eqx ¼ 0; eqy ¼ 1 (4.21a)
86 Modeling and dynamic behaviour of hydropower plants

4.2.2.2 Mathematical model of governor


There are many types of governor, such as auxiliary receiver, damping, accelera-
tion damping, and parallel proportional integral derivative (PID). In this chapter,
we apply a simple governor which is simplified from the parallel PID type gover-
nor. Its equation is presented as follows:
dy dx
¼ Kp  Ki x (4.22)
dt dt
where Kp is the proportional gain, and Ki is the integral gain.

4.2.3 Electricity submodel


The resisting moment of turbine is denoted as Mg and its relative deviation is denoted
as mg. If there is an unbalance between the resisting moment and the kinetic moment,
the unit frequency would change. They should satisfy the following equation.
First derivative differential equation of generator [14–18]:
dx
Ta ¼ mt  ðmg þ eg xÞ (4.23)
dt
where Ta is called unit inertia time constant, Ta ¼ GD2 n2r =365Nr . It means the
required time for the unit to accelerate from 0 to rated rotor speed nr (r/min) under
the condition of rated kinetic moment and the rated output power Nr (kW). GD2 is
the rotational inertia of unit (tm2). eg is the load self-regulation coefficient.

4.3 Stability analysis of turbine governing system


Under isolated operation mode, the operating conditions of governor are compli-
cated due to the influence of the value of load change and the load characteristics of
isolated grid. Hence, maintaining the grid frequency within a certain range is
indeed a challenge. Some national standards stipulate the limit value of decay rate
for frequency response of isolated HPP under load disturbance. In practical cases,
the frequency response is required to have good regulation quality on the premise
of the satisfaction of stability. In this section, we first introduce the basic concepts
and criterion of stability of dynamic system (Section 4.3.1). Then, the stability of
turbine governing system without surge tank and with surge tank are analyzed
(Sections 4.3.2 and 4.3.3). Finally, the critical stable sectional area of surge tank is
derived (Section 4.3.4).

4.3.1 Basic knowledge of stability of dynamic system


4.3.1.1 Basic concepts of stability
The turbine governing system of hydropower plant is a kind of typical dynamic
system. The basic requirement of turbine governing system is stability and there is
no practical significance for an unstable system. For a dynamic system, some
system parameters will deviate from the equilibrium position and fluctuate under
disturbance. This process is usually called the transient process of dynamic system.
Modeling and stability analysis of turbine governing system 87

Based on the changed process of the amplitude of system parameter, the transient
process can be classified into three types [19,20]: (a) damped oscillation, (b) persistent
oscillation, and (c) diverging oscillation.
1. Damped oscillation
The amplitude of system parameter gradually decreases with the time. Finally,
the dynamic system can enter into a new equilibrium state. The process is illu-
strated in Figure 4.3(a).
2. Persistent oscillation
The amplitude of system parameter gradually keeps unchanged with the time.
The process is illustrated in Figure 4.3(b).
3. Diverging oscillation
The amplitude of system parameter gradually increases with the time. Finally,
the dynamic system cannot enter into a new equilibrium state. The process is
illustrated in Figure 4.3(c).
Both of damped oscillation and persistent oscillation are stable oscillation, while
diverging oscillation is unstable oscillation. The system that leads to unstable oscil-
lation is generally nonfunctional. Although persistent oscillation system is stable, its
amplitude does not decrease with the time. Hence, this kind of system cannot be used
in practical engineering. Only damped oscillation system has practical value. In this
chapter, the stability means the damped oscillation.
For the dynamic system, its transient process can be described by the following
linear homogeneous differential equation under linear hypothesis:

dn x dn1 x dx
a0 n
þ a1 n1
þ    þ an1 þ an x ¼ 0 (4.24)
dt dt dt
Equation (4.24) can characterize the dynamic characteristics of the dynamic system.
Its characteristic equation is

a0 ln þ a1 ln1 þ    þ an1 l þ an ¼ 0 (4.25)

The solutions of (4.24) is

X
n
xðtÞ ¼ ci eli t (4.26)
i¼1

where li are the roots of (4.25), which are called characteristic roots. They may be
real numbers, or imaginary number.
If the characteristic roots are all real numbers, we can obtain that the system is
unstable when there is at least one positive characteristic root. Hence, when the
characteristic roots are all real numbers, the necessary and sufficient condition of
making the system stable is that all the real roots are negative.
If (4.25) has a dual conjugate complex roots:

l1;2 ¼ a  ib (4.27)
88 Modeling and dynamic behaviour of hydropower plants

(a)
x

(b)
x

(c)

Figure 4.3 Transient process of dynamic system: (a) damped oscillation,


(b) persistent oscillation, and (c) diverging oscillation

Then, the two particular solutions of (4.24) corresponding to those complex roots
are as follows:
c1  l1 t 
x1 ¼ e þ el2 t ¼ c1 eat cos bt (4.28)
2
c2  l1 t 
x2 ¼ e  el2 t ¼ c2 eat sin bt (4.29)
2i
Modeling and stability analysis of turbine governing system 89

Hence,

x ¼ x1 þ x2 ¼ eat ðc1 cos bt þ c2 sin btÞ ¼ ceat cos ðbt  qÞ


qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
where c ¼ c21 þ c22 ; q ¼ tan1 ðc2 =c1 Þ: (4.30)

According to (4.30), we can get that: only when a is negative can the ampli-
tude x of system parameter gradually decreases with the time.
In conclusion, the necessary and sufficient condition of making the dynamic
system (4.24) stable is that all the roots of (4.25) have negative real parts.

4.3.1.2 Criterion of stability


Based on the analyses in Section 4.3.1.2, the stability of the dynamic system can be
judged by the roots of the characteristic equation of system. But for the linear
homogeneous differential equation of transient process of hydropower plant, the order
is usually relatively high and the characteristic equation is usually higher-degree
algebraic equation. The roots are extremely difficult to solve. So, researchers
proposed some stability judgment methods, which can be avoided to solve the
differential equation or characteristic equation of the dynamic system, and gave the
judging conditions. The conditions used to judge the stability of the dynamic system
are usually called the criterion of stability.
In this chapter, the two most common types of criteria are presented, i.e.,
Routh criterion and Hurwitz criterion [19,20].
1. Routh criterion
It is a kind of algebraic criterion. It can not only provide the information about
the stability of linear time-invariant system but also points out the number of
characteristic roots which locate at the imaginary axis and the right-half plane
of the complex plane, respectively.
Routh criterion is established by the relationship between the roots li and
coefficients ai of the characteristic equation of the dynamic system. According
to (4.25), the n-order characteristic equation can be denoted as

DðlÞ ¼ a0 ln þ a1 ln1 þ    þ an1 l þ an


¼ a0 ðl  l1 Þðl  l2 Þ    ðl  ln Þ
¼ 0 (4.31)
Based on the relationship between the roots and coefficients we can obtain:
8 a1
>
> a0 ¼ ðl1 þ l2 þ    þ ln Þ
>
>
>
>
>
> a2
>
> ¼ ðl1 l2 þ l1 l3 þ    þ ln1 ln Þ
>
>
> a
< 0
a3
¼ ðl1 l2 l3 þ l1 l2 l4 þ    þ ln2 ln1 ln Þ
>
> a0
>
>
>
> ..
>
>
>
> .
>
>
> an
: ¼ ð1Þn ðl1 l2    ln Þ (4.32)
a0
90 Modeling and dynamic behaviour of hydropower plants

Equation (4.32) shows that the following two conditions must be satisfied if
we want to make all the characteristic roots l1 ; l2 ; . . .; ln have negative real
parts, i.e., make the system stable.
i. The coefficients a0 ; a1 ; . . .; an of the characteristic equation are all nonzero.
ii. The symbols of the coefficients a0 ; a1 ; . . .; an are the same.
Based on the above two conditions, the necessary condition of system
stability is all the coefficients a0 ; a1 ; . . .; an are positive, i.e., ai > 0.
Routh table:
ln a0 a2 a4 a6 ...
ln1
a1 a3 a5 a7 ...
ln2
b1 b2 b3 b4 ...
ln3
c1 c2 c3 c4 ...
... ... ...
l 2
f1 f2
l 1
g1
l 0
h1
where
     
 a0 a2   a0 a4   a0 a6 
     
a a  a a  a a 
1 3 1 5 1 7
b1 ¼ ; b2 ¼ ; b3 ¼ ;...
a1 a1 a1
     
 a1 a3   a1 a5   a1 a7 
     
b b  b b  b b 
1 2 1 3 1 4
c1 ¼ ; c2 ¼ ; c3 ¼ ;... (4.32a)
b1 b1 b1
c; d; . . .; f ; g; h are defined by the same way as b; c.
For the linear dynamic system, the necessary and sufficient condition of system
stability is that all the elements in the first line of the Routh table are positive.
2. Hurwitz criterion
For Hurwitz criterion, the coefficients a0 ; a1 ; . . .; an of the characteristic
equation are used to judge the stability of the dynamic system. The Hurwitz
determinant constituted by the coefficients a0 ; a1 ; . . .; an is
 
 a1 a0 0 0 0 0 . . . 
 
 a3 a2 a1 a0 0 0 . . . 
 
 
 a5 a4 a3 a2 a1 a0 . . . 
 
D ¼  a7 a6 a5 a4 a3 a2 . . .  (4.32b)
 
 .. 
 . 
 
 . .. .. .. .. 
 . 
. . . . . an
Modeling and stability analysis of turbine governing system 91

The necessary and sufficient conditions of system stability are as follows:

a0 > 0
D1 ¼ a1 > 0
 
 a1 a0 

D2 ¼  >0
a3 a2 
 
 a1 a0 0 
 
 
D3 ¼  a3 a2 a1  > 0
 
a a a 
5 4 3
..
.
Dn > 0 (4.32c)

4.3.2 Stability analysis of turbine governing system


without surge tank
According to the mathematical model established in Section 4.2 and the criterion of
stability presented in Section 4.3.1, we can carry out the analysis of stability of
turbine governing system without surge tank and with surge tank, respectively. Let
us first consider stability of turbine governing system without surge tank.

4.3.2.1 Derivation of overall transfer function


For the turbine governing system without surge tank, the basic equations are as
follows:

dqt 2ht0
h ¼ Twt  qt
dt H0
mt ¼ eh h þ ex x þ ey y

qt ¼ eqh h þ eqx x þ eqy y

dy dx
¼ Kp  Ki x
dt dt

dx
Ta ¼ mt  ðmg þ eg xÞ (4.32d)
dt

According to the Laplace transforms of the above equations, the following overall
transfer function of turbine governing system without surge tank is obtained:
X ðsÞ sðb2 s þ b3 Þ=Ki
GðsÞ ¼ ¼ 3 (4.33)
Mg ðsÞ a2 s þ a3 s2 þ a4 s þ a5
92 Modeling and dynamic behaviour of hydropower plants

Then, we can get the linear homogeneous differential equation of the system:

d3 x d2 x dx
a2 3
þ a3 2
þ a4 þ a5 ¼ 0
dt dt dt
The expressions of coefficients in overall transfer function (see (4.34)) are presented
in Section 4.3.3.1.
By applying Hurwitz criterion, the stability criteria of turbine governing system
without surge tank represented by (4.34) are listed as follows:

D1 ¼ ai > 0ði ¼ 2; 3; 4; 5Þ
D2 ¼ a3 a4  a2 a5 > 0 (4.34)

When the coefficients in (4.34) satisfy the discriminants D1 > 0 and D2 > 0 simul-
taneously, the system without surge tank is stable.

4.3.2.2 Stable domain of turbine governing system


By substituting the characteristic parameters of the governing system in different
conditions into the stability criterion given in Section 4.3.1.2, the domain in which
the stability criterion is satisfied can be drawn in coordinate system for which the
abscissa and ordinate are usually 1/Kp and Kp/Ki, respectively. This domain is
called stable domain and its boundary is called stability boundary. The opposite
side of stability boundary is unstable domain.
A hydropower plant without surge tank is taken as example to illustrate the
stable domain. The basic information of this hydropower plant is listed in Table 4.1.
The ideal hydroturbine transfer coefficients are: eh ¼ 1.5, ex ¼ 1, ey ¼ 1,
eqh ¼ 0.5, eqx ¼ 0, and eqy ¼ 1. And the values of other parameters are as follows:
eg ¼ 0 and g ¼ 9.81 m/s2. Then, the stable domain, stability boundary, and
unstable domain are illustrated in Figure 4.4. Three working points A, B, and C
located in stable domain, stability boundary, and unstable domain, respectively, are
selected and their frequency responses under load disturbance are damped oscil-
lation, persistent oscillation, and diverging oscillation, respectively.

Table 4.1 Basic information of an actual example


of hydropower plant without surge tank

Parameters Values
Rated power output (MW) 30.93
Rated head (m) 110.00
Rated discharge (m3/s) 30.95
Twt (s) 2.01
ht0 (m) 2.07
Ta (s) 10.52
Lt (m) 503.13
Rated rotor speed (r/min) 375
Modeling and stability analysis of turbine governing system 93

4.3.2.3 Analysis of influencing factors on stability


Twt and Ta are two important parameters for the stability of turbine governing system
without surge tank. Their effects on the stability are shown in Figures 4.5 and 4.6.
Figure 4.5 shows that Twt has a significant effect on stability. With the homo-
geneous decrease of Twt, the stable domain increases homogeneously, i.e., the stability

20

15
Stable domain
Kp /Ki (s)

A
10

5 B Stability boundary

C
Unstable domain
0
0.0 0.2 0.4 0.6 0.8 1.0
1/Kp

Figure 4.4 Schematic diagram of stability domain of turbine governing system


without surge tank

20
Twt = 1.0 s
Twt = 1.5 s
Twt = 2.0 s
15
Twt = 2.5 s
Kp /Ki (s)

10

0
0.0 0.2 0.4 0.6 0.8 1.0
1/Kp

Figure 4.5 Effect of Twt on stability of turbine governing system without


surge tank
94 Modeling and dynamic behaviour of hydropower plants

20
Ta = 7.5 s
Ta = 8.5 s

15 Ta = 9.5 s
Ta = 10.5 s
Kp /Ki (s)

10

0
0.0 0.2 0.4 0.6 0.8 1.0
1/Kp

Figure 4.6 Effect of Ta on stability of turbine governing system without surge tank

of system becomes better, which indicates that the effect of water inertia in penstock
on the stability of system is unfavorable.
Figure 4.6 shows that Ta has little effect on stability. With the increase of Ta,
the stable domain increases, i.e., the stability of system becomes better.

4.3.3 Stability analysis of turbine governing system with surge tank


By the same method used in Section 4.3.2, the stability of turbine governing system
with surge tank is analyzed in the following section.

4.3.3.1 Derivation of overall transfer function


For the turbine governing system with surge tank, the basic equations are:
dqy 2hy0
z ¼ Twy þ qy
dt H0
dz
qy ¼ qt  TF
dt
dqt 2ht0
h ¼ Twt  qt  z
dt H0
mt ¼ eh h þ ex x þ ey y
qt ¼ eqh h þ eqx x þ eqy y
dy dx
¼ Kp  Ki x
dt dt
dx
Ta ¼ mt  ðmg þ eg xÞ (4.34a)
dt
Modeling and stability analysis of turbine governing system 95

According to the Laplace transforms of the above equations, the following


overall transfer function of turbine governing system with surge tank is obtained:

X ðsÞ sðb0 s3 þ b1 s2 þ b2 s þ b3 Þ=Ki


GðsÞ ¼ ¼ 5 (4.35)
Mg ðsÞ a0 s þ a1 s4 þ a2 s3 þ a3 s2 þ a4 s þ a5
Then, we can get the linear homogeneous differential equation of the system:

d5 x d4 x d3 x d2 x dx
a0 þ a1 þ a 2 þ a3 þ a4 þ a5 ¼ 0 (4.36)
dt5 dt4 dt3 dt2 dt
The expressions of coefficients in overall transfer function (see (4.34) and (4.36))
are as follows:
a0 ¼ f1 f9 a1 ¼ f1 f10 þ f2 f9 þ f5 f12
a2 ¼ f1 f11 þ f2 f10 þ f3 f9 þ f5 f13 þ f6 f12 a3 ¼ f2 f11 þ f3 f10 þ f4 f9 þ f6 f13 þ f7 f12
a4 ¼ f3 f11 þ f4 f10 þ f7 f13 þ f8 f12 a5 ¼ f4 f11 þ f8 f13
b0 ¼ f1 b1 ¼ f2
b2 ¼ f3 b3 ¼ f4
  
2ht0 2hy0
f1 ¼ eqh TF Twy Twt f2 ¼ TF Twy 1 þ eqh þ Twt eqh
H0 H0
f3 ¼ eqh ðTwy þ Twt Þ
 
2hy0 2ht0 2ðhy0 þ ht0 Þ
þTF 1 þ eqh f4 ¼ 1 þ eqh
H0 H0 H0
 
2ht0 2hy0
f5 ¼ TF Twy Twt f6 ¼ TF Twy þ Twt
H0 H0
2hy0 2ht0 2ðhy0 þ ht0 Þ
f7 ¼ Twy þ Twt þ TF f8 ¼
H0 H0 H0
f9 ¼ Ta =Ki f10 ¼ ðeg  ex Þ=Ki þ ey Kp =Ki
f11 ¼ ey f12 ¼ eh eqx =Ki  eh eqy Kp =Ki
f13 ¼ eh eqy
(4.36a)
Note that the expressions of coefficients in (4.34) are the special cases of those in
(4.36) when Twy, hy0, and TF are all 0.
By applying Hurwitz criterion, the stability criterion of turbine governing
system with surge tank represented by (4.36) are listed as follows:
0
D1 ¼ ai > 0ði ¼ 0; 1; 2; 3; 4; 5Þ
0
D2 ¼ a1 a2  a0 a3 > 0
0
D4 ¼ ða1 a2  a0 a3 Þða3 a4  a2 a5 Þ  ða1 a4  a0 a5 Þ2 > 0 (4.36b)
0 0
When the coefficients in (4.36) satisfy the discriminants D1 > 0, D2 > 0, and
0
D4 > 0, simultaneously, the system with surge tank is stable.
96 Modeling and dynamic behaviour of hydropower plants

4.3.3.2 Stable domain of turbine governing system


A hydropower plant with surge tank is taken as example to illustrate the
stable domain. The basic information of this hydropower plant is listed in Table 4.2.
The ideal hydroturbine transfer coefficients are: eh ¼ 1.5, ex ¼ 1, ey ¼ 1,
eqh ¼ 0.5, eqx ¼ 0, and eqy ¼ 1. And the values of other parameters are as follows:
eg ¼ 0 and g ¼ 9.81 m/s2. nf ¼ 1.0, where nf ¼ F/Fth, is amplification coefficient of
sectional area of surge tank, and Fth is the critical stable sectional area. Then, the
stable domain, stability boundary, and unstable domain are illustrated in Figure 4.7.

Table 4.2 Basic information of an actual example


of hydropower plant with surge tank

Parameters Values
Rated power output (MW) 610.00
Rated head (m) 288.00
Rated discharge (m3/s) 228.60
Twy (s) 23.84
Twt (s) 1.26
hy0 (m) 12.92
ht0 (m) 2.91
Ta (s) 9.46
Ly (m) 16,662.16
Lt (m) 530.69
Rated rotor speed (r/min) 166.7

20

Stable domain
15
A
Kp /Ki (s)

10 B

Stability boundary
5 C

Unstable domain

0
0.0 0.2 0.4 0.6 0.8 1.0
1/Kp

Figure 4.7 Schematic diagram of stability domain of turbine governing system


with surge tank
Modeling and stability analysis of turbine governing system 97

Three working points A, B, and C located in stable domain, stability boundary, and
unstable domain, respectively, are selected and their frequency responses under
load disturbance are damped oscillation, persistent oscillation, and diverging
oscillation, respectively.

4.3.3.3 Analysis of influencing factors on stability


nf, Twt, and Ta are the two important parameters for the stability of turbine
governing system without surge tank. Their effects on the stability are shown in
Figures 4.8–4.10.
Figure 4.7 shows that nf has a significant effect on stability. There is a critical
value of nf about the stability and this critical value corresponds to the condition
that the critical stable sectional area of surge chamber is equal to the actual area. If
it is less than its own critical value, the stability will become better when nf
increases. And with the homogeneous increase of nf, the range of increase of sta-
bility domain becomes larger. If it is greater than its own critical value, the stability
domain will increase in some area while decrease in other area.
Figures 4.9 and 4.10 show that the effects of Twt and Ta on stability of turbine
governing system with surge tank are similar as their effects on stability of turbine
governing system without surge tank. Twt has a significant effect on stability. With
the homogeneous decrease of Twt, the stable domain increases homogeneously, i.e.,
the stability of system becomes better, which indicates that the effect of water
inertia in penstock on the stability of system is unfavorable. Ta has little effect on
stability. With the increase of Ta, the stable domain increases, i.e., the stability of
system becomes better.

50

40

30
Kp /Ki (s)

nf = 0.50
nf = 0.75
20 nf = 1.00
nf = 1.25

10

0
0.0 0.2 0.4 0.6 0.8 1.0
1/Kp

Figure 4.8 Effect of nf on stability of turbine governing system with surge tank
98 Modeling and dynamic behaviour of hydropower plants

20
Twt = 0.76 s
Twt = 1.26 s
Twt = 1.76 s
15
Twt = 2.26 s
Kp /Ki (s)

10

0
0.0 0.2 0.4 0.6 0.8 1.0
1/Kp

Figure 4.9 Effect of Twt on stability of turbine governing system without surge tank

20
Ta = 10.46 s
Ta = 9.46 s
Ta = 8.46 s
15
Ta = 7.46 s
Kp /Ki (s)

10

0
0.0 0.2 0.4 0.6 0.8 1.0
1/Kp

Figure 4.10 Effect of Ta on stability of turbine governing system with surge tank

4.3.4 Critical stable sectional area of surge tank


For the hydropower plant with surge tank, the fluctuation stability of the water level
in surge tank is the most important challenge. The design of the sectional area of
surge tank is the most effective way to maintain the stability of the hydropower
plant with surge tank, and the critical stable sectional area is the minimal area of
surge tank that can keep the water level fluctuation stable.
Modeling and stability analysis of turbine governing system 99

For the fluctuation of the critical stable sectional area of surge tank, the Thoma
assumption is made, i.e., the governor is absolutely sensitive to maintain the output
of turbine generator constant. This assumption contains two points: simplified
assumptions: (a) the fluid inertia in the penstock is neglected. (b) the output and the
rotational speed of hydraulic turbine remain invariant.
Then, (4.12) can be transformed into the following equation according to
simplified assumptions:
h ¼ z (4.37)
According to the simplified assumptions (b), we can get: mt ¼ 0; x ¼ 0:
Finally, the second order linear homogeneous differential equation of turbine
governing system with surge tank is obtained based on (4..11), (4.37), (4.13),
(4.20), (4.21), (4.23), and mt ¼ 0; x ¼ 0:

d2 z dz
þ 2d w2 z ¼ 0 (4.38)
dt2 dt
        
where d ¼ vy0 =2 2ag=Ly  Ay =F ðH0  2ht0 Þ , w ¼ gAy =Ly F 1  2hy0 =
ðH0  2ht0 ÞÞÞ; a ¼ hy0 =v2y0 ; and vy0 is the flow velocity in headrace tunnel.
The period of water-level fluctuation in surge tank is obtained according to
(4.38): Tst ¼ pffiffiffiffiffiffiffiffiffi
2p ffi
. If the friction is neglected, the formula of period is simplified to:
w2 d 2

sffiffiffiffiffiffiffiffi
Ly F
Tst ¼ 2p (4.39)
gAy

Based on the criterion of stability presented in Section 4.3.1.2, the stability criterion of
turbine governing system without surge tank represented by (4.38) are listed as follows:
● Condition 1: D1 ¼ 2d > 0
● Condition 2: D2 ¼ 2dw2 > 0
From Condition 1, we can get:

Ly Ay
F> (4.40)
2ag ðH0  2ht0 Þ

From Condition 2, we can get:


 
H0 > 2 hy0 þ ht0 (4.41)
 
H0 > 2 hy0 þ ht0 is easy to satisfy in actual engineering. Hence, the control condi-
tion is (4.40). From (4.40), we can obtain the expression of the critical stable sectional
area of surge tank:
Ly Ay
Fth ¼ (4.42)
2ag ðH0  2ht0 Þ
100 Modeling and dynamic behaviour of hydropower plants

4.4 Conclusions
This chapter describes the modeling and stability analysis of turbine governing
system of hydropower plant without and with surge tank, respectively. Under the
assumptions of isolated operation and rigid water hammer, the linearized complete
mathematical model for the hydroturbine governing system of hydropower plant
without and with surge tank, which is used for analyzing the transient process and
dynamic performance of the turbine governing system under load disturbance, is
established. The stability of turbine governing system without and with surge tank
is analyzed, which includes the criterion of stability, stable domain and the effects
of influencing factors on stability. Finally, according to the homogeneous differ-
ential equation of turbine governing system, the analytical formula of critical
stable sectional area is deduced.

Acknowledgments
This work was supported by the National Natural Science Foundation of China
(Project No. 51379158) and the China Scholarship Council (CSC).

References
[1] Yang J., Zeng Z., Tang Y., Yan J., He H., Wu Y. ‘Load frequency control
in isolated micro-grids with electrical vehicles based on multivariable
generalized predictive theory’. Energies. 2015;8(3):2145–64.
[2] Wang T., Yang K. ‘Hydraulic control simulation and parameters optimiza-
tion for water diversion systems’. Journal of Hydraulic Engineering.
2006;37:1071–7.
[3] Guo W.C., Yang J.D., Yang W.J., Chen J.P., Teng Y. ‘Regulation quality for
frequency response of turbine regulating system of isolated hydroelectric
power plant with surge tank’. International Journal of Electrical Power and
Energy Systems. 2015;73:528–38.
[4] Guo W.C., Yang J.D., Chen J.P., Teng Y. ‘Study on the stability of water-
power-speed control system for hydropower station with air cushion surge
chamber’. In Proceedings of the IOP Conference Series: Earth and Envir-
onmental Science; Montreal, Canada, Sep 2014.
[5] Guo W.C., Yang J.D., Chen J.P., Teng Y. ‘Effect mechanism of penstock on
stability and regulation quality of turbine regulating system’. Mathematical
Problems in Engineering. 2014;2014:1–13.
[6] Yang K.L. Hydraulic Transient and Regulation for Hydropower Plants and
Pump Stations. Beijing: Water Resources and Electric Power Press; 2002.
[7] Wei S.P. Hydraulic Turbine Regulation. Wuhan: Huazhong University of
Science and Technology Press; 2009.
[8] Chaudry M.H. Applied Hydraulic Transients, 3rd ed. New York: Springer;
2014.
Modeling and stability analysis of turbine governing system 101

[9] Fang H.Q., Chen L., Dlakavu N., Shen Z.Y. ‘Basic modeling and simulation
tool for analysis of hydraulic transients in hydroelectric power plants’. IEEE
Transactions on Energy Conversion. 2008;23(3):834–41.
[10] Guo W.C., Yang J.D., Wang M.J., Lai X. ‘Nonlinear modeling and stability
analysis of hydro-turbine governing system with sloping ceiling tailrace
tunnel under load disturbance’. Energy Conversion and Management.
2015;106:127–38.
[11] Guo W.C., Yang J.D., Chen J.P., Yang W.J., Teng Y., Zeng W. ‘Time
response of the frequency of hydroelectric generator unit with surge tank
under isolated operation based on turbine regulating modes’. Electric Power
Components and Systems. 2015;43(20):2340–54.
[12] Guo W.C., Yang J.D., Chen J.P. ‘Research on critical stable sectional area of
surge chamber considering the fluid inertia in the penstock and character-
istics of governor’. In Proceedings of the ASME 26th Symposium on Fluid
Machinery; Chicago, IL, Aug 2014.
[13] Franklin G.F., Powell D., Naeini A.E. Feedback Control of Dynamic Systems.
Englewood Cliffs, NJ: Prentice Hall; 2009.
[14] Krivehenko G.I., Kwyatkovskaya E.V., Lyubitsky A.E., Ostroumov S.V.
‘Some special conditions of unit operation in hydropower plant with long
penstocks’. In Proceedings of Eighth Symposium of IAHR Section for
Hydraulic Machinery, Equipment and Cavitation; Leningrad, Soviet Union,
Sep 1976, pp. 465–75.
[15] Pennacchi P., Chatterton S., Vania A. ‘Modeling of the dynamic response of
a Francis turbine’. Mechanical Systems and Signal Processing. 2012;29:
107–19.
[16] Demello F., Koessler R., Agee J., Anderson P., Doudna J., Fish J. ‘Hydraulic-
turbine and turbine control-models for system dynamic studies’. IEEE Trans-
actions on Power Systems. 1992;7:167–79.
[17] Strah B., Kuljaca O., Vukic Z. ‘Speed and active power control of hydro
turbine unit’. IEEE Transactions on Energy Conversion. 2005;20(2):424–34.
[18] Kishor N., Saini R.P., Singh S.P. ‘A review on hydropower plant models and
control’. Renewable and Sustainable Energy Reviews. 2007;11(5):776–96.
[19] Liu Q.Z., Peng S.Z. Surge Tank of Hydropower Station. Beijing: China
Water Power Press; 1995.
[20] Streeter V.L., Wylie E.B. Fluid Transients. New York, NY: McGraw-Hill;
1978.
Part II
Control of hydropower plants
Chapter 5
Dynamic simulation issues for hydropower
generation control
Joël Nicolas1 and Gérard Robert2

5.1 Introduction
Dynamic simulation studies are widely applied in hydropower by researchers and
engineers for different stages of the life cycle of a Hydro Power Plant (HPP), from
the design to the operation period:
● process design studies with hydraulic transient analysis and control systems
specifications,
● feasibility studies to evaluate the capability to offer new services (power
capacity increasing, provision of ancillary services such as frequency control)
or to improve the performances (optimisation of the operating points,
upgrading procedure for aging power plants),
● monitoring and fault diagnosis to anticipate the maintenance and solve process
control problems.
This chapter will develop the context of simulation studies related to the design and
tuning of turbine governing systems for HPP, in relationship with the operation
of these power plants in large interconnected grids. It does not deal with the other
types of simulation analysis (e.g., voltage stability), as they are usually developed
in the electrical power system analysis books.
Progress made on computer technologies allows the engineer to develop non-
linear simulation models which represent more accurately the real dynamic beha-
viour of the HPP and enables us to solve more and more efficiently some particular
problems relevant to hydraulic transients.
Thus, there is an increasing interest among industries to use modeling and
simulation because field tests are not always possible in all operating conditions.
Indeed, simulation is a very flexible tool as it enables us to evaluate many hydraulic

1
EDF Hydropower Generation and Engineering, DTG, 21 Avenue de l’Europe BP 41, F38040 Grenoble
Cedex 9, France
2
EDF Hydropower Generation and Engineering, CIH, Savoie Technolac, 73373 Le Bourget du Lac,
France
106 Modeling and dynamic behaviour of hydropower plants

configurations for different control laws. It is an industrial tool not only to direct a
design choice for the process (turbine and its actuators, surge tank, penstock etc.)
and the whole control system (control loops, programmable logic controller (PLC)
and supervisory control and data acquisition (SCADA)), but also for an owner of a
power plant to write up more realistic specifications for suppliers of control sys-
tems and to help in economic decision, for instance in case of an upgrading of a
turbine governing system.
Generally speaking, it is very useful for feasibility studies associated with the
improvement of generation control performances regarding grid code requirements
with the corresponding controller tuning, or for understanding transient phenomena
by comparing field tests and simulation results and also for diagnosis to figure out
the causes of the failure.
This chapter is structured in three parts. Section 5.2 presents the context linked
to the European grid code requirements for frequency control. The experience of a
French hydroelectric power producer (EDF) concerning the use of computer simu-
lations is developed in Section 5.3, for power-frequency controller specifications,
feasibility studies and field tests carried out on large HPP. Finally, Section 5.4
concludes on the role of hydropower for integrating the intermittent renewable
sources in electrical power systems.

5.2 Grid codes requirements for frequency control


and balancing: example of the European network

5.2.1 General overview


Faced by the increased technical complexity of the electrical power systems within
a market-based framework, the different TSOs (transmission system operators)
around the world are developing new grid codes, with new definitions and corre-
sponding requirements with one main objective to avoid major blackouts.
The present section will give the example of the recent developments in Eur-
opean grid codes prepared to face the market rules and the large increase in inter-
mittent renewable energy sources, and therefore, the consequences for the design
and tuning of the control systems of the power plants with a focus on frequency
control and balancing process issues for HPP.

5.2.2 The European institutional context


In 2009, the European Union (EU) institutions adopted the ‘Third Energy Package’,
which is a set of two European Directives (gas and electricity) and three regulations
for the development of the Internal Energy Market. This ‘Package’ established and
gave key legal mandates to two types of associations:
● the ‘Agency for the Cooperation of Energy Regulators’ (ACER)
● the ‘European Network of Transmission System Operators for Electricity’
(ENTSO-E) as the association of European electricity TSOs; and the same for
gas (ENTSO-G)
Dynamic simulation issues for hydropower generation control 107

ENTSO-E now represents 41 TSOs across 34 European countries, with the objective
of assisting in the development of a pan-European electricity transmission network
in line with EU energy policy goals. These include:
● ensuring a secure and reliable operation of the increasing complex network
● facilitating the cross-border network development and the integration of
intermittent renewable energy sources (RES) along with system flexibility
● enhancing the creation of the internal market with a market-based approach
To achieve these objectives, the Regulation (EC, European Commission) no.
714/2009 provides ENTSO-E with a toolbox of tasks and responsibilities, including
‘network codes (NCs)’ (i.e., grid codes), infrastructure planning and adequacy
forecasts.

5.2.3 Brief presentation of the European interconnected


network ENTSO-E
In 2014, the 41 TSOs of the ENTSO-E have served 532 million final consumers,
for a total electricity consumption of 3,174 TW h; this accounts for about 15% of
the world’s total electricity consumption.
The geographical area covered extends beyond the EU, as shown in Figure 5.1,
with a large amount of exchanges (424 TW h in 2014) between the different
member TSOs.
There are several AC synchronous areas, in many cases connected together
by high voltage direct current (HVDC) links; the largest one is the Continental
Europe (CE) synchronous area, which is also connected by AC links with Turkey,
Albania and Maghreb; the other important ones are Northern Europe (NE), Great
Britain (GB) and Ireland (IRE).
Concerning the corresponding production, the net electricity generating total
capacity reached 1,024 GW in 2014, with the corresponding ratios of the different
energy sources as in Figures 5.2 (in capacity) and 5.3 (in energy).
Concerning the ‘other renewable’, the intermittent RES generation like wind
power or photo-voltaic (PV) is increasing more and more in capacity and energy,
and at certain periods already exceeds the consumption inside the perimeter of
some TSOs like in Denmark or in Germany. It, therefore, needs to export the excess
if the interconnection links are dimensioned enough, or to curtail a certain amount
of such RES generation, but this last case is not economically viable because this
RES generation is generally subsidised.
The ENTSO-E projections for 2030 and 2050 take into account such devel-
opment of intermittent RES with different scenarios, as shown in Figure 5.4.
Figure 5.5 also gives an example of the 2050 scenario, where the intermittent RES
(plus the run-of-the river generation) exceeds the load consumption at the scale of
the whole ENTSO-E area in the October period [1].
There is consequently a need for more and more flexibility, which can be
provided either by non-intermittent generation (like HPP with reservoirs), storage
or demand side response. New challenges are coming, after many years of mixed
hydro–thermal systems [2].
108 Modeling and dynamic behaviour of hydropower plants

IS

ENTSO-E members
FI
NO
SE
EE RU
LV
DK LT
RU
BY
IE GB
NL PL
BE DE
LU CZ UA
SK MD
FR CH AT HU
SI HR RO
BA
RS
IT ME BG
MK
AL
ES
PT
GR TR

TN
DZ CY
MA

Figure 5.1 Map of the ENTSO-E member (ENTSO-E website, 2015)

22% 20%
Hydraulic

Nuclear
12%
Fossil fuels

Other renewables

46%

Figure 5.2 ENTSO-E net generating capacity in 2014


Dynamic simulation issues for hydropower generation control 109

14.4%
18.5%
Hydraulic

Nuclear

Fossil fuels
26.3%
40.5% Other renewables

Figure 5.3 ENTSO-E energy net generation in 2014

2,500
Intermittent
Non-intermittent
2,000

1,500
GW

40% 46%
1,000 58% 77% 55% 35% 69%
17% 31% 31%

500

83% 69% 69% 60% 54% 42% 23% 45% 65% 31%
0
2012 v1 v2 v3 v4 x5 x7 x10 x13 x16
ENTSO-E 2030-ENTSO-E TYNDP 2050-e-Highway2050

Figure 5.4 ENTSO-E projections of intermittent installed capacities [ENTSO-E]

5.2.4 The development of European network codes


Since 2011, the ENTSO-E has devoted a lot of effort into the development of
different network codes for electricity, under the supervision of the EC and ACER,
and with the participation of distribution system operators (DSOs) and other sta-
keholders from across the electricity sector.
These network codes (NCs) are of three types:
● the ‘Grid Connection’ related codes, such as the ‘requirement for grid con-
nection of generators’ network code (RfG NC)
110 Modeling and dynamic behaviour of hydropower plants

900,000
ENTSO-E area
800,000 October 2050 – Scenario X-5
700,000
Load
600,000
500,000 Hydro
MW

RoR
400,000
300,000 PV
200,000
100,000 Wind

0
18 19 20 21 22 23 24 25 26 27 28 29 30 31

Figure 5.5 ENTSO-E RES generation, 18–31 October 2050 [ENTSO-E]

● the ‘system operation’ related codes, such as the ‘load frequency control and
reserves’ network code (LFCR NC)
● the ‘market’ related codes, such as the ‘electricity balancing’ network code
(NC EB)
Since mid-2015, some codes are going through the process of entering in force
for a transposition in National laws: for example, the ‘RfG NC’ was adopted on
26 June 2015 by EU Member States in ‘comitology’. After a review by the European
Parliament and Council who are checking its compliance with the main principles of
the EU and the third Energy Package (scrutiny), it is expected to become a binding
regulation in Europe in early 2016, which will mark the start of a 3-year imple-
mentation period across Europe.

5.2.5 Focus on some European requirements for frequency control


5.2.5.1 The ‘load frequency control and reserves’ network code:
general overview and specific requirements for ‘frequency
containment reserves’
The technical aspects of the frequency control provided by the generating units or
demand users are managed through the above-mentioned LFCR NC [3], whereas
the market-based rules are largely developed in the ‘electricity balancing’ NC.
Concerning the provision of active power reserves, the LFCR NC give new
definitions of such reserves:
● the ‘frequency containment reserves’ (FCR), which allow to stabilise the fre-
quency after a system imbalance disturbance, at a steady-state value by a joint
quick activation within the whole synchronous area (for synchronous gen-
erating units, it corresponds to the classical ‘primary frequency control’, and it
is automatically provided by their turbine governing system).
Dynamic simulation issues for hydropower generation control 111

Joint action within


synchronous area

Power/ LFC area


frequency
Reserve activation

Frequency containment process Frequency restoration process

FCR FRR Manual FRR RR

Reserve replacement process

Frequency

Time to restore frequency


Occurrence of the
disturbance

Figure 5.6 Dynamic hierarchy of load–frequency control processes

● the ‘frequency restoration reserves’ (FRR), which restore the frequency


towards its nominal value and replace the activated FCR (manually and/or
automatically activated; for generating units, it is provided by modification of
the power output set point).
● the ‘replacement reserves’ (RR), which restore/support the required level of
FRR to be prepared for additional system imbalances.
Figure 5.6 gives an example of dynamic hierarchy of load–frequency control
processes, under the assumption that FCR is fully replaced by FRR, and that FRR
and RR are triggered by the disturbed load–frequency control (LFC) area.
As a focus on the FCR, the draft ENTSO-E LFCR NC requires, for all FCR
providing units, some properties in the different European synchronous areas, as on
Table 5.1: for example, in CE, the maximum combined effect of inherent insensi-
tivity and intentional dead band of the governor should be of 10 MHz, and the FCR
full activation time should be of 30 s for a frequency deviation of  200 MHz.

5.2.5.2 The ‘requirements for grid connection of generators’ network


code: active power response of generating units in ‘frequency-
sensitive mode’
The RfG NC [4] gives some design requirements for the connection of power
generating modules (there are also some considerations in the demand connection
NC for the demand users or DSOs).
These requirements are gradually severe, according to different categories
depending on the connection point voltage and on the maximum capacity of these
112 Modeling and dynamic behaviour of hydropower plants

Table 5.1 Frequency containment reserves properties in the different synchronous


areas [ENTSO-E, draft load frequency control and reserves network code]

Minimum accuracy of frequency CE, GB, 10 MHz or the industrial


measurement IRE and NE standard if better
Maximum combined effect of inherent CE 10 MHz
frequency Response insensitivity and GB 15 MHz
possible intentional Frequency response IRE 15 MHz
dead band of the governor of the FCR NE 10 MHz
providing units or FCR providing groups
FCR full activation time CE 30 s
GB 10 s
IRE 15 s
NE 30 s if system frequency
is outside standard
frequency range
FCR full activation frequency deviation CE 200 MHz
GB 500 MHz
IRE Dynamic FCR  500 MHz
static FCR  1,000 MHz
NE 500 MHz

modules (with thresholds depending on the synchronous area). They shall apply to
new power generating modules (including pump-storage power plants).
Concerning the existing power generating modules, they are not subject to
these requirements, except where such a module ‘has been modified to such an
extent that its connection agreement must be substantially revised’ or ‘a regulator
or, where applicable, the Member State decide to make an existing power gen-
erating module subject to all or some of the requirements’.
Concerning the frequency control capability of the power generating modules, the
draft LFCR NC requires, from a certain capacity of these modules, a classical
‘frequency-sensitive mode’ (FSM) control, as in Figure 5.7, with the corresponding time
response from frequency step change as in Figure 5.8, and the different required para-
meters ranges or values for full activation of active power response as in Table 5.2.
Particularly, we can notice the value of 30 s for the ‘maximum admissible
choice of full activation time t2, unless longer activation times are allowed by the
relevant TSO for reasons of system stability’.
In addition, there are some requirements for ‘limited FSMs’ for both under-
frequency and over-frequency large disturbances.

5.3 Application to the design and tuning of turbine governing


systems: the French EDF experience
The present section gives an industrial example of using dynamic simulation for the
design and tuning of turbine governing systems, based on recent experience of EDF
Dynamic simulation issues for hydropower generation control 113

ΔP
Pmax

⏐ΔP1⏐
Pmax

s1

Δf
fn

⏐ΔP1⏐

Pmax

Figure 5.7 Active power frequency response capability in frequency


sensitivity mode in the case of zero dead-band and zero insensitivity
[ENTSO-E, requirements for grid connection of Generators
Network Code]

ΔP
Pmax

⏐ΔP1⏐
Pmax

t1
t/s
t2

Figure 5.8 Active power frequency response capability in time response


[ENTSO-E, requirements for grid connection of Generators
Network Code]
114 Modeling and dynamic behaviour of hydropower plants

Table 5.2 Required parameters ranges or values for full activation of active
power frequency response resulted from frequency step change in
frequency-sensitive mode

Parameters Range or values


Active power range related to maximum capacity 1.5%–10%
(frequency response range) jPDP 1j
max
For power generating modules with inertia, the maximum 2s
admissible initial delay t1 unless justified otherwise in line
with Article 15(2)(d)(iv)
For power generating modules without inertia, the maximum As specified by the
admissible initial delay t1 unless justified otherwise in line with relevant TSO
Article 15(2)(d)(iv)
Maximum admissible choice of full activation time t2, unless longer 30 s
activation time are allowed by the relevant TSO for reasons of
system stability

Hydro Generation and Engineering Division, which operates and maintains a large
number of hydropower plants in France, and which is currently modernising a large
part of turbine governing systems of these HPP.
For such a project of modernisation, new specifications have been prepared,
with a special attention paid to the provision of frequency control functions, with
the corresponding simulations and field tests [5]. The topic dealing with the flow
control or water level control is not studied here, but it sometimes interferes with
certain requirements for frequency control: as example, the duration time of FCR
delivering could be difficult to respect for certain cascaded HPP along rivers and
need a preliminary analysis [6]: for such cases, specific control functions have to be
developed to respect the required performance [7].

5.3.1 Frequency control and turbine governing systems


specifications
5.3.1.1 General approach
As mentioned in Section 5.2.5 with the European example, the TSOs are requiring
precise performance in delivering of active power reserves by the generating units,
and particularly a fast response for primary frequency control (FCR of Section
5.2.5.1). The HPP supplied by reservoirs are flexible energy sources, generally able
to generate quickly an extra amount of power for balancing needs (FRR and RR of
Section 5.2.5.1).
Concerning the primary frequency control carried out by their turbine gov-
erning system, HPP performance could be limited by the hydraulic transient phe-
nomena due to the water hammer in the hydro circuit, especially in the case of
undersized surge tanks or long penstocks.
To meet the grid codes requirements for frequency control, the performances of
the turbine governing system are specified in terms of static and dynamic aspects
Dynamic simulation issues for hydropower generation control 115

through a methodology based on modeling and simulations. The dynamic perfor-


mance requirements are particularly developed, with settling time/stability/robustness
criteria, for interconnected grid connection mode and optionally for isolated grid
connection mode.
In the past years, concerning the frequency control in interconnected grid
connection mode, EDF specifications required to tune the dynamic parameters of
the governing systems with the same values as in isolated network; for many HPP
with reservoirs, the corresponding time response of power output to frequency steps
was relatively slow.
The objective now is to fulfil as far as possible the corresponding ‘FSM’ require-
ments of the CE synchronous area, e.g., a maximum full activation time of 30 s for a
frequency deviation of  200 MHz, unless stability conditions cannot be obtained.

5.3.1.2 Dynamic fundamentals and corresponding requirements


We consider, hereafter, a typical high or medium head HPP, with hydro circuits as
in the layout represented in Figure 5.9.
For such a HPP, there are different identified technical constraints that repre-
sent a barrier for its participation to the primary frequency control or FCR service.
In accordance with Figure 5.10, we can consider the three following constraints:
● Opening slew rate: the opening and closing maximum speeds of the actuators
of the turbine, limiting the rising time to the desired power after FCR release.
● Water inertia: the water inertia time in the penstock, limiting the rapidity of the
dynamic response.
● Stability: the surge tank cross section, limiting the hydraulic transient stability
(related to the amplitude, damping factor and natural frequency of water flow
oscillations).
Thus, before on-site implementation of the new governing system, a numerical
simulation study of the behaviour of the HPP units with the turbine governing

Surge tank

Storage reservoir

Tunnel

Penstock

Turbines Downstream outlet

Figure 5.9 Diagram of a generic HPP considered for the simulator’s modeling
116 Modeling and dynamic behaviour of hydropower plants

Power

ΔP
Speed saturation of the Water inertia time in Surge tank damping
turbine actuators penstock oscillations

Time

Figure 5.10 Active power response of a hydro unit for a frequency step in primary
frequency control

system and the whole hydro circuit including the different water passages (tunnel,
penstock and surge tank) shall be produced, in order to obtain a suitable set of
parameters, which satisfy the performance criteria for different operating points of
the HPP. This simulation study has to be carried out in time domain and also in
frequency domain, in order to analyse the interaction between the hydraulic tran-
sients and the tuning of the turbine governing system.
An example of specification is presented hereafter, based on the schematic
diagram of the whole system in closed loop given in Figure 5.11 with a corre-
sponding schematic block diagram as in Figure 5.12, where:
● C is the transfer function of the equivalent main corrector of the digital gov-
ernor (PID or other type).
● G is the transfer function of the whole controlled process (servo-positioner,
turbine-including hydro circuits, generator), between the control input signal u
delivered by the corrector and the power output p (u, p are given in per unit).
● R is the transfer function corresponding to the processing of the power output
deviation Po  p in the digital governor, including the filtering and the pre-
sence of the permanent droop, if implemented on this deviation.
● Q is the transfer function corresponding to the processing of the frequency
deviation fo  f in the digital governor, including the filtering and the presence
of the permanent droop, if implemented on this deviation.
For different operating points (power set-point, head value and number of units in
operation), the structure and parameters of the governing system should lead to the
following required performance:
● in time domain, the expected dynamic response of the active power for a fre-
quency step (typically 200 MHz) should respect the following requirements
(Figure 5.13):
– rising time (90%): tm  25 s (as far as possible according to the respect of
the stability criteria as below)
– overshoot 1 (first oscillation) D1  30%
– overshoot 4 (fourth oscillation) D4  5%
Frequency measurement

Frequency set Generator Transformer


point –
+ Turbine
controller Servo-positioner Turbine Grid
PID
+–
Opening or power
set point
Position
feedback
Control mode choice
Power measurement

Figure 5.11 Simplified diagram principle of a turbine governing system in primary frequency control
118 Modeling and dynamic behaviour of hydropower plants

fo + – f

Q d
+ +
u
C G
+
R
+ – p
Po

Figure 5.12 Schematic block diagram of the primary frequency control loop of a
hydro unit

30% ΔP
10% ΔP
ΔP
90% ΔP

0 tm t

Figure 5.13 Active power step response (for a frequency disturbance)

● In frequency domain, the stability criteria must be respected, with the classical
gain margin MG and phase margin Mj of the open loop transfer function L
( jw) ¼ C( jw) * G( jw) * R( jw). The corresponding required criteria are the
following:
– MG > 6 dB
– Mj > 40
In addition to the above stability criteria, robustness criteria should be fulfilled,
for example in relationship with the sensitivity transfer function between dis-
turbance and power output, to limit the influence of the uncertainties of the
modeling: Sdp( jw) ¼ Dp( jw)/Dd( jw) ¼ 1/(1 + CGR), with the corresponding
required criterion:
– Max Sdp < 6 dB
Other sensitivity functions can be taken in account, such as the sensitivity to fre-
quency and power output measurement noises.
Moreover, for specific cases where the HPP has to be able to operate in iso-
lated grid connection mode, EDF defines specific performance requirements to
respect stability criteria in such a mode, with corresponding governing structure
and parameters.
Dynamic simulation issues for hydropower generation control 119

5.3.2 Simulation numerical studies: general issues


As a prescriber for modernisation of turbine governing system, EDF asks manu-
facturers for a numerical simulation study, to predetermine the best structure and
parameters of the digital controller in compliance with the requirements given in
Section 5.3.1.2, before the on-site implementation of the governor.
The simulation study could be only in time domain for HPP supplied by a simple
and not critical hydro circuit, but for HPP supplied by water passages equipped with
surge tanks, or by long penstocks, a complete modeling of this hydro circuit is
required, and both time domain and frequency domain studies should be carried out.
Such simulation results are given in Section 5.3.5 for a particular case.
Before this manufacturer’s study, for the more critical HPP, EDF performs a
preliminary numerical simulation study, to check the possibility of fast and
stable frequency control with a classical PID structure of governor, or on the contrary
to point out the attention on the difficulties to obtain easily such performances. These
simulations are carried out with a specific simulation tool, according to different
levels of complexity in the modeling of HPP, as detailed in Section 5.3.3.
Finally, the chosen parameters of the turbine governing system are imple-
mented by the manufacturer and tested during specified on-site tests, and if
necessary, modified in a contradictory discussion with a specialised testing team
within EDF: the corresponding testing results need to be carefully documented, to
be provided to the TSO, according to the national regulation.

5.3.3 Preliminary simulation numerical studies: principles


After many years of experience in simulation of control systems for hydropower
plants, EDF has developed for this purpose a ‘corporate simulator’ using
MATLAB/Simulink [8,9], to simulate the power response in primary frequency
control, for a generic HPP represented by Figure 5.8.
The classical theory of modeling the hydraulic behaviour of a HPP is largely
developed in the literature [10–14]. Based on this theory and on more recent
developments [15,16], Chapter 3 of this book gives more details (with the different
corresponding equations that model each part of the HPP) about the selected
complete model which was validated by comparison with field test results (see
Section 5.3.4).
The simulator can also be customised as a dedicated simulator if the topology
of the HPP is more complex, e.g., for a HPP with two water supplying reservoirs
and two surge tanks, and/or many penstocks. In all cases, the model validation from
measured data is necessary.
Two models are used with two specific objectives: a simplified linear model
and a non-linear model:
● The linear model is used to obtain – via simplified criteria – a quick overview
of the stability/rapidity compromise, from well-known tools of automatic
control, and then making the analysis easier.
● The non-linear model is used for a more precise analysis, as a real tool for
hydraulic transient simulation in which the simulated data must comply as
much as possible with the real behaviour.
120 Modeling and dynamic behaviour of hydropower plants

Therefore, the assumptions on the structure of the different sub-systems differ


between linear and non-linear models:

● In the linear case, the different water supply pipes including the penstock are
considered like rigid pipes (i.e. only the ‘mass water hammer’ phenomenon is
included in the model, the ‘wave water hammer’ is not considered); the rela-
tionship between power and turbine water outflow is linear; the cross section of
the surge tank is constant.
● In the non-linear case, the surge tank is modeled like a reservoir with variable
cross section varying with the height, and the head losses are included at the
diaphragm; the turbines are described by their abacus power/water flow/
actuator position; all the pipes include the elastic dynamics.

The developed tool will test in a first-time three ability indicators [6,9] from the
linear model and based on the three technical constraints limiting the power
response (opening slew rate, water inertia and stability), as mentioned in
Section 5.3.1.2.
The evaluation of these three criteria can roughly define the dynamic ability of
the considered HPP and helps to identify the probable cause of eventual incap-
ability. In case a criterion is not reached, it requires moving to the next stage of the
hydraulic transient simulation, to confirm or not the primary frequency control
capability with more accurate data.
For critical HPP for which the stability constraint is not fully satisfied, the
second step using the non-linear model is essential as it enables to calculate with an
optimisation algorithm the PID controller parameters, in accordance with the
rapidity and stability criteria and robustness analysis as defined in Section 5.3.1.2
and to validate the simulation for intermediate and extreme operating points taking
into account the nonlinearity of turbine characteristic curves. These simulations are
used to identify the undetected problems in the controller tuning phase, such as:

● The actuator solicitation rate taking into account the anti-windup control loop
and knowing that the goal is to solicit the actuator without excess.
● The risk of dewatering or discharge of the surge tank due to the water hammer
mass effect.
● The risk of overpressure in the penstock due to water hammer wave effect.

The modification of the PID parameter setting is the mean to mitigate the possible
above problems until achieving the respect of all technical requirements for the best
compromise between fast response in primary frequency control performance and
hydraulic safety.

5.3.4 Preliminary simulation numerical studies: results for some


HPP cases
The models presented in Section 5.3.3 were validated with real data on different
HPP, before their integration into the simulation tool. Compliance of simulation
with field test results is illustrated in Figures 5.14 and 5.15 for two HPP located in
Dynamic simulation issues for hydropower generation control 121

Opening gate
220
215 Real data
Simulation
210
205
200
mm

195
190
185
180
175
200 400 600 800 1,000 1,200
Surge tank water level
657
Real data
656.5 Simulation

656
m

655.5

655

654.5

654
200 400 600 800 1,000 1,200

Power
52
Real data
51 Simulation
50
49
48
MW

47
46
45
44
43
200 400 600 800 1,000 1,200
Time (s)

Figure 5.14 Simulation with a non-linear model and field test results for a 52 MW
hydro generating unit (Curbans HPP): step response in primary
frequency control
122 Modeling and dynamic behaviour of hydropower plants

Command
0.85
Opening gate (mm)

0.8

0.75

0.7

0.65
0 100 200 300 400 500 600 700 800 900
Power
10.5
Power (MW)

10

9.5 Real data


Non-linear model

9
0 100 200 300 400 500 600 700 800 900
Outflow pressure
6
Pressure (bar)

5.8

5.6

5.4
0 100 200 300 400 500 600 700 800 900
Time (s)

Figure 5.15 Simulation with a non-linear model and field test results for a 12 MW
hydro generating unit (Hautefage HPP): step response in primary
frequency control

France, characterised by the limited section of their surge tank: one named Curbans
on the Durance river in the South East (3  52 MW Francis turbines, rated rota-
tional speed ¼ 200 rpm, maximum head of 82 m) and the second one named
Hautefage in the Massif Central (2  12 MW Francis turbines, rated rotational
speed ¼ 375 rpm, nominal head of 60.7 m).
After this validation, EDF has been performing a lot of different feasibility
studies for ‘critical’ HPPs, to adapt the specifications for the provision of mod-
ernised turbine governing systems by the chosen manufacturers. Two examples
are presented hereafter in order to show which kind of results we can obtain.
The first study concerns the HPP of Saint Chamas on the Durance River, in the
South East of France (3  50 MW Francis turbines, rated rotational speed ¼ 200 rpm,
maximum head of 72 m). Figure 5.16 is an example of Bode diagram and
step response results from simulation through the linear model to both extreme water
head values. It corresponds to the best response of the considered HPP with optimal
Dynamic simulation issues for hydropower generation control 123

Bode diagram
Magnitude (dB) Gm = 8.33 dB / Pm = 90.7 deg
50
0
–50
–100
720
Phase (deg)

360

0
10–2 10–1 100 101 102 103
Frequency (rad/s)
Step response
1.2
1
0.8
Amplitude

0.6
0.4
0.2
Low head/high discharge conditions
0 High head/low discharge conditions
–0.2
0 50 100 150 200 250 300
Time (s)

Figure 5.16 Example of Bode diagram with the linear model in primary frequency
control at both extreme water head values for a 50 MW hydro
generating unit (Saint Chamas HPP)

PID controller setting. The goal is to verify that the phase and gain margins respect the
stability criteria, but also the robustness criteria (as presented in Section 5.3.1.2).
The second study concerns the HPP of La Saussaz in the French Alps (2  86 MW
Francis turbines, rated rotational speed ¼ 333 rpm, nominal head of 191 m), equipped
with two Francis turbines of 86 MW each. Figure 5.17 gives an example of simulation
results in time domain for the non-linear model.

5.3.5 Application for modernised turbine governing systems with


manufacturer’s simulations and performance field tests
As example of recent modernisation of turbine governing systems with such
numerical studies and complex hydraulic scheme, we can give the case of the large
mixed HPP of Grand’Maison in the French Alps (Figure 5.18). With a capacity of
1,800 MW, this HPP is equipped with four Pelton turbines of 157 MW each (five
injectors, rated rotational speed ¼ 428 rpm, nominal head of 918 m) and eight
reversible pump-turbines of about 150 MW each (rated rotational speed ¼ 600 rpm,
nominal head of 900 m in turbine mode and 948 m in pump mode).
Opening gate
90
85
80
% 75
70
–100 0 100 200 300 400 500 600 700 800

Discharge
44
42
40
m3/s

38
36
34
–100 0 100 200 300 400 500 600 700 800

Power
68
66
64
MW

62
60
58
–100 0 100 200 300 400 500 600 700 800
Time (s)

Figure 5.17 Power response in primary frequency control simulated for the
parameter tuning of a turbine governing controller (86 MW hydro
generating unit in La Saussaz HPP)

RESERV1 PIPEN1 STANK


PIPEN2 PIPEN9 PTURB9

Hupmax = 1695
Hupmin = 1590 PIPEN3 PIPEN10 PTURB10
RESERV3
PIPEN5
PIPEN4 PIPEN11 PTURB11
H = 772.6
PIPEN12 PTURB12

PIPEN13 FTURB1 PIPEN21


PIPEN6

PIPEN14 FTURB2 PIPEN22


PIPEN29
PIPEN15 FTURB3 PIPEN23
PIPEN7 RESERV2

PIPEN16 FTURB4 PIPEN24


Hdownmax = 768.5
Hdownmin = 740
PIPEN17 FTURB5 PIPEN25

PIPEN30
PIPEN18 FTURB6 PIPEN26
PIPEN8

PIPEN19 FTURB7 PIPEN27

PIPEN20 FTURB8 PIPEN28

Figure 5.18 Hydraulic scheme of the 1,800 MW Grand’Maison HPP (mixed four
Pelton turbine and eight reversible pump-turbine)
Dynamic simulation issues for hydropower generation control 125

Hmax Pmin –200 MHz


24

22
Setting 1
Setting 2
(%)

20

18

16
0 10 20 30 40 50 60 70 80 90 100

65

60 130% P
105% P
55 90% P
95% P
(MW)

50

45

40
11.3 s 11.7 s
35
0 10 20 30 40 50 60 70 80 90 100

940
(m)

920

900
0 10 20 30 40 50 60 70 80 90 100
Time (s)

Figure 5.19 Simulation of time responses to frequency step change for one Pelton
unit at 25% Pmax/Hmax (Grand’Maison HPP, other units stopped)

The governing systems of the four Pelton turbines have been recently moder-
nised with new digital controllers. After a preliminary feasibility study performed
by EDF as described in Sections 5.3.3 and 5.3.4, the manufacturer chosen for the
provision of the new governors performed a simulation study according to the
specifications detailed in Section 5.3.1.2. Figures 5.19 and 5.20 show time domain
responses of injectors position, power output and net head associated to one Pelton
turbine for a 200 MHz frequency step in interconnected grid connection mode
and for two different operating points: initial power output Po ¼ 25% Pmax, Hmax,
other units stopped (Figure 5.19); and Po ¼ 80% Pmax, Hmin, other units at Pmax or
126 Modeling and dynamic behaviour of hydropower plants

HminPmax –200 MHz


85

80

75
(%)

Setting 1
70 Setting 2

65

60
0 10 20 30 40 50 60 70 80 90 100

140 130% P
105% P
135
P 90%
95% P
(MW)

130

125

120
19.7 s 29.3 s
115
0 10 20 30 40 50 60 70 80 90 100

765

760
(m)

755

750
0 10 20 30 40 50 60 70 80 90 100
Time (s)

Figure 5.20 Simulation of time responses to frequency step change for one Pelton
unit in frequency control at 80% Pmax/Hmin (Grand’Maison HPP,
other units started)

80% Pmax (Figure 5.20). The black and grey dynamic responses correspond to two
different parameter settings.
As we can see, Figure 5.20 corresponds to the most stable case (maximum load).
Finally, field tests have been carried out by an EDF special testing team [17],
after modernisation of one turbine governing system with close values of the stu-
died parameters. Frequency steps of 200 MHz were applied to the digital gov-
erning system of the concerned Pelton unit around 136.5 MW, with two other
Pelton units at around 150 MW and the reversible units stopped.
Dynamic simulation issues for hydropower generation control 127

(%)
(bar) (Hz)
80 (MW)
∆F = –200 MHz Frequency (Hz) 50
150
Power output (MW)
70 93
∆P = 49 140
16.2 Rising time (90%) = 11.5 s
MW Average nozzle position (%)
60 130
90 48

120
50
Water pressure (bar) 47
87
110
40
10 20 30 40 50 60 70 80 90 100 Time (s)

Figure 5.21 Field test results: time response of one Pelton unit to frequency step
of 200 MHz in frequency control (Grand’Maison HPP)

The corresponding time-domain responses are given in Figure 5.21, with a


stable behaviour and a rising time (90%) of about 11 s, with a similar behaviour to
the simulation study, taking in account the operation conditions.
These field tests are very important to be completely confident about the
simulation study conclusions [18–20]; conversely, because these tests can be only
performed at some operating points of the units, the simulation study allows to
predict the behaviour of the HPP for other operating points.

5.4 Conclusion
In a context of an increasing part of the intermittent renewable sources of energy in
the electrical power systems, the HPP have a major role to play in providing
reserves because of their flexibility. Concerning the frequency control, the structure
and parameter settings of turbine governing systems need to be adapted to the new
requirements of the NCs; this can be possible with the help of recent developments
in the digital control systems and numerical simulation techniques.

References
[1] ENTSO-E. ‘Power System Vision and Action Paper’. 22 Aug 2014. Avail-
able from https://www.entsoe.eu/Documents/Publications/RDC%20
publications/140822_Power_System_Vision_and_Action_Paper.pdf
[Accessed 28 Jan 2016].
128 Modeling and dynamic behaviour of hydropower plants

[2] Nicolas J., Caillault B., Bouilliez J. ‘Ancillary generation services for the
security of large interconnected power systems: the major role of hydro-
power plants in mixed hydro-thermal systems’, Proceedings of the HYDRO
Conference; Porto, Portugal, October 2004.
[3] ENTSO-E. ‘Load Frequency Control and Reserves Network Code’, final
version and associated supporting paper. 28 June 2013. Available from
https://www.entsoe.eu/fileadmin/user_upload/_library/resources/LCFR/
130628-NC_LFCR-Issue1.pdf and https://www.entsoe.eu/fileadmin/user_
upload/_library/resources/LCFR/130628-NC_LFCR-Supporting_Document-
Issue1.pdf [Accessed 28 Jan 2016].
[4] ENTSO-E. ‘Draft Commission Regulation (EU) Establishing a Network
Code on Requirements for Grid Connection of Generators’. 2015. Available
from https://www.entsoe.eu/Documents/Network%20codes%20documents/
NC%20RfG/draft_ec_networkCodesJune.pdf [Accessed 28 Jan 2016].
[5] Nicolas J., Taloud J.F., Koehl A., Robert G., Demaya O. ‘Standardized
performance of turbine governing systems and grid codes requirements for
frequency control: specifications, optimization studies and tests’, Proceed-
ings of the HYDRO 2015 Conference; Bordeaux, France, October 2015.
[6] Koehl A., Michaud F., Gubert S., Nicolas J. ‘A generic method for the cap-
ability evaluation of hydraulic power plant to participate to the load–frequency
control (LFC)’. La Houille Blanche Journal; 2015, vol. 5, pp. 46–54.
[7] Robert G., Michaud F. ‘A Simple multi-objective control for cascaded hydro
power plants’, Proceedings of the IFAC World Congress; Milano, Italy,
2011.
[8] Robert G., Michaud F. ‘Hydro power plant modeling for generation control
applications’, Proceedings of the ACC Conference; Montréal, Canada, 2012.
[9] Michaud F., Robert G. ‘Dynamic capability of hydro power plants for pri-
mary load–frequency control’, Proceedings of the IFAC Power Plant and
Power System Control Symposium; Toulouse, France, 2012.
[10] Working Group on Prime Mover and Energy Supply Models for System
Dynamics Performance Studies. ‘Hydraulic turbine and turbine control
models for system dynamic studies’. IEEE Transactions; 1992, vol. 7, no. 1,
pp. 167–179.
[11] Kundur P. Power System Stability and Control. McGraw-Hill Professional,
New York; 1994.
[12] Munoz-Hernandez G.A., Mansoor S.P., Jones D.L. Modelling and Control-
ling Hydropower Plants. Springer-Verlag, New York; 2014.
[13] Chaudhry M.H. Applied Hydraulic Transients. Springer-Verlag, New York
Inc.; 3rd ed. 2014.
[14] Brekke H. ‘Frequency response analysis of hydroelectric power plants with
influence from a non-linearized frictional damping and the turbine char-
acteristics’. Modeling, Identification and Control Journal; 1985, vol. 6, no. 1,
pp. 21–37.
[15] Nicolet C. ‘Hydroacoustic modelling and numerical simulation of unsteady
operation of hydroelectric systems’. Thesis, EPFL, 2007.
Dynamic simulation issues for hydropower generation control 129

[16] Robert G., Michaud F. ‘Reduced models for grid connected hydro power
plant – application to generation control’. Proceedings of the IEEE-CCCA
Conference; Hammamet, Tunisia, 2011.
[17] Nicolas J., Caillault B., Planque J.L. ‘Speed and voltage control systems
acceptance field tests’, Proceedings of the Modelling, Testing & Monitoring
for Hydropower plants Second Conference (Hydropower & Dams);
Lausanne, 1996.
[18] Hannett L.N., Fardnish B. ‘Field tests to validate hydro turbine-governor
model structures and parameters’, Proceedings of the IEEE/PES 1994 Win-
ter Meeting; New York. IEEE Paper 94WM190-9PWRS. 1994.
[19] Nicolas J., Libaux A., Planque J.L. ‘Control systems of hydro plants: from
the design of high-performance systems to their identification for power
system study models’, Proceedings of the Third Modelling, Testing &
Monitoring for Hydro Powerplants Conference (Hydropower & Dams);
Aix-en-Provence, 1998.
[20] IEC 60308 International Standard ‘Hydraulic Turbines – Testing of Control
Systems’. 2nd ed., International Electrotechnical Commission, IEC Central
Office, Geneva, Switzerland; 2015. Available from www.iec.ch.
Chapter 6
Methods of signal analysis for vibration control
at hydropower plants
Olga Shindor 1 and Anna Svirina1

6.1 Introduction
Hydropower plants are units that work solely on renewable natural energy source
are becoming one of the frequently used sources of energy. At the same time, these
elements of energetic infrastructure appear to be objects of high potential risk and
thus require consistent diagnostics, prediction of possible failures and correspond-
ing technogenic risks, and assessment of all kinds of defects as early as possible.
The setting requires implementation of diagnostics instruments which can find and
then localize the defect in the process of its development, assess the factors which
affect it, create an efficient algorithm for early diagnostics of hydraulic units that
aim to ensure equipment robustness within the working period—including defini-
tion of the best time to withdraw the unit for repair to avoid failure in critical modes
of operation [1].
As hydraulic units are complex objects, consisting of a large number of com-
ponents and work units, the probability of failure remains relatively high within
their lifetime period [2]. To avoid risks, complex methods of technical inspection
for hydropower plants as a whole and its components are implemented to assess
equipment condition at each stage of the work cycle. Within this process, one of the
most important factors is the nature and magnitude of vibration of hydraulic unit as
a whole or its individual nodes [3]. Vibration usually becomes the most significant
indicator of hydraulic units malfunctioning.
The nature and magnitude of the vibration are used to assess defects and faults
at hydropower plants—thus leading to significant interest given to this problem by
scientists and engineers worldwide. Their theoretical and practical research devel-
oped understanding of the nature of defects in the hydraulic units, followed by a
number of proposed diagnostics methods [1,3]. Each of these methods focuses on
specific fault by assessment of particular node or a specific defect.
The main methods of diagnostics include visual checkups, endoscopy, and
ultrasonic flaw detection, control of electrical insulation parameters during

1
Kazan National Research Technical University named after A.N. Tupolev, Kazan, Russia
132 Modeling and dynamic behaviour of hydropower plants

continuous or routine inspections, control of the air gap in the generator is stopped
or when the machine is running, vibration diagnostics, temperature diagnostics.
Existing research had proven that vibration analysis of hydraulic units’ fixed parts
allows detecting up to 76% of defects. Thus, vibration diagnostics of hydraulic unit
are to be conducted once a year, and elevated level of vibration increases the
number of surveys—and this suggestion is fixed in normative documentation.
One of the steps to estimate the technical state of hydraulic units is the control
of its structural unit’s vibration. This type of control is performed in accordance
with normative documents. Normally, the vibration tests are accompanied by visual
inspections of equipment. These actions allow precisely enough evidence to eval-
uate operating condition of the unit. Vibration for hydraulic unit in this case is
measured within double scope of vibrating displacement. Vibration is normally
being tested both before and after each repair. The prerepair tests aim to determine
deviations from the normal equipment operation, to reveal hidden defects which
were not detected during the operation and to clarify what is necessary to perform
repairing. Testing after repair is implemented to determine its quality. The results
of vibration tests are to form conclusion on existing technical condition of the
hydraulic unit. As one of the most efficient tests, the vibration control is a man-
datory element for hydro unit’s service that assesses its technical condition [1].
Vibration is assessed separately for the hydro generator steel structures and
for the hydraulic unit’s support structures. Hydro generator testing is produced
at the rotating rotor. The main reasons for high vibration of hydro generator’s
stator steel structures can be one of the following: the insufficient density of the
composite core’s butt connections; an unsuccessful scheme stator winding; gen-
erator’s equalizing current, rotor distortion, or coiled circuit in the windings of rotor
poles [2].
The two main vibration components in case of hydro generator steel structures
vibration include: (i) 100 Hz frequency vibration component, which is referred as
high-frequency vibration and (ii) poly-harmonic low-frequency vibration (as a sum
of 4–5 lower harmonic vibration elements), which is referred as low-frequency
vibration. The frequency of the first harmonic low-frequency component is defined
as the shaft rotation frequency of hydraulic unit, whereas the other components are
derived as multiples of the first [4]. Evaluation of vibration of the stator steel
structures is carried out separately for high and low frequencies.
As the vibration signal is a multi-component one, the main task of vibration
diagnostics is the selection of informative component of the signal. For this pur-
pose, one can apply different methods. The most spread methods are the method for
measuring the overall vibration level, the crest factor method, spectral analysis,
envelope spectrum of high frequency (HF) vibration, cepstrum analysis. Typically,
a combination of several analytical methods allows to accurately diagnose the state
of equipment. The most commonly used indicator of vibro-diagnostics signal is the
energy spectrum, which shows amplitude-frequency changes in the signal when a
defect appears. The methods of spectral diagnostics are effective in the case of
stationary vibration analysis when time power of random and periodic components
is a constant.
Methods of signal analysis for vibration control at HPPs 133

On the contrary, if localized in frequency and time domains nonstationary


vibration for the allotment of local features of the signal appears, it is recommended
to use an alternative method of processing vibro-diagnostics data–wavelet trans-
form. Spectral method allows determining the frequency content of the signal
without simultaneous localization of components in the time domain. Wavelet
transform allows analyzing the fine structure of the signal through the use the
locality property of wavelets, due to the multiplication by window contained in the
basic functions.

6.2 Hydro units vibration control methodology:


implementation of wavelet transform
Hydraulic unit’s vibrating condition control methodology is based on the vibration
velocity of structural unit’s measurement and analysis. Control steps include the
following ones:
1. Estimating the organization points of vibration measuring of hydro units
Ve1 ðtÞ; Ve2 ðtÞ ; . . .; VeN ðtÞ, which are to be performed according to the
guidelines in the normative documents [4]. At this stage, hydro generator sta-
tor’s steel structure’s as well as hydro unit support structure’s vibration are to
be separately measured.
2. Defining precalculated values of the wavelet coefficients corresponding to the
vibration condition of the hydraulic unit as ‘‘excellent’’—Wex , ‘‘good’’—Wg ,
‘‘satisfactory’’—Wsat , and ‘‘unsatisfactory’’—Wunsat , along with the parameters of
the wavelet transform (the mother wavelet, limits, and step changes in scale factor).
3. Registering and discrete input of data Ve1 ðti Þ; Ve2 ðti Þ ; . . .; VeN ðti Þ into the pro-
cessing module.
4. Processing vibration signal on the basis of wavelet analysis (discrete wavelet
analysis to eliminate the noise of the signal) and continuous wavelet transform
to localize the signal features and calculate the sum of the wavelet coefficients
Wsum at the preestimated level.
5. Comparing calculated values of the wavelet coefficients Wsum to precalculated
Wex , Wg , Wsat , and Wunsat to evaluate the current vibration state of hydro unit.
Forecasting the prophylaxis period or making a decision on an emergency stop.
The corresponding algorithm is presented in Figure 6.1.
For analysis of the signal, we have suggested wavelet transform. Wavelet
transform is a relatively new mathematical tool for the time/frequency analysis of
nonstationary signals [5,6]. Up to now, it has been successfully used in the fol-
lowing areas: image processing [7], chemical engineering [8], and technical diag-
nostics [9].
The movable time/frequency window equally well identifies the low-frequency
(LF) and HF components of the signal: this provides an incomparably great advan-
tage in the analysis of signal’s local features, which is essentially absent in the
Fourier transform.
134 Modeling and dynamic behaviour of hydropower plants

А
Start
Calculate sum of absolute evaluation
wavelet coefficients в Wsum
Organization measuring vibration points
(Ve1(t),Ve2(t), … VeN (t))
Yes Vibration Prediction the
Input pre-calculated values of wavelet Wsum < Wex condition period of
coefficients “excellent” prophylaxis
Wоmp , Wхор, Wуdob , Whеуdоb and
No
parameters of wavelet transform
Yes
Wsum < Wg Vibration condition
“good”
Data registration and discrete
input VeN (ti) No
Yes
Vibration condition
Wsum < Wsat
Discrete wavelet transform (denoise) “satisfactory”

No
Continuous wavelet transform Yes
Wsum < Wunsat Vibration condition
“unsatisfactory”

А No
Vibration condition
“inadmissible”

Emergency stop

End

Figure 6.1 Wavelet transform–based algorithm for control of hydraulic


unit’s vibration

The wavelet transform decomposes the signal into a set of basic functions
[5,10,11]. The basic functions are scaled and shifted versions of the mother wavelet
function. A wavelet yðtÞ, as its name implies, is a small wave that increases and
decreases essentially in a small time period. To become a wavelet, a function has to
meet the following conditions.
First, it should have a zero mean:
ð
1

yðtÞdt ¼ 0 (6.1)
1

Second, it needs to meet the condition of boundedness:


ð
1

jyðtÞj2 dt < 1 (6.2)


1

Third, the wavelet function has to be localized in time and frequency domain, i.e.,
the function has to be within a finite interval on the time axis, and within its Fourier
transform on the frequency axis.
Methods of signal analysis for vibration control at HPPs 135

Scaled and shifted version of yðtÞ is defined as follows:


 
1 tb
ya;b ðtÞ ¼ pffiffiffi y (6.3)
a a
where a is the scaling parameter, and b is a time translation.
Decomposition of the signal into multiple wavelet bases allows analyzing band
signal with a different frequency. All the above advantages can be achieved by
implementation of the continuous wavelet transform.
Direct continuous wavelet transform of the signal SðtÞ is formalized in the
following form [10]:
ð
1  
1 tb
W ða; bÞ ¼ pffiffiffi SðtÞ  y dt (6.4)
a a
1

As it can be seen from (6.4), wavelet transform has the combined information of the
signal, of the analyzing wavelet y, b—shift of function y which is proportional to
the time dimension, a—scale that evaluates frequency. Continuous wavelet trans-
form has redundant information, which in this case is a plus, as it allows allocating
local features of the signal [12].
Continuous wavelet transform allows simultaneous analysis of the same signal
at different frequency scales with the level of detail inherent to a chosen scale.
Scale a is positively related with frequency resolution and negatively related to
time resolution: when scale a increases, frequency resolution increases and time
resolution decreases, but when scale a decreases, the frequency resolution decrea-
ses, whereas time resolution increases for the corresponding components of the
signal [12].
The main challenge of the wavelet transform implementation is visualization
of the wavelet coefficients. Wavelet transform of one-dimensional signal, that
converts a function of one variable into a set of wavelet coefficients, is a function of
two variables; scale and shift. One way to visualize the wavelet coefficients is to
place it on the scale/shift coordinates or projection on the plane ab enable to trace
amplitudes of intensity wavelet coefficients at different scales in time.
The wavelet transform is an effective mathematical instrument for localization
and classification of nonstationary signal’s singular points, which allows simulta-
neous analysis in time and frequency domains. Application of continuous wavelet
transform allows to localize the time points when the nature of the vibration signal
changes with high level of accuracy.
At present, spectral analysis is mainly used for vibrating condition control,
whereas for certain purposes, wavelet analysis seems to have an advantage. Spec-
tral analysis has a weak resolution in the localization of the signal’s frequency
component in the time domain, and wavelet analysis solves this problem. The
theory of the wavelet transform is in fact the advanced direction of spectral analysis
theory, and its main advantage is the simultaneous localization of time and fre-
quency domains features.
136 Modeling and dynamic behaviour of hydropower plants

6.3 Hydropower plant vibration diagnostics case study


6.3.1 Controlling object and measurement equipment
characteristics
In this case study, we have considered the following objects of control:
1. Vertical hydro generator—rated power, Pn 78,000 kW; rated rotational speed,
nn 57.7 revs/min; frequency, f 50 Hz; manufacturing plant ‘‘Sibelectrotyazhmash’’;
2. Rotary blade hydro turbine—diameter of the impeller, D1 10 m; head of water:
max Hmax 18.5 m, calculated Hcal 12.4 m, estimated minimum Hmin 6.5 m; turbine
power at Hcal Nt 80.5 MW; rotational speed nrmin 57.7 min; frequency accelerating
namin 119.0 min.
Measuring complex included the following elements:
1. Vibration sensors (18 pcs), whose output signal is proportional to the vibration
velocity;
2. The sensor of synchronization (laser timer), used to estimate the phase rela-
tions in the measurement input to the vibration of each rotor pole;
3. The matching unit which matches output signals from the sensors with the
input signals of the analog-to-digital converter;
4. Analog-to-digital converter (32-channel, 12-bit);
5. Portable personal computer.
Functional diagram of measurement system can be seen in Figure 6.2.
Vibration sensors measuring vibration of hydro generator stator’s steel struc-
tures are installed on both sides of each joint sector in the appropriate place as
required by legislating documentation [4], according to the arrangement of sensors
shown in Figure 6.3.
The measurements were performed on different vibration levels on running
generator in the following modes:
1. idle stroke without excitation;
2. idle stroke with excitation;
3. connected to a resistive load (20, 30, and 40 MW).
Hydroelectric support structure’s testing was performed by measuring: (1) the
radial vibration of the turbine and generator bearings in two mutually perpendicular

Vibration sensors

Analog-to-digital
Preamplifier Personal computer
converter

Phase sensor

Figure 6.2 Functional diagram of the implemented measuring equipment


Methods of signal analysis for vibration control at HPPs 137

Downstream

Number of
the sector
4

3 5
Phase sensor

2 6

1
Joint of
the sector
Vibration
sensor
Upstream

Figure 6.3 The layout of vibration sensors on the stator of hydraulic generator

directions, (2) the shaft’s beat in the zone of turbine and generator bearings in the
same directions, and (3) support of a heel and upper bracket’s vertical vibration
(Figure 6.4).
For vibration measurement of the supporting structures of hydraulic unit, the
following operating modes of the equipment were considered:
1. idle stroke speed variation from 0.4 to 1.0 times nominal;
2. idle stroke with excitation;
3. work under a load of 20, 30, and 40 MW.
The tests were performed before and after each major overhaul of hydraulic unit.
The vibration displacement data obtained were averaged for every 20 measurements,
4 s each.

6.3.2 Hydraulic unit’s vibration condition monitoring on the


basis of diagnostics data wavelet analysis
To assess the level of vibration of hydraulic units, technicians currently use hydro-
power plant’s spectral analysis. Assessment is performed for the level of vibration in
the double amplitude of vibration displacement [13]. The measured quantity is, as a
rule, velocity, and to determine the value of vibro displacement, spectral analysis is
used. On the spectrogram, the amplitude of the vibration component is determined.
138 Modeling and dynamic behaviour of hydropower plants

12

Relative vibration of Absolute vibration of


generator bearing RB generator bearing DS
Absolute vibration of Relative vibration of generator
generator bearing RB bearing DS
Absolute vertical vibration of generator
bearing

Absolute vertical vibration of


thrust bearing
Relative vibration of supporting Relative vibration of supporting
bearing RB bearing DS
Absolute vibration of Absolute vibration of
supporting bearing RB supporting bearing DS

Figure 6.4 The layout of vibration sensors on the supporting structures of


hydraulic unit

To evaluate the magnitude of vibration velocity of vibro displacement, one should


use the following expression:
V
S¼ (6.5)
2pf
The average value of velocity and the results of spectral analysis of vibration in
joint sections 3–4 of the stator of hydraulic generator is presented in Figure 6.5.
Methods of signal analysis for vibration control at HPPs 139
2,200 2,200
1,000
Ve, µm/s

Ve, µm/s
1,000
0 0
–1,000 –1,000
–2,200 –2,200
0 1 2 3 4 0 1 2 3 4
(a) t, s (b) t, s

400 400
350 350
300 300
250 250
Ve, µm/s

Ve, µm/s
200 200
150 150
100 100
50 50
0 0
0 25 50 75 100 125 150 175 200 225 250 0 25 50 75 100 125 150 175 200 225 250
(c) f, Hz (d) f, Hz

Figure 6.5 The average value velocity of vibration for the left side of the
section 4 of the hydro generator stator (number 7) and the spectrum
of vibration: (a) and (c) prior to repair; (b) and (d) after repair

Table 6.1 Left part six of hydraulic stator’s vibration level (pre-
and postrepair)

Before repair After repair


Level of low-frequency vibration 112 108
(magnitude of amplitude in microns)
Level of high-frequency vibration 2 2
(magnitude of amplitude in microns)

This hydraulic generator is a part of the hydraulic unit number 7, which is shown in
Figure 6.5(a) and (c) prior to repair and in Figure 6.5(b) and (d) after the repairing
was done, respectively.
The magnitude of left part of four hydro generator’s stator sector’s vibration
before and after the repair can be seen in Table 6.1. The data shows that repairing
had reduced low-frequency vibration, whereas high-frequency vibration remained
unchanged, i.e., decreased very slightly, as shown in Figure 6.5(d).
According to the suggested methodology, wavelet transform of the analyzed
signal is performed under the following parameters:
1. The mother wavelet—Daubechies 3 (db 3); the mother wavelet in this case is
selected on the basis of center frequency wavelet, as it is necessary to analyze
the signal components of 1.4, 2, 3, 4, 5, and 100 Hz. The central frequency of
the Daubechies 3 (db 3) wavelet is Fr ¼ 0:8 Hz [14,15].
2. Scale factor—the maximum scale factor was chosen on the basis of the signal
analysis at a frequency of 1.4 Hz; thus, using wavelet Daubechies 3, a is estimated
140 Modeling and dynamic behaviour of hydropower plants

1,600

1,250
1,000
W (8,b)

750
500
250
0
0 50 100 150 200 250 300 350 400 450 500
(а)

1,600

1,250
1,000
W (8,b)

750
500
250
0
(b) 0 50 100 150 200 250 300 350 400 450 500

Figure 6.6 The wavelet coefficients at scale a ¼ 8 (100 Hz): (a) prior to repair
and (b) after the repair

equal to 2,048, and for a step change in the scale factor, we select Da ¼ 7:85. To
reduce the cost of hardware and the time to compute step changes of the scaling
factor, it is selected so that for each scale no more than 512 wavelet coefficients
are calculated. Fewer coefficients are calculated, as determined by experiment,
with no effect on their value or the interpretation of their results.
The results of the wavelet transform signal in correspondence with parameters
given below are shown in Figure 6.6.
Next, the vibration speed sensor, on the left side of the section 1, is considered
before and after the repair of hydraulic generator GA-7. The corresponding
(Figure 6.5(a) and (b)) and the results of the wavelet transform of the signal
(Figure 6.6(a) and (b)).
Signal identified prior to the repair of the hydro generator has an instant
increase of the vibration level at time points marked 13, 17, 33, 37, 57, 62, and 69,
and also the short-term changes at the point of 20 and 50 s. These changes are
absolutely absent in the signal after the repair of hydraulic generator was performed.
Application of spectral analysis evaluates HF and LF vibration and allows
estimating reduction of vibration as a result of repairs, while the wavelet analysis
provides information about the frequency character of these features. In case of
regular evaluation of hydraulic unit’s vibration state at the time of withdrawal from
equipment repair and for corresponding routine inspection, the use of spectral
analysis is effective, as it allows evaluating the presence of a defect in the signal
Methods of signal analysis for vibration control at HPPs 141

components of certain frequency. But for control of critical equipment operation


modes, wavelet analysis appears more efficient as it allows localizing particular
signal in the frequency and time domains. For the control of a vibrating condition
of the hydraulic unit, the total value of the wavelet coefficients should be used.
As it can be seen in Figure 6.6(b), after the planned repairs, the signal contains
a periodic component that does not lead to significant changes of wavelet
coefficients. Instant increase in the vibration level is expressed as an abrupt change
of the wavelet coefficients at the frequency of 100 Hz (see Figure 6.6(b)). After
repairs all nonstationary that was present in the signal prior to the repair were
eliminated.
Visualization of wavelet analysis results can be used to localize changes in the
time domain-vibrating signal and determine the frequency range of the fault. Thus,
this type of analysis can define any change in the vibration signal, localized in time
and frequency domains.
To follow this path, one should also consider the possibility of localizing
vibration signal characteristics in the frequency domain. Figure 6.7 shows the
absolute values of the wavelet coefficients at frequencies of 1.4, 2, 3, 4, and 5 Hz.
A significant increase in the wavelet coefficients can be observed at a fre-
quency of 1.4 Hz (see Figure 6.7(a)) and 2 Hz (see Figure 6.7(c)). At the frequency
of 3 Hz (Figure 6.7(e)) and 4 Hz (Figure 6.7(g)), one can also find a short-term
increase in wavelet coefficients, but its magnitude is much lower. At a frequency of
5 Hz the values of wavelet coefficients prior and after the repair are almost iden-
tical (Figure 6.7(h) and (i)). Thus, the wavelet transform allows determining both
frequency and nature of short-term changes in the values of instantaneous vibration
velocity. The obtained data indicates high sensitivity of the wavelet coefficients,
which increases the possibility to acquire instantaneous information about emer-
ging changes in the signal with the ability to assess it in frequency domain.
As vibration control of hydraulic units is required, and the procedure for its
implementation and results evaluation is regulated, the main task in creating
wavelet analysis based control systems is to define the relationship between nor-
mative evaluation criteria and wavelet coefficients. At present, normative evalua-
tion criterion is the average value of the vibration displacement, and its relation to
wavelet coefficients needs to be defined.
The average value of the vibration displacement is evaluated as a sum of the
wavelet coefficient’s absolute values at frequencies 1–10 Hz assessed at each 1 Hz
step:
X
Wsum ¼ jW ða; bÞj (6.6)
a

The values of the coefficient are defined within a selected range of 80–800
(scale factor spacing is unfixed) as the choice of scale factor is estimated regarding
the need to evaluate coefficients of at a particular frequency level. Thus, the
following values are selected as the scale factor: 800, 400, 267, 200, 160, 133,
114, and 100. For all of the calculations, the chosen mother wavelet, Daubechies 3
(db 3), is used.
142 Modeling and dynamic behaviour of hydropower plants
15 15

10 10

5 5

0 0
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80
(a) t, s (b) t, s

15 15

10 10

5 5

0 0
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80
(c) t, s (d) t, s
15 15

10 10

5 5

0 0
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80
(e) t, s (f) t, s
5 5
4 4
3 3
2 2
1 1
0 0
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80
(g) t, s (h) t, s

5 5
4 4
3 3
2 2
1 1
0 0
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80
(i) t, s (j) t, s

Figure 6.7 Absolute values of the wavelet coefficients 1 section steel structures
at 1.4 Hz: (a) prior to repair; (b) after repair; (c) and (d) 2 Hz;
(e) and (f) 3 Hz; (g) and (h) 4 Hz; and (i) and (j) 5 Hz

Analysis of the experimental data allows to define the relationship between the
value of the wavelet coefficients and the average value of vibration displacement
(Table 6.2).
Control of supporting structures of hydraulic unit’s vibrating condition
includes the following vibration measurements: absolute support and generator
bearings vibration measuring, as well as measuring of vertical vibration for the
support heel and upper bracket.
The results of measuring hydraulic unit’s GA-10 radial vibration of generator
bearing in two mutually perpendicular dimensions, working under a load of 20 MW
are shown in Figure 6.8(a) and (b).
Methods of signal analysis for vibration control at HPPs 143

Table 6.2 The correspondences between the total value of the wavelet
coefficients and average magnitude of vibro displacement

Evaluation Excellent Good Satisfactorily Unsatisfactorily Inadmissible

Parameter
The average <50 50–100 101–144 145–180 >180
magnitude of
vibro displace-
ment A, mm
Sum of absolute <0.17 0.17–0.57 0.58–0.79 0.8–1 >1
values of the
wavelet coeffi-
cients, Wsum

1,500 1,500
1,000 1,000
Ve, µm/s
Ve, µm/s

0 0

–1,000 –1,000
–1,500 –1,500
0 10 20 30 40 50 60 70 80 0 10 20 30 40 50 60 70 80
(а) t, s (b) t, s

1.2 1.2
Inadmissible Inadmissible
1 Unsatisfactory 1 Unsatisfactory
Wsum

Wsum

0.75 Satisfactory 0.75 Satisfactory


0.5 Good 0.5 Good
0.25 0.25
0 0
0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80
(c) t, s (d) t, s

Figure 6.8 Fluctuations of vibration velocity of the generator bearing: (a) the
right bank; (b) the downstream; the sum of the absolute values of the
wavelet coefficients: (c) the right bank; and (d) the downstream

The results of the wavelet transform of these signals evaluated as the sum of
wavelet coefficients Wsum absolute values are calculated according to (6.2) and
illustrated in Figure 6.8(c) and (d). According to Table 6.2, one can state that
generator bearing of the hydraulic unit GA-10’s absolute vibration level in the
searched mode of operation can be evaluated as one within the ‘‘good’’ area. The
values of the wavelet coefficients are homogeneous in nature with no visible spikes
or dips. The values of the wavelet coefficients of generator bearing vibration mea-
sured in the direction to the right bank are in the range 0.247, the average magnitude
of the vibro displacement is 61 mm. The values of the wavelet coefficients for
vibration generator bearing, measured in the downstream direction are in the range of
0.26, and the average magnitude of the direction vibro displacement is 67 mm.
144 Modeling and dynamic behaviour of hydropower plants

Thus, the average value of the absolute vibration generator bearing appears to
be 64 mm, and the average value of the wavelet coefficients absolute values sum is
Wsum ¼ 0:25. Evaluation of the generator bearing’s absolute vibration level, which
was obtained by means of spectral analysis, matches the one defined by means of
wavelet transform. Thus, same data is derived from both wavelet and spectral
analysis that evaluate vibrations from other sensors—and hence wavelet transform
can be used for periodic monitoring of structural units at hydropower plants as a
part of system of continuous monitoring system that assesses minimal vibration
fluctuations.

6.4 Conclusions
Development of vibration condition monitoring system of hydraulic unit on the
basis of wavelet transform allows efficient controlling of equipment in the oper-
ating mode and has several advantages over spectral analysis. The use of wavelet
transform includes not only vibration analysis that aims to define in time interval
moment of change in the state of equipment but also to predict the time for its
development. This increases effectiveness of detecting defects at an early stage of
development, which is very important in the case of hydroelectric power plant as
such analysis provides more flexibility in avoidance of hydraulic units mal-
functioning prevention.

References
[1] Shirman, A.R., Soloviev, A.B. Practical vibration diagnostics and mon-
itoring of mechanical equipment. Moscow, 1996. 276p.
[2] Glazyrin, G.V. Development of models and methods for vibration diag-
nostics of hydraulic units. Ph.D. thesis, Novosibirsk, Novosibirsk Technical
University, 2006, 338p.
[3] Vladislavlev, L.A. Vibration of hydraulic unit hydro power stations.
Moscow: Energy, 1972. 176p.
[4] STO 17330282.27.140.001-2006 Estimation methodologies of technical
condition of the main equipment of hydro power plants. RAO ‘‘UES of
Russia’’, 1995. 8p.
[5] Daubechies, I. Ten lectures on wavelets. Philadelphia, Pensylvania, Society
for industrial and applied mathematics, 1992, 378p.
[6] Massoputs, P.R. Fractal functions, fractal surfaces and wavelets. Amster-
dam: Academic Press, 1994, 379p.
[7] Vetterli, M., Kovacevic, J. Wavelets and subband coding. Englewood Cliffs,
NJ, 1995, 235p.
[8] Planisic, P., Petek, A. Characterization of corrosion processes by current
noise wavelet-based fractal and correlation analysis. Electrochimica Acta,
2008, vol. 53, no. 16, p. 5206–5214.
Methods of signal analysis for vibration control at HPPs 145

[9] Korepanov, V., Kulesh, M., Shardakov, I. Using of wavelet analysis for the
processing of experimental vibrodiagnostics data. Perm, 2007. 50p.
[10] Astafijeva, N.M. Wavelet analysis: basic theory application examples.
Successes Physical Sciences, 1996, vol. 11, p. 1145–1170.
[11] Mallat, S. A wavelet tour of signal processing. San Diego: Academic Press,
1998, 478p.
[12] Smolentsev, N.K. Foundations of the wavelets theory. Moscow: DMK Press,
2005. 304p.
[13] RD 34.31.303-96. Guidelines for operational control of a vibrating condition
of structural units of hydraulic units, Moscow: Standartinform, 1996, 26p.
[14] Nemtarev, V. The strategy of building systems, diagnostics of technical
condition of hydraulic units of vibratory field. Tatarstan Energy, 2005,
vol. 2, p. 76–79.
[15] Barkov, A.V. Basic requirements to modern condition monitoring and
diagnostics of machines and equipment [online]. VAST; vibroacoustic sys-
tems and technologies, St. Petersburg, 2011 [accessed 28.04.2014]. Avail-
able: http://www.vibrotek.ru.
Part III
Operation, scheduling, etc. of hydropower plants
(including pumped storage)
Chapter 7
Island mode operation in hydropower plant
Roshan Chhetri1 and Karchung2

7.1 Introduction

It is said the unit is islanded when the particular unit is disconnected from the
grid and is made to supply a certain area of load by itself. It is generally very
difficult to successfully supply the island load as both the frequency and voltage
has to be maintained at the range based on the sudden change of load from
interconnected to island load. It is said that unit can supply island load if the unit
can successfully supply load without getting tripped. In most of the cases, this is
not possible as the unit should have a best controller to overcome the problem
stated earlier.
In grid mode, the power generating station or power plant is in parallel with the
grid. As the total plant is connected to grid, the frequency and voltage is
uncontrollable and depends on the grid voltage and frequency. But the load is
controllable so is the power factor, as we can set the load of an unit to 5 or 8 MW
and set the power factor to 0.80 and alternator will generate power following the
pre-set commands.
In grid mode:
● Frequency and voltage depends on grid.
● Active and reactive power supplied to the grid can be controlled.
In Island mode, there is no grid connected, only a certain isolated area is
covered. This happens in case of marine or ship and factories with own supply and
consumption. We cannot control the amount of load supplied, as it totally depends
on load demand in the covered area, but we can control the voltage and frequency.
In islanded mode:
● Active and reactive power supplied to the bus bar depends on load.
● Frequency and voltage of the bus bar can be controlled.

1
Department of Electrical Engineering, College of Science and Technology, Phuentsholing, Bhutan
2
Jigme Namgyel Engineering College, Bhutan
150 Modeling and dynamic behaviour of hydropower plants

7.2 Performance in island mode


The first task is to investigate the stationary and dynamic behavior with the existing
hardware and controller settings. The stationary equation in speed control mode is:
ðPref  yD Þ  s þ ðnref  n þ nTrim Þ ¼ 0 (7.1)
The deflector position yD is a function of Pact (actual power), the upper water level
(gross head), and the load of the water hydraulic system. Dependent from the
flow through the other units the function can be derived from the model data
(Figure 7.1).
The opening position of the deflector in no load operation is about 31.3%. In
the controller, nTrim is set to 2.051% and the permanent droop s with closed/opened
Generator Circuit Breaker (GCB) is set to 6.0/10.0%. For nominal speed (1 p.u.)
before synchronization, nTrim should therefore be set to:
nTrim ¼ ðPref  yD Þ  s ¼ 3:13% (7.2)
Obviously, nTrim is also used to calibrate the speed feedback signal or the speed
reference is meant to set smaller than 1 p.u. With correct calibrated signals the
speed in no load operation (before synchronization) will be:
n ¼ nref þ nTrim  yD  s ¼ 98:922% (7.3)

Deflector position/active power


100
2 needles
90 operation

80

70
Deflector position in %

60

50 5 needles
operation
40

30

20
6 units operation
10
1 unit operation
0
0 20 40 60 80 100 120 140 160 180 200
Power in MW

Figure 7.1 Different needle operating cases


Island mode operation in hydropower plant 151

The speed reference in interconnected operation with activated speed track option
is always in the range 47:5  nref  52 Hz ð0:95  1:04 p:uÞ. After transition to
islanded load supply, the speed reference will usually be on the upper or lower limit
of the speed band. Also, the limits of over- and underspeed protection have to be
taken into account where an overshoot leads to a governor stop signal or GCB trip
signal. In case of islanded load without frequency dependence of the load, the new
theoretical stationary frequency can be calculated. It depends from the old and the
new operational point. Assuming there is only one unit supplying the island load,
the speed settles for islanded load of 30, 90, and 150 MW.
It can be seen that there are only a few stable settlement points between the
speed band and the overspeed and underspeed protection. Point inside of the band
causes intermittent switches between speed control and power control mode. This
can be described as controller instability.
Figure 7.3 shows some examples of stable and unstable transition. For the
transient process, the limit of electrical over speed detection at 55 Hz ð1:1 p:u:Þ as
well as the plant disconnection limit of 47 Hz ð0:94 p:u:Þ has to be taken into
account (lines marked with * in Figure 7.1). The speed band limit in power control
mode is marked with solid lines; the smaller speed band entry because of the
hysteresis is marked with dashed lines. The speed band settings for different loads
are shown in Figure 7.2.
For cases when (7.1) gives a speed settlement between 1:04  n  1:1, a
stable transition can be achieved, when the transient overshoot doesn’t cross the
limiting lines. Figure 7.3 shows only for the transition from 70 to 50 MW. For the

Speed settlement in speed control mode


1.25
Stationary speed 30 MW island load (2no)
Stationary speed 90 MW island load
1.2 Stationary speed 150 MW island load
Over-/underspeed
Speed band exit
1.15
Speed band entry
Speed in p.u.

1.1

1.05

0.95

0.9
0 50 100 150 200
Power reference in MW

Figure 7.2 Speed settlement in speed control mode at different island loads
152 Modeling and dynamic behaviour of hydropower plants

Turbine speed
1.15
Transient limit overshoot Unstable transition
1.1 Stable transition
Speed in p.u.

1.05

Controller instability
1

0.95

0.9
0 50 100 150 200 250 300
Time in s

Figure 7.3 Speed simulation of transitions from interconnected to islanded


operation, initial load 70 MW, islanded load 40, 50 (stable),
and 80 MW

transition from 70 to 80 MW, a controller instability can be expected, where the


controller continuously switches between NPC and PC mode and foments sustained
oscillations.
Figure 7.4 shows a generalization of the transition combinations in a grid of
10 MW steps with the above-mentioned boundary conditions. There is a wide band
of steady stable working points, but due to the transient overshoot, only a small
band of dynamically stable transitions exists. A maximal load step of only 30 MW
can be balanced. Also, the change of operation mode, particularly from two- to
five-nozzle operation is problematic because of associated speed turbulences.
Because of transient overshoot, not all of the shown valid points are stable. The
stable point investigated in the grid for 10 MW steps is shown in Figure 7.4. The
dynamic behavior of the load considered in the simulation has usual frequency load
factor of 2 p.u.
It can be seen that there is very small band of stable island mode transitions.
Load steps of maximal 30 MW can be balanced, if the power reference settings are
well chosen. Higher steps cause speed limit overshoots, even if the stationary points
are stable.
The areas marked 1 and 2 are points, where stationary speed is in the valid
band, but only the yellow points are successful transitions. Individual anomalies
can occur due to transient events or different net frequencies.
In the following sections, as an example one stable and two unstable transitions
are presented and discussed.
Figure 7.5 shows a typical case during which the unit generating of 70 MW in
interconnected mode is suddenly made to run at the 50 MW islanded load. As we can
see, there are some transients created during this transition, but it could successfully
gain transient stability after 70 s approximately, at 1:06 p:u: ð53 HzÞ of frequency.
Island mode operation in hydropower plant 153

190
180
ers ry
170
d
und ationa
160 pee
St

150

5 needles operation
140
Islanded load in MW

130 1
120
110
nd
100 ba
d 2
90 ee
Sp
80
70 1
60
50
40

operation
2 needles
30
20
y
onar
10 Stati eed
s p
0 over
0 10 20 30 40 50 60 70 80 90 100 110 120130 140150 160170 180 190
Power in interconnected operation/power reference in MW

Stable range
Steady stable but transient speed overshoot
Instability due to high-speed deviation
Controller instability due to mode switching
Semi-stable because of load frequency dependence

Figure 7.4 Island mode operating region for test HPP

In this case, we can say that the unit can successfully withstand island load. Now, let
us see how the unit behaves if the load at islanded condition is of decreased and
increased magnitude as demonstrated in Figures 7.6 and 7.7, respectively.
As we can see from Figure 7.6, the transient overshoot is of higher magnitude
as compared to the case in Figure 7.7. Although it can gain transient stability after
1 min, however, due to transient overshoot, as it crosses the speed band, machine
trips due to operation of over-speed relay.
On other hand, it is indicated in Figure 7.7 that with increased islanded load of
80 MW, the unit suddenly experiences reduced speed, followed by convergence
into transient instability having sustained oscillations in speed. As every generator
is required to operate successfully with in defined speed band, and whenever the
unit’s speed crosses this band, over-speed or under-speed relay trips the generator.
This safe guards the generator from going into unstable region of generator.
Another case study conducted is the change from two-to-five needle operation
(both directions) in islanded mode of generator. This transition in most of the cases
154 Modeling and dynamic behaviour of hydropower plants

Turbine speed
1.12

1.1

1.08

1.06

1.04
Speed in p.u.

1.02
Simulation 70 MW –> 50 MW
Over-/underspeed
1
Speed band exit
0.98 Speed band entry

0.96

0.94

0.92
0 50 100 150 200 250 300
Time in s

Figure 7.5 Frequency plot showing stable transition from 70 MW in


interconnected to 50 MW islanded load

Turbine speed
1.15
Electrical speed limit
overshoot
1.1
Speed in p.u.

1.05

Simulation 70 MW –> 40 MW
Over-/underspeed
1
Speed band exit
Speed band entry

0.95

0.9
0 50 100 150 200 250 300
Time in s

Figure 7.6 Frequency plot showing transient overshoot at transition from


70 MW in interconnected to 40 MW island load
Island mode operation in hydropower plant 155

Turbine speed
1.15
Simulation 70 MW –> 80 MW
Over-/underspeed
Speed band exit
1.1 Speed band entry
Speed in p.u.

1.05

Controller instability
1

0.95

0.9
0 50 100 150 200 250 300
Time in s

Figure 7.7 Frequency plot showing controller instability at the transition from
70 MW in interconnected mode to 80 MW islanded load

Turbine speed
1.12

1.1
Transition from 5- to
1.08 2-needles-operation
1.06
Speed in p.u.

1.04

1.02

0.98

0.96 Simulation 60 MW –> 50 MW –> 30 MW


Over-/underspeed
0.94 Speed band exit
Speed band entry
0.92
0 50 100 150 200 250 300
Time in s

Figure 7.8 Transition from five-to-two-needle operation

develops interaction with crossings of the speed band and leads to generator
instability. A change of number of needles in operation during islanded mode
causes higher turbulences. Figures 7.8 illustrates the variation in speed, during
transition to islanded mode with load of 50 MW and further load reduced to 30 MW
at 150s. As observed, initially, speed over-shoots but settles down reasonably.
Turbine speed
1.2

1.15

1.1

1.05
Speed in p.u.

0.95

0.9 Simulation 40 MW –> 30 MW –> 50 MW


Over-/underspeed
Speed band exit
0.85
Speed band entry

0.8
0 50 100 150 200 250 300
(a) Time in s

Turbine speed
1.12

1.1
Transition from 2- to
1.08
5-needles operation
1.06

1.04
Speed in p.u.

1.02

0.98

0.96 Simulation 40 MW –> 30 MW –> 50 MW


Over-/underspeed
0.94 Speed band exit
Speed band entry
0.92
0 50 100 150 200 250 300
(b) Time in s

Figure 7.9 Transition from two-to-five-needle operation. There is change in


nozzle opening from five to two in Figure 7.9(a) and from two to five
in Figure 7.9(b) at the time period of 150 seconds. When there is
change of nozzle from five to two, there is a greater vibration created
on the turbine as this proved that the turbine is hunting to stabilize
for the correct load but the transition is very smooth when there is
change of operation of nozzle opening from two to five as the pressure
in the penstock is released quickly and the turbine speed is stabilized
with load. The former is usually case when there is load shading
and latter is due to load being increased
Island mode operation in hydropower plant 157

Turbine speed
1.1

1.05

0.95
Speed in p.u.

0.9

0.85

0.8

0.75
Simulation 40 MW –> 70 MW
0.7
Over-/underspeed
0.65 Speed band exit
Speed band entry

0 50 100 150 200 250 300


Time in s

Figure 7.10 Two-to-five-needle operation

However, further reduction in load followed by transition from five-to-two needles


operation, results into one swing of oscillation, but the generator remains stable.
In case, as shown in Figure 7.9(a), during transition from two-to-five needle
operation, frequency dependency load is not considered, a sustained oscillation is
observed. While, using frequency dependent load, the response settles down briefly
after few swings of oscillations. This is suggested in Figure 7.9(b).
In general the transition from two-to-five-needle operation is critical in island
operation, because the control action of needles tend to be slow. This fact leads
higher deviations of speed during the transition. The transition from five-needle to
two-neddle operation is much faster because the closing action of the needles is
higher and in addition, the deflector is also able to stabilize the speed.
There is also an inherent instability in islanded operation during the transition
from two-needle to five-needle operation, which occurs during certain transitions,
shown in Figure 7.10. In this case of simulation, load changes to 70 MW in islanded
operation, with initial load being at 40 MW.

7.3 Measures to improve the island mode performance


There are some theoretical and also practicable solutions to improve the perfor-
mance of the plant. Some of the suggested solutions can be summarized as follows:
● Increase the speed of reaction (improves the overshoot performance).
● Eliminate the problems with the deflector feedback (ensures fastest reaction).
● Downsize the speed band.
● Operate without speed track (automatic controller path switchover must be chan-
ged, only manual switch back to power control mode after island mode detection).
158 Modeling and dynamic behaviour of hydropower plants

Turbine speed
1.12

1.1
Simulation 130 MW –> 50 MW
1.08 Over-/underspeed
1.06

1.04
Speed in p.u.

1.02

1
Without speed track and
0.98 automatic switch back
to power control
0.96

0.94

0.92
0 50 100 150 200 250 300
Time in s

Figure 7.11 Simulation without speed track

● Change of parameter settings.


● Change of controller structure.
A first simulation without speed track is shown in Figure 7.11. The system is
able to withstand a load step of 80 MW. Further improvements can be reached by
controller/parameter optimization.

7.4 Conclusion

Island operation of hydropower plant is fully discussed. Problems associated during


island operation are also explained. Different measures to operate a hydropower
plant in island operation are also mentioned.

Bibliography
[1] IEEE Committee. 1973. Dynamic models for steam and hydro turbines in
power system studies. IEEE Transactions on Power Apparatus and Systems;
92:1904–1915.
[2] Qijuan C. and Zhihuai X. 2000. Dynamic modeling of hydroturbine
generating set. In: IEEE International Conference on Systems, Man and
Cybernetics, IEEE, 8–11 Oct. 2000, pp. 3427–3430.
Island mode operation in hydropower plant 159

[3] Acakpovi A., Hagan E. B., and Fifatin F. X. 2014. Review of hydropower
plant models. International Journal of Computer Applications (0975–8887);
108(18), pp. 33–38.
[4] Bosona T. G. and Gebresenbet G. 2010. Modeling Hydropower Plant System
to Improve Its Reservoir Operation. Department of Energy and Technology,
Swedish University of Agricultural Sciences, Box 7032, 750 07 Uppsala,
Sweden.
[5] Kozdras K. 2015. Modeling and Analysis of a Small Hydropower Plant and
Battery Energy Storage System Connected as a Microgrid. University of
Washington, Seattle, WA, USA.
[6] Machowski J., Bialek J., and Bumby J. 2008. Power System Dynamics, 2nd ed.
West Sussex: Wiley.
[7] Yang W., Yang J., Guo W., et al. 2015. A mathematical model and its
application for hydro power units under different operating conditions.
Energies; 8:10260–10275; doi:10.3390/en80910260.
[8] Holst A., Golubovic M., and Weber H. 2007. Dynamic model of hydro
power plant ‘‘Djerdap I’’ in Serbia. In: IYCE Conference, Hungary.
[9] Holst A., Karchung, Chhetri R., and Sharma D. 2015. Analysis and modeling
of HPP Tala/Bhutan for network restoration studies. In: IYCE Conference,
Italy.
Chapter 8
Hydro generation scheduling: non-linear
programming and optimality conditions
Lucas S.M. Guedes1,2, Adriano C. Lisboa1,2,
Douglas A.G. Vieira1,2, Pedro M. Maia1
and Rodney R. Saldanha2

8.1 Introduction
The purpose of this chapter is to discuss the deterministic hydro generation sche-
duling (D-HGS) and a non-linear mathematical programming approach for it. To
obtain an efficient operation in a short or midterm horizon, i.e. a few days to a year,
hydropower plant characteristics are modeled in more detail at expense of inflow
uncertainties. In this context, non-linear programming provides significant gains [1].
Some important issues arise when mathematical programming is applied. The
formulation should include main physical aspects, such as an accurate power gen-
eration model and water flow conservation equations. On the other hand, from an
optimization standpoint, it is always preferable to define a convex formulation with
global optimality guarantees [2]. In this sense, these characteristics, especially the
hydropower generation efficiency, should be detailed towards a realistic model
while global optimization is still ensured.
Thus, many studies in power systems and optimization focused on D-HGS,
whose general form is
T X 
X 
max f vi;t ; ui;t ui;t
(8.1)
t¼1 i2I
subject to
" #
X 
vi;t ¼ vi;t1 þ gi;t  ui;t  si;t þ uj;t þ sj;t Dt ; 8i; t (8.2)
j2Wi

1
ENACOM Handcrafted Technologies, Rua Prof. José Vieira de Mendonça 770, 31310-260 Belo
Horizonte, Brazil
2
Graduate Program in Electrical Engineering, Federal University of Minas Gerais, Av. Antônio Carlos
6627, 31270-901 Belo Horizonte, Brazil
162 Modeling and dynamic behaviour of hydropower plants

v i  vi;t  v i ; 8i; t (8.3)


u i  ui;t  u i ; 8i; t (8.4)

si;t  0; 8i; t (8.5)

given a set of hydro plants I indexed by i and a time horizon of T time periods of
length Dt indexed by t, where v, u and s are the volume, discharge and spill vari-
ables, respectively. Volume lower v and upper n bound and discharge lower u and
upper u bound are defined in constraints (8.3) and (8.4). Water flow conservation is
set by constraint (8.2) given an inflow g. The inner summation is responsible for
addition of immediately upstream plants Wi outflow, in case of a hydropower
cascade, as shown in Figure 8.1. The initial volume of all reservoirs vi;0 ; 8i is
known. The objective aims to maximize hydropower generation (8.1) considering a
variable efficiency f. Efficiency is based on net head, the difference between
reservoir and tailrace height, which is dependent on volume and discharge variables
(in some cases also on spill variables). The head cannot be ignored in a storage
plant if there is a strong relationship between inflow and capacity.
This formulation can be extended by adding a demand and the minimization of
deficit or thermal complement. However, hydropower generation function (8.1)
determines how difficult the problem can be. If it is a concave function, then
D-HGS (and also the derivative problem) is convex. Unfortunately, concavity
property of hydropower generation function cannot be generally established [3].
Concerning the water-flow conservation equations, this constraint increases the
formulation complexity (it increases with number of periods and plants).

γ2t

γ1t

u2t + s2t

u1t + s1t

γ3t

u3t + s3t

Figure 8.1 Cascade with plants with reservoir (triangle) and without (circle).
Representation of flows in a time instant t: gi;t is the natural inflow
and ui;t þ si;t is the sum of the discharge and spill of plant i.
Hydro generation scheduling 163

Table 8.1 presents a non-exhaustive literature review focused on hydropower


generation function (Generation), water flow conservation equations (Water flow),
optimality guarantee and optimization method (Optimality).
Network flow algorithms exploit special structure of a hydropower cascade,
which various hydroelectric power plants are installed in same river basin (see
Figure 8.1) [4,5]. A second-order algorithm deal with non-differentiability in
hydropower functions by monitoring ‘‘break’’ points along line search [4]. Another
network algorithm, a simplex-type method, finds extreme local optimal solutions [5].
This method is applied after the linearization of the bilinear term in generation
function, net height times discharge, by a convex envelope. A compact formulation
has been solved by a conjugate gradient direction method by taking advantage of the
network structure of the linear system defined by water-flow conservation equations
considering the discharge (plus spill) as a dependent variable [3].
Sequential linear programming has no global optimality guarantee when
applied in an accurate and nonconvex D-HGS formulation [6], as well as, interior
point algorithm, which has to deal with an indefinite Hessian matrix using a
heuristic procedure [7]. These two works used fourth-order polynomial to fit
reservoir and tailrace height. An indefinite Hessian matrix is also defined by a
quadratic function of discharge and hydropower efficiency [8]. Due to non-con-
vexity, metaheuristics were applied to D-HGS, as Particle Swarm optimization [9].
A semidefinite programming relaxation is applied to a nonconvex quadratically
constrained quadratic formulation [10]. This relaxation is convex and, moreover,
authors prove the optimal solution obtained is that of the original problem. It is based on
quadratic form of hydro generation function and thermal complement minimization.
In its turn, Vieira et al. [12] achieved global optimality by a joint analysis of
physical and mathematical properties of hydropower plants. On the basis of an
efficiency function dependent on reservoir geometry, they have shown, under mild
conditions, that the resulting generation function is strongly increasing and pseudo-
concave. Then, a single plant D-HGS was established. A global optimal solution
can be found by traditional methods, e.g. ellipsoid algorithm.

Table 8.1 Literature review

Paper Generation Water flow Optimality


[3] Cont. first deriv. Linear system Local (Gradient)
[4] 2nd diff., disc. deriv. Equality Local (Network flow)
[5] Bilinear Equality Local (Network flow)
[6] Cont. first deriv. Equality Local (Seq. Linear Prog.)
[7] 2nd diff. Equality Local (Interior Point)
[8] Indef. quadratic Equality Local (Quadratic Prog.)
[9] Quadratic Equality Local (PSO)
[10] Quadratic Equality Global (Semidefinite Prog.)
[11] Bilinear Equality Global (B&B, envelopes)
[12] Pseudo-concave Inequality (v; u) Global (Ellipsoid alg.)
164 Modeling and dynamic behaviour of hydropower plants

A mixed integer non-linear programming formulation for cascades of hydro


plants, each one with multiple turbines, and a non-linear hydropower generation
function is solved by a spatial branch-and-bound (B&B) algorithm to address the
global optimum in [11]. Besides the binary variables associated with the status of
the turbines, the B&B principle is used as a partition scheme for the relaxation of the
bilinear terms with semi-continuous variables to provide a tight linear overestimation
of the non-convex function [13], i.e. a valid upper bound on the generation.
The next sections present an accurate D-HGS formulation for a cascade system
in order to obtain the global optimum guarantee based on [12]. The issues discussed
about bilinear functions [5,11] are also incorporated, as well as the transformation
of the water conservation equations in equivalent inequality constraints [14].
Firstly in this chapter, hydropower efficiency is analysed to establish an
accurate approximation with suitable mathematical properties. Then, cascade water
conservation modeling is detailed. As a result, a linearly constrained problem with
an increasing objective function is derived. It is shown this formulation belongs to
the class of jointly constrained biconvex problems under mild conditions, and a
B&B approach [15], which converges to a global optimal, is described and tested.

8.2 Hydropower generation function


Basically, hydropower plant transforms potential energy of water in electric power,
so plant efficiency comes from the water net head. This efficiency can be defined as

fðv; uÞ ¼ k½hðvÞ  oðuÞ (8.6)

where k is the plant productivity, which encompasses the acceleration of gravity,


the water density and the turbine and generator efficiency, hðvÞ is volume to
reservoir height function and oðuÞ is discharge (or output flow) to tailrace height
function. Naturally, this function can be more detailed [16] depending on applica-
tion, i.e. a day ahead scheduling or planning levels (short, mid and long term).
In planning level, the goal is to model net head with high accuracy. A widely
used approach is to fit these functions using a well-known non-linear function. For
example, Brazilian plants provide a fourth-order polynomial to represent volume to
reservoir height and discharge to tailrace height function [4,7,17]. Quadratic
function has also been used to approximate generation [8–10,18–20]. Although
these approaches may represent an accurate model, they are not able to ensure the
formulation convexity.
Some works have established a convexification procedure. In [21], a convex
hull of selected points of generation function is defined by a piece-wise linear
function. An initial approximation is defined by only considering volume and dis-
charge, and then, spillage effects is included. A single step procedure using a
convex hull algorithm was also developed [22]. But these methods do not consider
specific properties of each plant, setting a generic convex approximation that may
contain estimation errors.
Hydro generation scheduling 165

Thus, a detailed analysis on plant efficiency and its auxiliary functions, hðvÞ and
oðuÞ, enriches the approximation. In general, it is not possible to determine a unique
model for auxiliary functions without considering the plant physical properties.
Firstly, based on real-word operational restrictions and physical nature, domain
and codomain of all functions (hydro generation, efficiency and its auxiliaries)
could be characterized as
● domain is defined by the closed convex set D ¼ fðv; uÞj0 < v  v  v;
0 < u  u  ug, where volume v and discharge u values are both strictly
positive;
● codomain is also a subset of the strictly positive set, because the power
generation, the efficiency, reservoir and tailrace height are always strictly
positive.
From this basic premise, remaining properties are detailed.

8.2.1 Physical properties of geometric functions


Hydropower plant physical properties are fundamentally given by reservoir volume
to height function and tailrace flow to height function. In general, they are depen-
dent on the area of water surface, as these functions map volume (or output flow) to
height.
Firstly, based on general conditions of geometry, it is shown that both func-
tions are strongly increasing. Afterwards, the convexity or concavity condition is
defined based on the behaviour of the water surface area (decreasing, increasing or
constant). Then, some special reservoir and tailrace models are presented. For
simplicity, these components are modeled through standard forms.

8.2.1.1 Increasing property


A strongly increasing function is defined as

Definition 8.2.1 (Strongly increasing function [23, p. 8]). A function f : Rn ! R


is strongly increasing if for x1 and x2 2 Rn , x1;j  x2;j for all j ¼ 1; . . .; n and
x1;l < x2;l for some l implies f ðx1 Þ < f ðx2 Þ.
Firstly, volume to reservoir height function hðvÞ is analysed. Independently of
reservoir geometry, the basic assumption is the bigger the water volume is, the
bigger the reservoir height will be. This condition is adherent to real world because
the plant’s project defines efficiently the active portion of reservoir. Main irregu-
larities, e.g. unconnected valleys, are submerged below minimum operational
height. As a conclusion, function hðvÞ is strictly positive and strongly increasing,
therefore, v2 > v1 ) hðv2 Þ > hðv1 Þ [12].
Likewise, discharge to tailrace height function is also a strongly increasing and
strictly positive function in operational domain. In a given time, the bigger the
discharged flow into tailrace is, the bigger the water level immediately downstream
will be. So, u2 > u1 ) oðu2 Þ > oðu1 Þ. Even if spillage was considered, this func-
tion would still be strongly increasing.
166 Modeling and dynamic behaviour of hydropower plants

8.2.1.2 Convexity and concavity


Normally, water surface area varies with height h: the so-called water surface
function AðhÞ, as illustrate by a generic geometry in Figure 8.2.
This function is fundamental to characterize the relationship between height h
and volume v, since by definition:

ðh
v ¼ AðhÞ dh (8.7)
0

Given an infinitesimal change in height dh, a well-defined surface area is asso-


ciated and the associated volume variation dv is

dv ¼ AðhÞdh (8.8)

so, water surface function can be defined as

dv
AðhÞ ¼ (8.9)
dh
Regarding this function, three distinct situations may occur
● water surface increases with increasing height (volume), so, it is a strongly
increasing function;
● water surface decreases with increasing height (volume), so, it is a strongly
decreasing function; or,
● water surface is constant.
These situations impact on volume to height function hðvÞ according to the next
theorem, under differentiability assumption.

Theorem 8.2.1. The relationship between volume and height is increasing and
● concave, if the associated water surface area is increasing;
● convex, if the associated water surface area is decreasing;
● linear, if the associated water surface area is constant.

A(h) dv
dh

Figure 8.2 Generic geometry and the relationship between height h, volume v
and water area A
Hydro generation scheduling 167

Proof. If water surface function is increasing, its derivative is strictly positive [23],
and similarly, if water surface function is decreasing, its derivative is strictly
negative. The derivative of (8.9) with respect to height is

dAðhÞ d 2 v
¼ 2 (8.10)
dh dh
It is equivalent to the second derivative of volume in relation to height, vðhÞ. So, by
second-order conditions [2, p. 71], this function will be convex if water surface
function AðhÞ is increasing or concave if water surface function AðhÞ is decreasing.
Independently of reservoir geometry, this function is strongly increasing, because
greater the height, greater the water volume, as discussed in the previously section.
Function vðhÞ is the inverse of volume to height function hðvÞ. As both functions
are increasing by definition, convexity of one of them implies concavity of the other
[24]. Then, if volume to height function hðvÞ is concave, function vðhÞ is convex, i.e.
water surface function AðhÞ is increasing. And if volume to height function hðvÞ is
convex, function vðhÞ is concave, i.e. water surface function AðhÞ is decreasing.
In the last case, water surface area A is constant and function vðhÞ ¼ A h is linear,
so, its inverse (volume to height function hðvÞ) is also linear, hðvÞ ¼ ð1=AÞ v. &
Convexity or concavity property does not induce stationary points in the
volume to height function hðvÞ, because this function is increasing.
A similar theorem can be defined to discharge to height function oðuÞ
regarding tailrace water–surface area variation in relation to discharge flow.

8.2.2 Special cases of geometric functions


Figures 8.3–8.6 present four special geometries. In all cases, functions are
increasing. These forms illustrate Theorem 8.2.1. Normally, variation of height
with volume (discharge) is more pronounced at the bottom of reservoir (tailrace),
and it will continually being reduced due to increase in water surface area, i.e. an
increasing water surface function. This situation is approximated by continuous and
stair-like models with an associated concave function. In the stair-like case, function
could be fitted by a concave piecewise function. While a linear function is closely
Height

l
w Volume
(a) Geometry (b) Function example

Figure 8.3 Continuous case


168 Modeling and dynamic behaviour of hydropower plants

related to a uniform approximation of component, because surface is constant. An


irregular model induces a non-linear function neither convex or concave, because in
the region where water surface area decreases, function becomes convex, and where
it increases, function becomes concave.

Height
l

w1

w2 Volume
(a) Geometry (b) Function example

Figure 8.4 Stair-like case


Height

l
w Volume
(a) Geometry (b) Function example

Figure 8.5 Uniform case


Height

w Volume

(a) Geometry (b) Function example

Figure 8.6 Irregular case


Hydro generation scheduling 169

It is important to highlight some specific situations. For example a run-off-


the-river plant has a constant reservoir height, as its storage capacity is considered
null. Tailrace height could also be assumed constant because of topographical
characteristics or planning horizon length that result in very small tailrace height
variation.

8.2.2.1 Examples
Although these special cases do not fully contemplate variability of real geome-
tries, they are extremely useful to model in a simple and precise manner these
components. Official fourth-order polynomials used to fit auxiliary functions of
three Brazilian storage plants, Furnas, Emborcação and Sobradinho, are presented
in Figures 8.7–8.9. These polynomials are the most accurate approximation.
However, all reservoirs could be fitted by continuous model, as volume to reservoir

770

765
Reservoir height, m

760

755

750
0.5 1 1.5 2 2.5
Reservoir volume, hm3 × 104
(a) Furnas: reservoir height

673

672.8
Tailrace height, m

672.6

672.4

672.2

672

671.8

671.6
0 500 1,000 1,500 2,000
Discharge, m3/s
(b) Furnas: tailrace height

Figure 8.7 Auxiliary functions for Furnas hydropower plant


170 Modeling and dynamic behaviour of hydropower plants

670

660

Reservoir height, m
650

640

630

620

610
0.6 0.8 1 1.2 1.4 1.6
Reservoir volume, hm3 × 104
(a) Emborcao: reservoir height

524

523
Tailrace height, m

522

521

520

519
0 200 400 600 800 1,000 1,200
Discharge, m3/s
(b) Emborcao: tailrace height

Figure 8.8 Auxiliary functions for Emborcação hydropower plant

height is clearly a concave function. A simple concave piecewise linear function,


stair-like case, could represent these functions with a small loss of precision. This
approximation could be also used to fit tailrace height at Sobradinho plant. In a real
world plant, water surface area usually increases with increasing height while a
decreasing water surface, and consequently a convex function, is an unlikely phe-
nomenon. A uniform model is enough to describe the tailrace in the other two
plants. The most relevant advantage of these two models, uniform and stair-like, is
that both can be used in a linear programming framework.

8.2.3 Mathematical properties


This section establishes the relationship between physical conditions and mathe-
matical properties of generation function and also D-HGS problem. This analysis is
based on the assumption that spill effect in tailrace level could be ignored. In many
Hydro generation scheduling 171

394

392

Reservoir height, m
390

388

386

384

382

380
0.5 1 1.5 2 2.5 3 3.5
Reservoir volume, hm3 × 104
(a) Sobradinho: reservoir height

365

364
Tailrace height, m

363

362

361

360

359
0 1,000 2,000 3,000 4,000 5,000
Discharge, m3/s
(b) Sobradinho: tailrace height

Figure 8.9 Auxiliary functions for Sobradinho hydropower plant

plants, spillway does not direct water to tailrace, being located in another river bed
downstream location. Furthermore, due to operational constraints, such as river
flow and flood control, spillage occurs only in special situations.
Properties can be added to generation function when some physical models are
considered. In this analysis, tailrace and reservoir height will be considered con-
stant or variable. Each configuration is treated separately.
The case where both heights are constant defines a simple approximation with
efficiency constant and linear hydro-generation function. This approximation is
traditionally used in long-term planning. The next sections detail other cases.

8.2.3.1 Run-off-the-river plant with variable tailrace height


A run-off-the-river plant has a constant reservoir height ~h: In this case, efficiency and
generation are only functions of discharge. A condition to guarantee a concave hydro-
generation function, and consequently, a convex D-HGS formulation, is defined.
172 Modeling and dynamic behaviour of hydropower plants
 
Theorem 8.2.2. Let the efficiency function be fðuÞ ¼ k ~h  oðuÞ and the dis-
charge to tailrace height oðuÞ be strongly increasing and twice differentiable
function. Then, the hydropower generation function defined as gðuÞ ¼ fðuÞu is
concave in set D if:

d2o 2 do
2
 ; 8u 2 D (8.11)
du u du

Proof. A function in a convex domain is concave if and only if its Hessian is


negative semidefinite. For a single variable function, this reduces to simple con-
dition that second derivative must be non-positive [2, p. 71]. Set D is convex by
definition and the second derivative is
"   #
d2g d 2 ~hu d 2 ðoðuÞuÞ
¼k 
du2 du2 du2
d ðuðdo=duÞ þ oðuÞÞ (8.12)
¼ k
du
 2 
d o do
¼ k u 2 þ 2
du du
So, it is non-positive if the inner term is not negative as k > 0:

d2o do
u 2
þ2 0 (8.13)
du du
As tailrace height function oðuÞ is strongly increasing, its derivative is strictly
positive [23], and by definition u > 0, then, (8.13) is equivalent to

d2o 2 do
 (8.14)
du2 u du
&
For example uniform tailrace model, oðuÞ as a linear function, always respect this
condition.

8.2.3.2 Storage plant with constant tailrace height


This case was already studied by Vieira et al. [12]. The authors establish and prove
a condition to guarantee a pseudoconcave hydro generation function. Pseudo-
concave functions have the property that every local maximizer is also a global
maximizer over a convex set [25].
Given a constant tailrace height z; the efficiency function becomes a strongly
increasing function:

fðvÞ ¼ k½hðvÞ  z (8.15)


Hydro generation scheduling 173

in as much as volume to reservoir height function hðvÞ is strongly increasing. If it is


twice differentiable, the condition to a pseudoconcave generation function is

d2f 2 df 2
 ; 8v 2 D (8.16)
dv2 fðvÞ dv
where the proof can be found in [12].
If reservoir height function hðvÞ is concave, and consequently the efficiency
function, then the condition is satisfied, because second derivative is always non-
positive [2], and the right term is always positive since fðvÞ > 0.
A single plant D-HGS formulation with optimality guarantee is derived from
this condition [12].

8.2.3.3 Stair-like reservoir with a uniform tailrace


As seen in previous examples, tailrace height function oðuÞ could be fitted by a
linear function, uniform model:
oðuÞ ¼ au þ b (8.17)
and reservoir height function hðvÞ by a concave piece-wise function, stair-like model:

hðvÞ ¼ min cp v þ dp (8.18)


p

In this case, the generation function is nonconvex. However, the maximization of


generation function could be defined as

max k½h  oðuÞu ¼ k½h  ðau þ bÞu ¼ khu  kau2  kbu


subject to h  cp v þ d p ; 8p
(8.19)
hhh
u; v 2 D
where h is the reservoir volume variable with operational lower and upper bounds
h; h, and a; b are the coefficients of the linear tailrace height function oðuÞ. This
formulation is basically a static D-HGS without water conservation constraint, i.e. a
single time period and plant maximization problem.
In this formulation, hydro generation is written as a biconcave function, a
generalization of the bilinear concept.

Definition 8.2.2 (Bilinear and biconcave function). Let f ðx; yÞ be a continuous


function defined over a nonempty compact set S. If functions f ð; yÞ and f ðx; Þ are
both linear over S, then function f is bilinear. If functions f ð; yÞ and f ðx; Þ are
both concave functions over S, then function f is biconcave.
Although this class of problems is not concave, global optimization methods
[15,26,27] have been developed to guarantee optimality. This approach will be
detailed and applied to a complete D-HGS formulation in the last section.
174 Modeling and dynamic behaviour of hydropower plants

8.2.3.4 Storage plant with variable tailrace height


This is the most generic case. Unfortunately, it is unknown a form to establish a
concavity condition (restricted or extended). However, generation function can be
defined as a strongly increasing function. This condition does not exclude the exis-
tence of multiple local optimum, but it directs search to boundary of the feasible set,
if it is a convex, closed and nonempty set.
A strongly increasing function has all partial derivatives non-null (Definition
8.2.1), then, the sum of the gradients of the objective function and the active con-
straints in a particular solution will be null as required by the Karush–Kuhn–Tucker
conditions [2], if there is a set of active constraints to offset the gradient of the
objective function, i.e. the solution is in the boundary.
Proper conditions for the increasing property are established in the next theorem.

Theorem 8.2.3. Let efficiency function fðv; uÞ ¼ k½hðvÞ  oðuÞ be differentiable,


volume to reservoir height function hðvÞ and discharge to tailrace height function
oðuÞ be strongly increasing and differentiable functions. Then, hydropower gen-
eration function defined as gðv; uÞ ¼ fðv; uÞu is strongly increasing in set D if:
@oðuÞ
hðvÞ > oðuÞ þ u ; 8v; u 2 D (8.20)
@u

Proof. If a function is strongly increasing and differentiable, all of its partial derivatives
have to be strictly positive [23, p. 8]. The partial derivative with respect to volume is
@g @f @h
¼u ¼ uk (8.21)
@v @v @v
Set D is established by operational limits of each plant, i.e. only positive discharge
u > 0. As discussed in previous sections, function hðvÞ is strongly increasing in these
same operational limits, i.e. @h=@v > 0. So, partial derivative with respect to volume is
always strictly positive as k > 0 by definition, i.e. a real plant has a positive productivity.
Finally, partial derivative with respect to discharge is
@g @f
¼ fðu; vÞ þ u
@u @u  
@o
¼ k½hðvÞ  oðuÞ þ u k (8.22)
@u
 
@o
¼ k hðvÞ  oðuÞ  u
@u
All components are positive by definition, so this partial derivative is strictly
positive, and consequently generation function is strongly increasing, if:
@oðuÞ
hðvÞ > oðuÞ þ u ; 8v; u 2 D (8.23)
@u
&
Hydro generation scheduling 175

This condition can be simplified using the water head h0 ¼ hðvÞ  oðuÞ. In the real
world, this variable is always positive since the hydropower plant transforms
hydraulic potential energy, i.e. a positive water head, into electric power. So:

hðvÞ > oðuÞ; 8v; u 2 D (8.24)

holds true, and (8.20) can be rewritten as

@oðuÞ
h>u ; 8u 2 D (8.25)
@u
where h is the smaller water head defined by the plant project.

8.3 Water conservation and discharge limits


The increasing property holds true for the most generic case as stated in Theorem
8.2.3. In fact, this could be the unique common property of hydro-generation
function. As a result, if the formulation has only linear inequality constraints,
optimization methods could exploit the boundary of the feasible set. So, this section
demonstrates how to transform water balance equation (8.2) into inequalities within
the maximizing generation framework.
Discharge is established with only reservoir volume, natural inflow and dis-
charge upper limit [14], because hydropower cascade is an one-way direction
network flow. So, the maximization of an increasing hydro generation function
leads to the maximization of discharge.
For instance, in Figure 8.1, downstream flow of riverhead plants 1 and 2 at
period t is defined by (8.2) as
1 
u1;t þ s1;t ¼ v1;t1  v1;t þ g1;t (8.26)
Dt
1 
u2;t þ s2;t ¼ v2;t1  v2;t þ g2;t (8.27)
Dt
The immediately downstream plant is number 3 for both plants. In this run-off-the-
river plant, water conservation equation is
u3;t þ s3;t ¼ u1;t þ s1;t þ u2;t þ s2;t þ g3;t (8.28)

So, substituting (8.26) and (8.27) into (8.28),


1  1 
u3;t þ s3;t ¼ v1;t1  v1;t þ g1;t þ v2;t1  v2;t þ g2;t þ g3;t (8.29)
Dt Dt
In summary, downstream flow of a plant i at period t is defined as
X1  X
ui;t þ si;t ¼ vk;t1  vk;t þ gj;t (8.30)
D
k2L t j2Y
i i
176 Modeling and dynamic behaviour of hydropower plants

where set Li contains plant i, if it has reservoir, and all upstream plants with
reservoir, and set Yi contains plant i and all upstream plants.
In the previous section, it was established an assumption that spill does not
interfere in hydro generation function. Moreover, this analysis assumes that gen-
eration is an increasing function, then discharge could be defined as:
( )
X1  X
ui;t ¼ min vk;t1  vk;t þ gj;t ; u i (8.31)
D
k2L t j2Y
i i

which is equivalent to inequalities constraints


X1  X
ui;t  vk;t1  vk;t þ gj;t ; 8i; t
D
k2L t j2Y
i i
(8.32)
0  ui;t  u i;t ; 8i; t
v i  vi;t  v i;t ; 8i; t

8.3.1 Head sensitive discharge limits


Maximum water discharge can be considered a function of the water head [8]. This
represents an important physical characteristic related to plant efficiency. In some
storage plant, maximum water discharge may be different for each reservoir level
according to head. So, discharge equation (8.31) could be rewritten:
( )
X1  X  
ui;t ¼ min vk;t1  vk;t þ gj;t ; u hi;t (8.33)
D
k2L t j2Y
i i

where hi;t is the reservoir height, and u i ðhi;t Þ is the maximal discharge function.
A hydroelectric plant has a nominal water head, where nominal power is
produced if discharge is maximal, u i . The real discharge limit is less than this
maximum, otherwise. If head is lower than nominal, reduction is caused by turbine
limitations, and if head is bigger, the cause is the generator. Bellow nominal height,
discharge limit increases, and above it, it decreases [17].
Despite this function has a non-linear nature, it could be fitted by a concave
piecewise function to maintain all constraints linear:
ui;t  ri;1 hi;t þ ei;1 ; 8i; t (8.34)

ui;t  ri;2 hi;t þ ei;2 ; 8i; t (8.35)

where r and e are slope and constant in each piece. A function with two pieces can
be defined considering a given nominal head. Slope is positive in first range
ri;1 > 0 and negative in second one ri;2  0.
This constraint set with (8.32) defines discharge feasible values only with
linear inequalities, i.e. the lower value between water flow and discharge limit.
These constraints represent the complete discharge equation (8.33).
Hydro generation scheduling 177

8.4 Cascade D-HGS formulation


The problem (8.1)(8.5) induces a linearly constrained formulation with a non-
concave objective function. However, this formulation can evolve in the mathematical
and physical sense of the problem to retain optimality conditions based on the
mathematical framework proposed in this work.
Firstly, it is assumed that all generation functions are increasing, i.e. all
hydropower plants respect Theorem 8.2.3. So, the objective function is also
increasing (summation operator). Moreover, all constraints are linear inequali-
ties, i.e. each one defines a closed half-space Hk ¼ x 2 Rn jaTk x  bk , and the
T
intersection set Kk¼1 Hk (polyhedron) is a convex set [2]. Thus, the Karush–
Kuhn–Tucker conditions [2] state that the optimal solution belongs to boundary
of the polyhedron.
As concavity is not ensured, new properties must be added to the geometric
functions, and consequently, to the generation function to guarantee the con-
vergence to the global optimal boundary solution. The reservoir and tailrace in the
cascade are modeled by the standard form presented in the previous section. The
stair-like reservoir and a uniform tailrace can be an accurate approximation.
A stair-like reservoir with an increasing water surface area induces a concave
piecewise linear volume to height function, whereas the uniform tailrace is repre-
sented by a linear function. Then, a biconcave formulation is defined as
X
T X   X
T X  
max f ¼ f vi;t ; ui;t ui;t ¼ ki hi;t  ai ui;t  bi ui;t (8.36)
t¼1 i2I t¼1 i2I

subject to

hi;t  ci;p vi;t  di;p ; 8i; pi ; t (8.37)

ui;t  ri;1 hi;t  ei;1 ; 8i; t (8.38)

ui;t  ri;2 hi;t  ei;2 ; 8i; t (8.39)


X1  X
ui;t  vk;t1  vk;t  gj;t ; 8i; t (8.40)
D
k2L t j2Y
i i

hhh (8.41)

vvv (8.42)

uuu (8.43)

to maximize the total hydropower generation (8.36). Although it is not a general


formulation as (8.1)(8.5), it provides a realistic modeling of the reservoir and the
generating function; at a same time, it is solved to optimality by a global optimi-
zation approach [15].
178 Modeling and dynamic behaviour of hydropower plants

Reservoir stair-like model is defined by a concave piecewise linear function


(8.37). Another concave piecewise function is set by (8.38) and (8.39). These
constraints are related to head sensitive discharge limits. Water conservation
inequalities (8.40) and variables limits (8.41)–(8.43) complete the set of constraints.
The initial volume Vi;0 is known for all reservoirs.

8.5 Global optimization approach


The formulation can be solved to optimality if all plants have a constant or a linear
tailrace height. In the first case, generation function becomes bilinear:
khu  kbu (8.44)
where b is the constant tailrace level, considering a ¼ 0. Otherwise, function is
biconcave as previous discussed in mathematical analysis, see system (8.19). The
cascade could also have run-off-the-river plants, because these plants will induce a
concave function (see Theorem 8.2.2).
Thus, non-convexity is related to the bilinear term h u. The procedure to handle
this non-linearity is based on concave envelope and B&B algorithm [15]. The first
one sets a convex relaxation, while B&B refines it. n o
Concave envelope of a ¼ h u over set W ¼ h  h  h; u  u  u is
defined as
n o
min uh þ ðu  uÞh; hu þ ðh  hÞu (8.45)

which overestimates a over W and agrees with a on @W (boundary). This function


could be written as a concave piecewise linear function since objective is to max-
imize an increasing generation function. So, this relaxation becomes a convex
formulation. For example, the relaxation of a biconcave formulation (8.19) is

max f ¼ ka  kau2  kbu


subject to a  uh þ ðu  uÞh
a  hu þ ðh  hÞu
h  cp v þ d p ; 8p
hhh
u; v 2 D
uh  a  uh (8.46)

which is a linearly constrained formulation with a concave quadratic objective


function. This relaxation is generalized to the case of multiple periods and hydro-
power plants, and the introduction of the water balance equations does not change
any mathematical properties. So, formulation (8.36)–(8.43) has a similar relaxation
(linearly constrained with a concave quadratic objective function).
Hydro generation scheduling 179

Relaxed objective function value is always bigger or equal the original func-
tion value due concave envelope. Precisely, these values are equal when any
variables h or u is at its limit, i.e. a point in boundary @W. So, the approach aims at
adjusting (by reducing) the set W in order to best evaluate the original function
through relaxation problem.
Then, B&B establishes the search tree through partition of W. Given the
optimal solution ðh ; u Þ of the relaxation problem, the set W is divided into
four parts
1. W1 ¼ fh  h  h ; u  u  u g
n o
2. W2 ¼ h  h  h; u  u  u
n o
3. W3 ¼ h  h  h; u  u  u
n o
4. W4 ¼ h  h  h ; u  u  u
and limits are updated. A relaxation problem is associated to each of these subsets
k ¼ 1; 2; . . . and, then, they are optimized and the partition procedure is repeated.
At each step, the (best) relaxation becomes closer to the original function, since
solutions are closer to boundary of these new narrow sub-sets Wk . Note that
objective function is increasing, so, optimal solutions of all relaxation problems are
at the border of their feasible set. Same rules used in the traditional B&B method,
applied to a mixed integer problem, are sufficient to achieve global optimum
[15,26]. Pseudo-code 1 details this algorithm.
In the first line, the concave envelope relaxation over the original set E is
solved. The optimal solution of this relaxation is the current optimal solution
(line 2). Then, the relaxed function value is saved as upper bound (line 3) and
the original function value as lower bound (line 4). These limits are important
because the difference between them is the convergence criterion. In the fifth
step, the feasible set E is divided into four subsets as previously described.
A list is created with these four subsets. This list represents the subproblems
that will be solved in order to refining the relaxation, i.e. the search tree in the
B&B algorithm. The refining procedure is detailed in the loop from lines 6
to 19. There are two convergence criteria: (i) no more subproblems (empty List) or
(ii) the lower and upper bound converged numerically considering an tolerance e.
At each iteration, a subproblem in the List is selected and solved (lines 7
and 8). If the current original function value is greater than the lower bound,
the lower bound and the optimal solution are updated (lines 9–12). This new
solution is the new power generation maximum. If the relaxed function value is
greater than the lower bound, then, current feasible set Ek is split into four
subsets and stored in the List (line 14), i.e. a new branch is added because the
relaxation can be improved. Moreover, if this relaxed function value is less
than the upper bound, this limit is updated (lines 15–17). Note that the upper
bound is the best relaxed function value, i.e. the value closer to the maximal
original function value.
180 Modeling and dynamic behaviour of hydropower plants

Algorithm 1 Branch-and-bound algorithm based on concave envelope relaxation


for D-HGS.

Input: fðv; uÞ, W, D-HGS formulation, e


Output: u ; v ; h , f
1: Solve the concave envelope relaxation of D-HGS formulation, u0 ; v0 ; h0 ; f 0
2: u ¼ u0 ; v ¼ v0 ; h ¼ h0 ⊳ Current optimal solution
3: f ¼ f 0
⊳ Upper bound: relaxed function
4: f ¼ fðv0 ; u0 Þu0 ⊳ Lower bound: original function
5: Split set W into four subsets W1 ; . . .; W4 and store in List
6: While List not empty & f  f > e do
7: Select a subproblem k
8: Solve (relaxed) subproblem k, uk ; vk ; hk ; f k
 
9: If f vk ; uk uk > f
 k k k
10: f ¼ f v ;u u ⊳ Improve lower bound
 k  k 
11: u ¼ u ;v ¼ v ;h ¼ h k
⊳ Current optimal solution
12: end If
13: If f k > f
14: Split set Wk into four subsets and store in List
15: If f k < f
16: f ¼ fk ⊳ Improve upper bound
17: end If
18: end If
19: end While

8.5.1 Computational results


A Brazilian cascade was selected to test this method. The instance has 12 plants
installed in Paranaı́ba, Araguari, Corumbá and São Marcos rivers. Seven plants
have reservoir, as shown in Figure 8.10. Hydropower plants technical information
as well as historical natural inflows are available [28]. The global optimization
method was implemented in MATLAB and the quadratic formulations (relaxed
subproblems) were solved by a second-order cone programming solver [29]. The
computer used was an Intel 1366 XEON, 2.4 GHz and 16 GB RAM.
Monthly inflow observed in two recent years, 2010 and 2014, were used
because of differences in inflow patterns as shown in Figure 8.11. Initial reservoir
Hydro generation scheduling 181

Nova Ponte

Miranda
Corumbá IV

C. Branco I Corumbá III


Serra
Facão

C. Branco II Corumbá I
Emborcação

Itumbiara

C. Dourada

São Simão

Figure 8.10 Paranaiba river basin cascade, Brazil

450
2014
400
2010
350
Inflow at Nova Ponte, hm3

300

250

200

150

100

50

0
Feb Apr Jun Aug Oct Dec
Month

Figure 8.11 Inflow at Nova Ponte hydroelectric plant, Brazil


182 Modeling and dynamic behaviour of hydropower plants

10,000
2014
8,000 2010

Time, s 6,000

4,000

2,000

0
3 4 5 6 7 8
Number of time periods

Figure 8.12 Runtime for Araguari river instance

volume are also set differently. They are defined based on the observed value at the
beginning of each year. For example Nova Ponte reservoir starts 2010 with 77%,
whereas in 2014, initial volume was 35%. In general, year 2010 has a bigger initial
volume in all reservoirs. The algorithm was run in six different horizon settings,
from 3 to 8 months starting in January.
Firstly, it is considered just plants installed in Araguari river: Nova Ponte,
Miranda, C. Branco I and C. Branco II. Figure 8.12 presents the average running
time for each inflow case. A gap e smaller than 106 was set as convergence
criterion. Considering this tight numerical criterion, it was observed an expo-
nential time complexity in the 2014 case, as expected in a worst case for a B&B
algorithm. This behaviour is not repeated for 2010. Clearly, computational
complexity is strongly influenced by initial volume and inflow pattern, since
the relaxation is worst when reservoir height is lower. For example, in the
eight-month instance, the initial relaxation gap was 0.5% for 2010 and 4.7%
for 2014. A bigger gap increases the number of B&B iterations, and so, the
runtime.
Runtime for the complete instance is presented in Figure 8.13. A maximal
number of iterations was added due to bigger instance complexity (12 plants
instead of 4). In all cases, the limit of 100,000 iterations was reached. This new
criteria prevented exponential behaviour and approximate the runtime in both
cases. However, the final relaxation gap in 2014 is bigger than 2010 as shown in
Figure 8.14, so the runtime to numerical convergence for 2014 is expected to be
bigger than 2010 case.
Runtime could be improved by refining the feasible bounds of variables h and
u [5] and by applying a parallel B&B version.
Hydro generation scheduling 183

× 104
2.5
2010
2014
2

1.5
Time, s

0.5

0
3 4 5 6 7 8
Number of time periods

Figure 8.13 Runtime for Paranaı́ba river basin instance

0.7
2010
0.6
2014
Final relaxation gap, %

0.5

0.4

0.3

0.2

0.1

0
3 4 5 6 7 8
Number of time periods

Figure 8.14 Final relaxation gap for Paranaı́ba river basin instance

8.6 Conclusions
A mathematical analysis based on hydroelectric plant characteristics, i.e. reservoir
and tailrace geometry, is the central chapter’s subject. Deterministic mid-short term
scheduling planning could be solved to optimality if generation function is at
least biconcave, i.e. a discontinuous reservoir and a uniform tailrace. Although the
184 Modeling and dynamic behaviour of hydropower plants

formulation is not proved to be unimodal, a B&B method based on concave envelope


relaxation converges to global optimum. In a broader sense, increasing property is
established for hydro-generation function. As the formulation is linearly constrained,
i.e. only linear inequalities, special algorithms could exploit boundary of the feasible
set (a polytope) to find a global optimum. Unimodularity property could also be
studied in general case, i.e. arbitrary reservoir and tailrace height.

References
[1] Martins LSA, Azevedo AT, Soares S. Nonlinear Medium-Term Hydro-
Thermal Scheduling With Transmission Constraints. IEEE Transactions on
Power Systems. 2014;29(4 (July)):1623–1633.
[2] Boyd S, Vandenberghe L. Convex Optimization. New York, NY: Cambridge
University Press; 2004.
[3] Sylla C. A subgradient-based optimization for reservoirs system management.
European Journal of Operational Research. 1994; 76(1): 28–48.
[4] Oliveira GG, Soares S. A second order network flow algorithm for hydro-
thermal scheduling. IEEE Transactions on Power Systems. 1995; 10(3):
1635–1641.
[5] Feltenmark S, Lindberg PO. Network Methods for Head-dependent Hydro
Power Scheduling. In: Pardalos P, Hearn D, Hager W, editors. Network
Optimization, vol. 450 of Lecture Notes in Economics and Mathematical
Systems. Springer: Berlin; 1997. p. 249–264.
[6] Barros MTL, Tsai FTC, Yang S-l, Lopes JEG, Yeh WWG. Optimization of
large-scale hydropower system operations. Journal of Water Resources
Planning and Management. 2003; 129(3): 178–188.
[7] Azevedo AT, Oliveira ARL, Soares S. Interior point method for long-term
generation scheduling of large-scale hydrothermal systems. Annals of
Operations Research. 2009; 169(1): 55–80.
[8] Catalao JPS, Mariano SJPS, Mendes VMF, Ferreira LAFM. Scheduling of
head-sensitive cascaded hydro systems: a nonlinear approach. IEEE Trans-
actions on Power Systems. 2009; 24(1): 337–346.
[9] Mahor A, Rangnekar S. Short term generation scheduling of cascaded hydro
electric system using novel self adaptive inertia weight PSO. International
Journal of Electrical Power & Energy Systems. 2012; 34(1): 1–9.
[10] Zhu Y, Jian J, Wu J, Yang L. Global optimization of non-convex hydro-
thermal coordination based on semidefinite programming. IEEE Transactions
on Power Systems. 2013; 28(4 (Nov)):3720–3728.
[11] Lima RM, Marcovecchio MG, Novais AQ, Grossmann IE. On the com-
putational studies of deterministic global optimization of head dependent
short-term hydro scheduling. IEEE Transactions on Power Systems. 2013;
28(4): 4336–4347.
[12] Vieira DAG, Guedes LSM, Lisboa AC, Saldanha RR. Formulations for
hydroelectric energy production with optimality conditions. Energy Conversion
and Management. 2015; 89(1): 781–788.
Hydro generation scheduling 185

[13] McCormick G. Computability of global solutions to factorable nonconvex


programs: Part I – Convex underestimating problems. Mathematical Pro-
gramming. 1976; 10(1): 147–175.
[14] Guedes LSM, Vieira DAG, Lisboa AC, Saldanha RR. A continuous compact
model for cascaded hydro-power generation and preventive maintenance
scheduling. International Journal of Electrical Power & Energy Systems.
2015; 73(1): 702–710.
[15] Al-Khayyal FA, Falk JE. Jointly constrained biconvex programming.
Mathematics of Operations Research. 1983; 8(2): 273–286.
[16] Cordova MM, Finardi EC, Ribas FAC, de Matos VL, Scuzziato MR. Perfor-
mance evaluation and energy production optimization in the real-time operation
of hydropower plants. Electric Power Systems Research. 2014; 116(1): 201–207.
[17] Hidalgo I, Fontane D, Soares FS, Cicogna M, Lopes J. Data consolidation
from hydroelectric plants. Journal of Energy Engineering. 2010; 136(3): 87–94.
[18] Soares S, Lyra C, Tavares H. Optimal generation scheduling of hydro-
thermal power systems. IEEE Transactions on Power Apparatus and Systems.
1980; PAS-99(3 (May)):1107–1118.
[19] Naresh R, Sharma J. Two-phase neural network based solution technique for
short term hydrothermal scheduling. IEE Proceedings-Generation, Transmis-
sion and Distribution. 1999; 146(6 (Nov)): 657–663.
[20] Yuan X, Wang Y, Xie J, Qi X, Nie H, Su A. Optimal self-scheduling of
hydro producer in the electricity market. Energy Conversion and Management.
2010; 51(12): 2523–2530.
[21] Diniz AL, Maceira MEP. A four-dimensional model of hydro generation for
the short-term hydrothermal dispatch problem considering head and spillage
effects. IEEE Transactions on Power Systems. 2008; 23(3): 1298–1308.
[22] Ramos TP, Marcato ALM, da Silva Brandi RB, Dias BH, da Silva Junior IC.
Comparison between piecewise linear and non-linear approximations
applied to the disaggregation of hydraulic generation in long-term operation
planning. International Journal of Electrical Power & Energy Systems. 2015;
71(1): 364–372.
[23] Miettinen KM. Nonlinear Multiobjective Optimization. Boston, MA: Kluwer
Academic Publishers; 1999.
[24] Mršević M. Convexity of the inverse function. The Teaching of Mathe-
matics. 2008; XI(1): 21–24.
[25] Diewert WE, Avriel M, Zang I. Nine kinds of quasiconcavity and concavity.
Journal of Economic Theory. 1981;25(3):397–420.
[26] Al-Khayyal FA. Jointly constrained bilinear programs and related problems:
an overview. Computers & Mathematics with Applications. 1990;19(11):53–62.
[27] Bloemhof-Ruwaard JM, Hendrix EMT. Generalized bilinear programming:
an application in farm management. European Journal of Operational
Research. 1996; 90(1): 102–114.
[28] Brazilian Chamber of Electric Energy Commercialization. NW201506.zip;
2015. Available from http://www.ccee.org.br/ccee/documentos/NW201506.
[29] Gurobi Optimization, Inc. Gurobi Optimizer Reference Manual; 2015. Available
from http://www.gurobi.com.
Chapter 9
A PV hydro hybrid system using residual
flow of Guarita Hydro Power Plant,
in southern Brazil
Rafael Schultz1, Alexandre Beluco2, Roberto Petry
Homrich1 and Ricardo C. Eifler3

Abstract
The current situation of depletion of energy resources and population growth makes
feasible the use of the remaining potential in power plants already built and in
operation. A recent research trend is the study of performance of photovoltaic–
hydroelectric hybrid energy systems with PV modules installed on the reservoir
surface. The hydroelectric power plant of Guarita was launched in 1953 and has an
installed capacity of 1.86 MW, with 40 m height. This study assesses the feasibility
of utilization of the residual flow of 370 l/s in a machinery house placed 12 m
below the water level of the dam, operating in conjunction with PV modules
installed over the surface of the water reservoir. The study was conducted based on
simulations with well-known HOMER software, Legacy version. The hydroelectric
potential will be deployed with low-cost alternatives such as the use of centrifugal
pumps as turbines. The photovoltaic potential will be exploited with modules
installed on floating structures. The optimization space obtained with HOMER
indicates feasible solutions with the combination of 34.8 kW hydro and 30 kW PV.
Lower costs of PV modules can make viable some solutions with 60 and 90 kW PV.
This work also indicates useful conclusions in the design process and implementa-
tion of the hybrid system under study.

Keywords

Hybrid system, PV hydro hybrid system, PV modules on floating structures,


residual flow, southern Brazil, computational simulation, software Homer

1
Escola de Engenharia, Universidade Federal do Rio Grande do Sul (UFRGS), Porto Alegre, RS, Brazil
2
Instituto de Pesquisas Hidráulicas, Universidade Federal do Rio Grande do Sul (UFRGS), Porto Alegre,
RS, Brazil
3
Companhia Estadual de Energia Elétrica, Salto do Jacuı́, RS, Brazil
188 Modeling and dynamic behaviour of hydropower plants

9.1 Introduction
A global overview of economic crisis is configured for some years, worsened by
the steady increase in the consumption of fossil fuels and the increasing demand for
energy supplies. The increase in energy demand is a quantitative increase, but there
is also a growing demand for better quality of power supplies. In this scenario, it is
important to increase the availability of energy supplies, either through new plants
or through repowering of old plants.
In hydroelectric plants in which the power house is not located near the dam
base, there is a piece of river with reduced flow. There are studies and laws that
establish suggested values and minimum values for these reduced flows. However,
even accounting small flows, especially when compared to the power provided by
the plant to which they belong, these reduced flows may represent the availability
of important energy supplies.
A new plant to be implemented to generate energy from the ecological flow
should be planned within the current legislation. This new plant should not lead to
pieces of river without any flow. Thus, it is likely that a small dam is built a little
ahead of the main dam, and the new powerhouse is located next to the new dam.
Obviously, it will be a run of the river plant possibly generating energy also from
the water that is flowing through the main dam spillway.
A hybrid system consisting of a hydroelectric power plant and a photo-
voltaic system may simply represent an increase in installed capacity for power
supply, but can also represent a better use of available energy if there is a
possibility of energy storage (by means of batteries or water reservoir or other
methods). Moreover, energy can be managed according to the energetic com-
plementarity [1] of the existing energy resources on the site where the system is
implemented.
This chapter presents a prefeasibility study on the use of ecological flow in
Guarita Hydro Power Plant, in southern Brazil. An additional small hydropower
plant is proposed for power generation, constituting a hybrid system with photo-
voltaic modules installed on floating structures on the surface of the small reservoir
formed for this purpose. The next section presents the Guarita power plant, and the
following section explains the use of ecological flow. Then, the components of the
hybrid system are presented, and simulations with HOMER software are explained.
Finally, results are presented and discussed.

9.2 The Guarita hydroelectric power plant


Guarita Hydro Power Plant is located on the river Guarita, which belongs to the
basin of the river Uruguay. The location of the plant has the coordinates
27 360 24.700 South and 53 340 26.600 West. The dam lies on the border between the
municipalities of Erval Seco [2] and Redentora [3], in the State of Rio Grande do
Sul [4], in southern Brazil, with the powerhouse located in ErvalSeco. The dam is
at a distant of approximately 440 km from Porto Alegre, the capital of the state.
Use of residual flow for a PV hydro hybrid system 189

The Guarita power plant belongs to the Division of Generation and Transmission of
the State Company for Electric Power and became operational in 1953.
The steady flow is 6.15 m3/s, surpassed or equaled 95% of the time, the
average monthly minimum flow is 3.03 m3/s, and the average water discharge is
5.78 m3/s. The value of the ecological flow is not provided by the company
responsible for the plant, but for this work, it has been calculated as the difference
between the steady flow and the average water discharge, resulting in 0.37 m3/s.
The area contributing to the reservoir is 829 km2, and the total extension of
data used for the determination of flow rates is from 1964 to 2009. This ecological
flow rate is considered in this work for power generation purposes in a supple-
mentary plant.
The dam is built in cyclopic concrete, with a crest length of 100 m. Figures 9.1
and 9.2 show, respectively, upstream and downstream views of the dam. The
maximum height of the dam on the ground level is 7 m, but the height in the region
of the spillway is 4 m. At the reservoir, the water level is always between the
normal minimum of 408.548 m and the normal maximum of 410.048 m with
exceptional maximum equal to 413.048 m. The area flooded by the dam in the
normal maximum level is 0.0287 km2. Below the reservoir, the normal maximum
level is 366.596 m, and the exceptional maximum level is 368.688 m at the
powerhouse.
An intake tunnel with a length of 960 m and diameter of 2.55 m carries the water
to the nearest point of the powerhouse. The standpipe has a diameter of 8 m and a
height of 18.75 m, and the penstock has a diameter of 1.7 m and length of 76 m.
A single Francis turbine is used with 2 MW and nominal height of 43.45 m and
nominal flow rate of 5.78 m3/s. The generator unit has 2.2 MVA with a power factor
of 0.8 and terminal voltage of 8 kV. The rated speed of the turbine and the generator
is 450 rpm. Figure 9.3 shows the hydraulic turbine and synchronous generator used

Figure 9.1 Upstream view of the dam


190 Modeling and dynamic behaviour of hydropower plants

Figure 9.2 Downstream view of the dam

Figure 9.3 Hydraulic turbine and electric machine at the powerhouse

in the Guarita Plant. The connection with the energy system uses a 23 kV. The plant
has an installed capacity of 1.76 MW with steady capacity of 1.10 MW.

9.3 The use of residual flow of Guarita


This paper proposes the implementation of a new powerhouse that will not influ-
ence the operation of the main plant described in the previous section. There is a set
of natural falls along the river, shown in Figure 9.4, a little below the dam shown in
Use of residual flow for a PV hydro hybrid system 191

Figure 9.4 Natural waterfall along the river

Figures 9.1 and 9.2. A new dam located in a position below these waterfalls will
ensure a gross height of 12 m to generate power with the ecological flow identified
in a preliminary way in the previous section. This new dam should fully exploit the
height difference between its position and the dam of the Guarita power plant, thus
ensuring that environmental legislation will not be infringed.
Figure 9.5 shows a satellite image [5] of the area of the Guarita power plant
with its main elements identified. The dam is located at the bottom of the figure,
whereas the powerhouse is located at the top. The reduced flow path extends
between the dam and the powerhouse that is where the water returns to the natural
course of the river. This figure also shows the location of the new dam, proposed in
this work, a little ahead of the position of the existing dam. The portions of the river
between the new dam and existing dam will be completely flooded, forming a small
reservoir, however without any usable energy storage capacity.
The proposed hydroelectric power plant will be designed with low costs and
equipment with large-scale production. The power house will be made with cen-
trifugal pumps used as power turbines and induction motors used as self-excited
asynchronous generators. The compact configuration for the proposed plant and the
proximity of its components, in addition to the small volume necessary for the
small dam, ensure that energy is generated with very low costs.
The new plant will have a steady capacity of 34.8 kW.

9.4 Components of the PV hydro hybrid system


The hybrid system to be constituted will then have the hydroelectric power plant
described in the previous section, and a set of PV modules placed on floating
structures installed on the free surface of the small reservoir will be formed as a
192 Modeling and dynamic behaviour of hydropower plants

Power house

Part of the river


with ecological flow
Rio
Gua
rita

Natural waterfalls

Dam

Google

Figure 9.5 Location of main elements of Guarita Hydro Power Plant

result of the proposed small dam located as shown in Figure 9.5. These two gen-
erator sets will be connected to the national grid and be liable for an electric charge
with a certain value. Figure 9.6 shows a schematic diagram of the system. All
components are connected by the AC bus that describe how the hydro power plant
and the PV modules will be connected to the grid and responsible for a given
electrical charge.
The PV module assembly will be installed on floating structures, as recently
suggested by Ferrer-Gisbert et al. [6] and Redon-Santafé et al. [7]. They propose
and test a system with polyethylene floating modules that occupy an area that
would not be used in a better way and that also contribute reducing evaporation.
This study has not yet detailed the floating structures, having been restricted so
far to the economic feasibility. The basic floating structure considered in this
study has dimensions suitable for 30 kW in PV modules. The water surface
formed with the dam is small but sufficient for several tens of structures having
these dimensions.
Use of residual flow for a PV hydro hybrid system 193

Figure 9.6 Schematic drawing of the PV hydro hybrid system

9.5 Simulations with HOMER


HOMER [8] is a software for optimization of micro and small-hybrid energy sys-
tems. It was originally developed by National Renewable Energy Laboratory and is
available for universal access in its version called ‘‘Legacy.’’ HOMER simulates a
system for power generation over the time period of 25 years at intervals of 60 min,
presenting the results for a period of 1 year [9,10]. The HOMER software performs
simulations of the hybrid system aiming to build a space of optimal solutions, while
still allowing to gather these optimization spaces according to sensitivity analyses.
The cost of the hydro power plant should be assessed more accurately in the
next stages of the project. The powerhouse will be built with pumps used as tur-
bines, for cost reduction, and the overall cost was estimated a US$ 69,600 (con-
sistent with IRENA [11]), with a replacement cost of US$ 55,680 and annual
operation and maintenance costs of US$ 2,784. The turbine can operate between
50% and 150% of the design flow rate, and the efficiency of the machines is 80%.
A loss of about 3% is estimated for the hydro power plant intake system. Figure 9.7
shows the parameters of the hydroelectric power plant, and Figure 9.8 shows the
available energy.
The HOMER software simulates hydroelectric power plants operating as ‘‘run
of the river’’ power plants. In these plants, the flow that is used to generate energy
is at the maximum equal to the flow of the river by the lack of a reservoir with
a reasonable storage capacity. The dam planned for this new plant will have a
194 Modeling and dynamic behaviour of hydropower plants

Figure 9.7 Parameters of the hydroelectric power plant proposed in this work

maximum level equal to the output level of the existing dam, where there is the
release of ecological flow rate. Thus, there will be no stretch of the river without
the ecological minimum flow rate, even with the installation of this new dam and
power generation from residual flow.
The cost of the PV modules was US$ 4,380/kW, and it is compatible with
usual costs found, for example, by Feldman et al. [12]. The installation of floating
structures, as suggested by Ferrer-Gisbert et al. [6] and Redon-Santafé et al. [7],
increases the cost by 30%. The lifetime of the PV system is considered to be
12.5 years, the replacement cost of the PV system at the end of the useful life is
80% of the initial cost and annual cost of operation, and maintenance is 5% of the
installation cost. The reflectance of the water surface is considered to be 10% at the
installation site. Figure 9.9 shows parameters of the PV modules, and Figure 9.10
shows the available solar energy.
Simulations with the system of Figure 9.6, with the PV modules assembled
on floating structures installed over the flooded surface of the reservoir, were
performed. A set of 9,984 simulations, with 64 combinations for optimization
analysis and 15 combinations for sensitivity analysis. Among the results, 3,444
were feasible solutions and 6,540 were unfeasible (2,400 due to the capacity
Use of residual flow for a PV hydro hybrid system 195

Figure 9.8 Hydro energy considered in this work

Figure 9.9 Parameters of the PV modules


196 Modeling and dynamic behaviour of hydropower plants

Figure 9.10 Solar energy availability considered in this work

shortage constraint and 4,140 to the renewable fraction constraint). The optimiza-
tion variables considered were 0, 30, 60, 90, 120, 150, 180, and 210 kW for PV
array capacity and 0, 50, 100, and 150 kW for grid purchases.
The sensitivity inputs were 725 kWh/d, 750 kWh/d, 775 kWh/d, 800 kWh/d,
825 kWh/d, 850 kWh/d, 875 kWh/d, 900 kWh/d and 925 kWh/d for AC load; 0.50,
0.75, 1.00 and 1.30 for PV capital cost multiplier, PV replacement cost multiplier,
PV operation, and maintenance cost multiplier, these three linked; US$ 69,600,
US$ 52,200 and US$ 34,800 for hydro capital cost; US$ 55,680, US$ 41,760 and
US$ 27,840 for hydro replacement cost; and US$ 2,784, US$ 2,088, and US$ 1,392
for hydro operation and maintenance cost.
A constraint of 95% of energy supplies must be obtained from renewable
resources limits the grid purchases. The values for AC load are adopted to deter-
mine the dimensions of the main components of the hybrid system. PV costs
multipliers were chosen to assess the impact of floating structures, adding 30% to
the costs, and to evaluate possible cost reductions obtained through some kind of
financial or economic incentives on the price of PV modules.
The simulations were repeated a few times with different data to understand
the effects of generating power with a flow rate slightly higher than the ecological
flow to reduce dependence on the grid overnight. This possibility can be made
easier if the small reservoir formed with the proposed hydro power plant present
daily regulation capacity of the flow rate. Some additional simulations were also
performed to study the effects of small variations in the height used for the turbines
on the storage capacity and the system performance.
Use of residual flow for a PV hydro hybrid system 197

9.6 Results and discussion


The results are presented in nine figures and are discussed in this section. The first
five figures (Figures 9.11–9.15) show directly results obtained with HOMER. The
following four figures (Figures 9.16–9.19) were established by collecting the
results of various simulations with HOMER.
Figure 9.11 shows the optimization space for the system showing PV capital
cost multiplier as a function of the load, when the capital cost of the hydraulic
system is US$ 69,600, with a limitation of grid purchases at 5% of the load. The
optimization space indicates that the hydropower plant is the solution for loads up
to approximately 850 kWh/d, whereas the PV hydro hybrid system is the best
solution for loads above 850 kWh/d. The PV cost multiplier and the different hydro
capital costs apparently did not influence this result.
Figure 9.12 shows the PV array capacity for the system, showing PV capital
cost multiplier as a function of the load, when the capital cost of the hydraulic
system is US$ 69,600, also with limitation of grid purchases at 5% of the load. For
load consumption of less than about 850 kWh/d, the installation of photovoltaic
panels is not recommended (as also shown in the previous graph). For values above
850 kWh/d, the power of the PV modules increases gradually until it reaches 90 kW
when consumption is close to 925 kWh/d.

Optimal system type

1.2
PV capital multiplier

1.0

0.8

0.6

750 800 850 900


Load (kWh/d)
System types
Grid/hydro
Grid/hydro/PV
Fixed
Hydro capital = $69,600

Figure 9.11 Results for the optimization space obtained for the system of
Figure 9.6
198 Modeling and dynamic behaviour of hydropower plants

PV array capacity
Legend
100 kW
0 0 0 0 0 0 30 30 30 30 60 60 90 90
1.2 80
70
PV capital multiplier

60
50
40
1.0 30
0 0 0 0 0 0 30 30 30 30 60 60 90 20
10
0
0.8 Superimposed
PV array capacity (kW)
0 0 0 0 0 0 30 30 30 30 60 60 90 Fixed
Hydro capital = $69,600
0.6
0 0 0 0 0 0 30 30 30 30 60 60 90

750 800 850 900


Load (kWh/d)

Figure 9.12 Results for the PV array capacity obtained for the system of
Figure 9.6

Grid purchases
Legend
25,000kWh
22,500
1.2 20,000
17,500
PV capital multiplier

15,000
12,500
10,000
1.0 7,500
5,000
2,500
0
0.8 Fixed
Hydro capital = $69,600

0.6

750 800 850 900


Load (kWh/d)

Figure 9.13 Results for the grid purchases obtained for the system of Figure 9.6

Figure 9.13 shows the grid purchases for the system, showing PV capital cost
multiplier as a function of the load, when the capital cost of the hydraulic system is
US$ 69,600. The purchase of energy from the grid will only occur when the load
consumption exceeds the approximate value of 810 kWh/d. For smaller values, the
installed capacity of the hydro-electric power plant is always greater than the
minimum rated power in these simulations.
In the range between 810 and 850 kWh/d, the interconnected system is
responsible for supplying the load that is not supplied by the hydroelectric power
plant. For values greater than 850 kWh/d, the photovoltaic modules are responsible
for generating part of the energy required during the time where there is sunlight,
but during the night, the interconnected system completes the required energy.
Use of residual flow for a PV hydro hybrid system 199

Leveled cost of energy


Legend
0.30 S/kWh
0.031 0.030 0.029 0.028 0.031 0.036 0.118 0.118 0.118 0.118 0.200 0.199 0.278
0.27
1.2 0.24
0.21
0.18
0.15
PV capital multiplier

0.12
1.0 0.036 0.099 0.099 0.099 0.099 0.162 0.162 0.222
0.09
0.031 0.030 0.029 0.028 0.031
0.06
0.03
0.00
0.8 Superimposed
0.031 0.030 0.029 0.028 0.031 0.036 0.082 0.082 0.083 0.083 0.131 0.130 0.176 Leveled COE
(S/kWh)
Fixed
0.6
Hydro capital
0.031 0.030 0.029 0.028 0.031 0.036 0.065 0.066 0.067 0.067 0.099 0.099 0.130
= $69,600
750 800 850 900
Load (kWh/d)

Figure 9.14 Results for the leveled cost of energy obtained for the system of
Figure 9.6, with hydro capital cost equal to US$ 69,600

Leveled cost of energy


Legend
0.016 0.015 0.015 0.014 0.017 0.022 0.105 0.105 0.106 0.106 0.188 0.187 0.266
0.30 S/kWh
0.27
1.2 0.24
0.21
PV capital multiplier

0.18
0.15
0.12
1.0 0.016 0.015 0.015 0.014 0.017 0.022 0.085 0.086 0.086 0.087 0.150 0.149 0.210
0.09
0.06
0.03
0.00
0.8 Superimposed
Leveled COE (S/kWh)
0.016 0.015 0.015 0.014 0.017 0.022 0.069 0.069 0.070 0.071 0.118 0.118 0.164 Fixed
Hydro capital = $34,800
0.6
0.016 0.015 0.015 0.014 0.017 0.022 0.052 0.053 0.054 0.055 0.086 0.087 0.118

750 800 850 900


Load (kWh/d)

Figure 9.15 Results for the leveled cost of energy obtained for the system of
Figure 9.6, with hydro capital cost equal to US$ 34,800

Figure 9.12 shows that it is possible to find viable solutions with PV modules
with 30, 60, and 90 kW. A hybrid system with the 34.8 kW for the proposed
hydroelectric power plant and 30 kW for PV modules is already a viable combi-
nation of these components. For values of 30, 60, and 90 kW for the PV modules,
Figure 9.13 indicates increasing values for the grid purchases. The implementation
of a PV hydro hybrid system will meet a greater load, which however will not have
energy at night without the adoption of some kind of energy storage. The energy
purchased from the grid will then be required.
Figures 9.14 and 9.15 show, respectively, the leveled cost of energy obtained
for the system, with hydro capital costs equal to US$ 69,600 and US$ 34,800,
200 Modeling and dynamic behaviour of hydropower plants

also with limitation of grid purchases at 5% of the load. Considering consumption


lower than about 850 kWh/d, the cost of energy is lower than US$ 0.06/kWh in
both the graphs. The average cost of energy remains less than US$ 0.06/kWh for
the results of Figure 9.15 when consumption is greater than about 850 kWh/d and
less than 890 kWh/d for the PV cost multiplier near 0.5.
In other areas of both graphics, the cost of energy increases considerably and
may exceed values of US$ 0.25/kWh when daily consumption is around 925 kWh
and PV cost multiplier is equal to the maximum value (considered in the simula-
tions) of 1.3. Lower values of PV capital cost multiplier obviously lead to lower
values for the cost of energy. Among Figures 9.14 and 9.15, the cost per kWh is
lower for the hydro capital cost equal to US$ 69,600/kW installed, getting between
US$ 0.012/kWh and US$ 0.015/kWh.
Figures 9.16 and 9.17 show the cost of energy as a function of the load to the
system of Figure 9.6, with hydro capital cost equal to US$ 34,800 and respectively
with PV cost multiplier equal to 0.5 and 0.75. The region in these graphs in which
the use of the photovoltaic modules is economically more favorable is examined to
see the influence of power generation from the use of a higher flow than the eco-
logical flow during the evening. This additional flow could simply be used when
there is spillage of water through the spillways of Guarita power plant.
In both graphs, the cost of energy decreases with increasing load, mainly due to
the replacement of expensive energy provided by the grid for a cheaper electricity
generated by the hydroelectric power plant. The jump in the cost of energy in both
figures when daily consumption is respectively close to 895 and 920 kWh happens
by increasing the size of the photovoltaic system to be installed. The difference in
the cost of energy is also highlighted when larger loads are supplied.
Figure 9.18 shows the additional flow to be fueled overnight as a function of
daily consumption, considering a conversion efficiency of hydraulic energy into

0.12
With extra hydro
Cost of energy (US$/kWh)

0.10
Without extra hydro
0.08

0.06

0.04

0.02

0.00
800 820 840 860 880 900 920 940
Load (kWh/d)

Figure 9.16 Cost of energy as a function of the load to the system of Figure 9.6
with hydro capital cost equal to US$ 34,800 and PV cost multiplier
equal to 0.5
Use of residual flow for a PV hydro hybrid system 201

0.18
With extra hydro

Cost of energy (US$/kWh)


0.15 Without extra hydro

0.12

0.09

0.06

0.03

0.00
800 820 840 860 880 900 920 940
Load (kWh/d)

Figure 9.17 Cost of energy as a function of the load to the system of Figure 9.6
with hydro capital cost equal to US$ 34,800 and PV cost multiplier
equal to 0.75

60

50

40
Flow (l/s)

30

20

10

0
800 820 840 860 880 900 920 940
Load (kWh/d)

Figure 9.18 Additional flow to be fueled overnight as a function of daily


consumption

electricity equal to 80% and the height of the hydroelectric plant equal to 12 m.
The ecological flow is 370 l/s and, for example, a consumption of 890 kWh/d
will correspond to an extra flow to turbine, approximately equal to 10% of ecolo-
gical flow.
Figure 9.19 shows height increase necessary to ensure daily regulation of flow,
as a function of daily consumption, considering a flooded area of 0.0287 km2 with
the water reservoir formed by the proposed hydro power plant. For the daily con-
sumption of 890 kWh, the water reservoir allowing daily regulation of flow
corresponds to an increase in height of just 0.045 m and for a daily consumption of
920 kWh corresponds to an increase of just 0.07 m.
202 Modeling and dynamic behaviour of hydropower plants

0.08
0.07
0.06
0.05
Height (m)

0.04
0.03
0.02
0.01
0.00
800 820 840 860 880 900 920 940
Load (kWh/d)

Figure 9.19 Height increase necessary to ensure daily regulation of flow

9.7 Conclusions

This chapter presented a pre-feasibility study for the increase of hydroelectric


capacity in the existing hydroelectric power plant of Guarita, in southern Brazil,
generating energy from its ecological flow. The proposed hydropower plant will
have a height of 12 m and flow rate of 370 l/s, providing steady supply of 34.8 kW.
An additional power can be obtained with the installation of 30 kW in PV modules
on floating structures on water surface of the small reservoir formed with the pro-
posed plant. The PV modules allow the delivery of higher energy supply, but these
will require power supplies overnight, which should be obtained from the grid.
However, further studies may assess whether a small increase in height or flow
(if possible) may exempt the acquisition of these extra supplies.

Acknowledgments
This work was developed as a part of research activities on renewable energy
developed at the Escola de Engenharia and Instituto de Pesquisas Hidráulicas, at
Universidade Federal do Rio Grande do Sul. The authors acknowledge the support
received by the institution. The second author acknowledges the financial support
received from CNPq for his research work.

References
[1] Beluco, A., Kroeff, P.K., Krenzinger, A. (2012) A method to evaluate the
effect of complementarity in time between hydro and solar energy on the
performance of hybrid hydro PV generating plants. Renewable Energy,
vol. 45, pp. 24–30.
Use of residual flow for a PV hydro hybrid system 203

[2] Google Maps. Location of the municipality of ErvalSeco, State of Rio


Grande doSul, Brazil. Available at https://goo.gl/maps/MGz56eiJgCw.
Accessed January 25, 2016.
[3] Google Maps. Location of the Municipality of Redentora, State of Rio
Grande doSul, Brazil. Available at https://goo.gl/maps/pM15sqqCtsG2.
Accessed January 25, 2016.
[4] Google Maps. Location of the State of Rio Grande doSul, Brazil. Available
at https://goo.gl/maps/GTdxfjFcQiR2. Accessed January 25, 2016.
[5] Google Maps. Location of the Guarita Hydroelectric Power Plant,
ErvalSeco, Rio Grande do Sul, Brazil. Available at https://goo.gl/maps/
ZAwdFfifj182. Accessed January 25, 2016.
[6] Ferrer-Gisbert, C., Ferran-Gonzalvez, J.J., Redon-Santafé, M., Ferrer-Gisbert,
P.S., Sanchez-Romero, F.J., Torregrosa-Soler, J.B. (2013) A new photovoltaic
floating cover system for water reservoirs. Renewable Energy, vol. 60,
pp. 63–70.
[7] Redon-Santafé, M., Ferrer-Gisbert, P.S., Sanchez-Romero, F.J., Torregrosa-
Soler, J.B., Ferran-Gonzalvez, J.J., Ferrer-Gisbert, P. (2013) Implementation
of a photovoltaic floating cover for irrigation reservoirs. Journal of Cleaner
Production, vol. 66, pp. 568–570.
[8] Software HOMER, version 2.68 beta. The Micropower Optimization Model,
Homer Energy. Available at www.homerenergy.com.
[9] Lilienthal, P.D., Lambert, T.W., Gilman, P. (2004) Computer modeling of
renewable power systems. In: Cleveland, C.J. (ed.) Encyclopedia of Energy,
Elsevier, Boulder, Colorado, USA, vol. 1, pp. 633–647. NREL Report
CH-710-36771.
[10] Lambert, T.W., Gilman, P., Lilienthal, P.D. (2005) Micropower system
modeling with Homer. In: Farret, F.A., Simões, M.G. (eds.) Integration of
Alternative Sources of Energy, John Wiley & Sons, Hoboken, New Jersey,
USA, pp. 379–418. ISBN 0471712329.
[11] IRENA, International Renewable Energy Agency (2012) Renewable energy
technologies: cost analysis series, Hydropower. Volume 1: Power Sector.
Available at www.irena.org/documentdownloads/publications/re_technolo
gies_cost_analysis-hydropower.pdf. Accessed on February 23, 2016.
[12] Feldman, D., Barbose, G., Margolis, R. et al. (2014) Photovoltaic system
pricing trends: historical, recent and near term projections. National
Renewable Energy Laboratory, US Department of Energy. Report No.
62558. Available at www.nrel.gov/docs/fy14osti/62558.pdf. Accessed on
February 23, 2016.
Chapter 10
A PV wind hydro hybrid system with pumped
storage capacity installed in Linha Sete,
Aparados da Serra, southern Brazil
Alfonso Risso1, Fausto A.Canales1, Alexandre Beluco1
and Elton G.Rossini2

Abstract
The intermittency and variability of various renewable energy resources, such as
wind power and photovoltaic solar energy, can overcome with the use of these
resources in conjunction with energy storage devices. The energy storage as
hydraulic power, so before energy conversion, can guarantee high efficiency to the
storage process. This study aims to identify the technical and economic feasibility
of using wind power and PV modules in conjunction with a reversible hydroelectric
power plant installed in Aparados da Serra, in the south of the Serra Geral, a
geological structure in southern Brazil that allows topographical height differences
of approximately 600 m. In this work, specifically, a hydropower plant installed at
Linha Sete with 610 kW and at 400 m height. This study explores the feasibility of
this pumped storage plant operating in conjunction with existing wind turbines and
PV modules installed on the surface of reservoirs. The work is based on simulations
and optimization performed with well-known software HOMER. The results indi-
cate that a group of 10–50 2-MW wind turbines may have an increased capacity
factor from usual 0.34 to values between 0.50 and 0.60. The results also relate the
power capacity and costs per kW installed for PV modules to be feasible. This work
also indicates useful conclusions in the design process and implementation of the
hybrid system under study.

Keywords
Wind energy, wind diesel hybrid systems, Weibull shape parameter, southern
Brazil, computational simulation, software HOMER

1
Instituto de Pesquisas Hidráulicas, Universidade Federal do Rio Grande do Sul (UFRGS), Porto Alegre,
Rio Grande do Sul, Brazil
2
Universidade Estadual do Rio Grande do Sul (UERGS), Porto Alegre, Rio Grande do Sul, Brazil
206 Modeling and dynamic behaviour of hydropower plants

10.1 Introduction
Brazil is blessed with one of the largest water resource systems and hydroelectric
potential in the world. Thus, Brazil has in its territory some of the largest hydro-
electric power plants and a lot of water reservoirs with large bodies of water arti-
ficially formed. As a result, Brazil is one of the few countries that have an energy
system that is largely based on hydropower. As it is a nonintermittent source of
energy, a wide base made with hydropower favors the use of renewables.
The current time of crisis in the global scenario, for various reasons, con-
tributes to the increasing encouragement of the use of renewable energy. Among
the different alternatives, recent years have seen a considerable increase in the use
of wind turbines and photovoltaic modules. Both for wind energy and photo-
voltaics, as for other alternatives for power generation, a greater number of new
plants will result in higher production of equipment and a trend of reduction in
installation costs as well as in operation and maintenance costs.
In this scenario, it is almost obvious to consider the implementation of pho-
tovoltaic modules on the water surface of reservoirs formed by hydroelectric plants.
The PV modules will not shadow useful areas and, covering the surface flooded by
the reservoirs, they will contribute to the reduction of water loss by evaporation.
Thus, it will be possible to generate photovoltaic energy and to have a larger
amount of water to hydroelectric power generation. The panels can be installed on
floating structures modulated with a given power, possibly produced in series.
The association between hydroelectric power plants and photovoltaic power
plants might seem strange in the past when photovoltaic plants with reasonable
values of power did not exist. But hydropower is ‘‘constant’’ and ‘‘more available,’’
whereas photovoltaics is ‘‘intermittent’’ by weather issues and ‘‘less available’’ by
its own characteristics. It is precisely the constancy of hydroelectric power plants
(and notoriously large hydropower with large storage capacity) that enables greater
investment in photovoltaic farms.
A prime example is the hydroelectric power plant in Longyangxia Dam, on the
Yellow River, in northwest China. The hydropower plant was installed in 1992,
with 1,280 MW of installed capacity and four machines with 320 MW each. Few
years ago, a project for a PV hydro hybrid system was started culminating in the
installation of 320 MW in 2013, a first phase covering 9 km2 and a further 530 MW
in 2015, covering another 15 km2. The PV hydro hybrid system and the photo-
voltaic power plant now constitute the largest in the world.
The design operation of photovoltaic hydro hybrid systems of this kind can
also be decisively influenced by the possible energetic complementarity between
hydro and solar energy availability [1–3]. The greater availability of solar energy
can occur in periods of low water availability, as well as less availability of solar
energy can coincide with increased water availability. The use of stored water in
the reservoir can be managed to increase this effect of energetic complementary.
This chapter presents a feasibility study for the implementation of a pumped
storage hydroelectric power plant (or reversible hydroelectric power plant) oper-
ating with wind turbines and photovoltaic modules. The study is based on the
A PV wind hydro hybrid system with pumped storage capacity 207

results obtained with the well-known software HOMER. The next section describes
the reversible power plant planned at a place called ‘‘Linha Sete,’’ in southern
Brazil and also describes how this plant will be simulated with HOMER.
Subsequent sections describe the components of the hybrid system under study,
the results and discussions and finally the conclusions.
This chapter presents the results of an exploratory study on the operation of a
planned pumped storage plant with a set of wind turbines in operation in southern
Brazil, in the city of Osório, in a place where wind potential is known and is
currently being explored. This chapter shows results from a project that also led to
an article [4] reporting conclusions already obtained on the operation of this
pumped storage plant.

10.2 The Linha Sete pumped storage power plant


The hydraulic system considered in this work was identified in an earlier work of
the research group [5]. It is a set of areas and storage volumes that allow the
implementation of two reservoirs in a region in southern Brazil, where there are
strong topographical height differences. Figure 10.1 shows the upper and lower
reservoirs and their watersheds. The content of Figure 10.1 was prepared from a
region that appears on Google Maps and is located according to [6].
The reservoirs have been sized for a storage volume of 1510,000 cubic meters.
The lower reservoir has a maximum quota at 290 m and the upper reservoir at
840 m. The total height is 655 m, whereas the machine room is placed 105 m below
the maximum level of the lower reservoir. Simulations have limited accuracy as

Figure 10.1 Upper and lower reservoirs in ‘‘Linha Sete’’ and their watersheds
208 Modeling and dynamic behaviour of hydropower plants

Hydro resource
1,000

800
Stream flow (L/s)

600

400

200

0
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Figure 10.2 Monthly average stream flow rate available to the turbine at power
house, already considering the residual flow

the variations in height resulting from operation of the pumped storage system are
not simulated by HOMER, as discussed below and presented by Canales and
Beluco [7].
The natural flow available to the lower reservoir was determined by Canales
et al. [4] and is equal to 0.539 m3/s. This flow will be available for generation in
addition to the flow rates obtained with the management of reversible plant, starting
from the moment that the lower reservoir is full. Based on the Tennant method
describer by Benetti et al. [8], 10% of the annual average flow was adopted as
residual flow. Figure 10.2 shows the stream flow rate available to turbine each
month, already considering the residual flow.

10.3 Components of the PV wind hydro hybrid system


The hybrid system consists of the pumped storage hydropower plant described in
the previous section, operating in conjunction with wind turbines and photovoltaic
modules with diesel generators support. The operation of the pumped hydro and
wind turbines has been the subject of a recent article [4] pointing that the operation
of the two reservoirs, even demanding higher initial costs, leads to lower environ-
mental impacts.
Wind farms in Osório, in southern Brazil, were considered in this study. The
three wind farms contain 75 turbines model Enercon E-70 E4, providing a total
power of 75 MW operated at a capacity factor of about 34%. Based on Braciani [9],
the average cost per installed kilowatt in wind farms in Brazil is around US$
2,156.50/kW. By using this value, the initial cost of each E-70 turbine was set at
US$ 4,313,000 in HOMER.
Figure 10.3 shows the wind data used for simulations with HOMER. The
monthly average wind speed in Osório at 100 m above the ground was extracted
from Silva [10]. These data were used to obtain a synthetic series of hourly wind
A PV wind hydro hybrid system with pumped storage capacity 209

Scaled data monthly averages


30
Average value (m/s) Max
25
Daily high
20
Mean
15
Daily low
10 Min
5
0
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec Ann
(a) Month
Scaled data m/s
24 30
24
Hour of day

18 18
12
12
6
6 0

0
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
(b) Day of year

Figure 10.3 Wind resource input for the case study

speed data to the operation site of the wind turbines of the wind parks at Osório.
Figure 10.3 shows two graphs. At first, the average wind speed for each month, the
deviations around these averages and maximum and minimum values are shown.
This graph shows the typical variability of the wind. The second graph, with strong
variation in color, enhances the variability of the wind over days and months.
The photovoltaic modules will be installed on floating structures, as recently
proposed by Ferrer-Gisbert et al. [11] and Redon-Santafé et al. [12]. The basic
model for the floating structure considered in this study has dimensions suitable for
50 kW of PV modules. The total area of the water surfaces formed with the two
dams is small but sufficient for several tens of structures having these dimensions.
Figure 10.4 shows the incident solar radiation data used in the simulations and
obtained automatically by HOMER in a NASA database.
Figure 10.4 also shows two graphs. At first, the average incident solar radiation
on a horizontal plane for each month, the deviations around these averages and
maximum and minimum values are shown. The maximum insolation occurs in
January, whereas the minimum occurs in June. In the second graph, it is evident that
the variation of sunlight available throughout the hours of the day, with the lowest
values available in the first and the last hour of the day, and the available peak near
midday. Also evident is the change in hours of the day throughout the year.
The cost of the PV modules was considered as US$ 4,380/kW, and it is com-
patible with usual costs found, for example by Feldman et al. [13]. The installation
of floating structures, as suggested by Ferrer-Gisbert et al. [11] and Redon-Santafé
et al. [12], raise the cost by 30%. The lifetime of the PV system is considered to be
12.5 years, the replacement cost of the PV system at the end of the useful life is
80% of the initial cost and annual cost of operation, and maintenance is 5% of the
210 Modeling and dynamic behaviour of hydropower plants

Scaled data monthly averages


1.4
1.2 Max
1.0 Daily high
0.8 Mean
0.6 Daily low
0.4 Min
0.2
0.0
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec Ann
(a) Month
Scaled data kW/m2
24 1.40
1.12
Hour of day

18 0.84
0.56
12
0.28
6 0.00

0
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
(b) Day of year

Figure 10.4 Incident solar radiation on a horizontal plane for the reservoirs
location, obtained with software HOMER, considered in this study

Scaled data monthly averages


35,000
Average value (kW)

30,000 Max
Daily high
25,000 Mean
20,000
Daily low
15,000 Min
10,000
5,000
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec Ann
Month

Figure 10.5 Scaled monthly averages of load profile considered in this study

installation cost. The reflectance of the water surface was considered to be 10% at the
installation site. Figure 10.5 shows the monthly average output power of PV modules.
Diesel generator sets were considered as support in the simulations, for the
times when the availability of renewable energy is not enough to meet the energy
demand. The average cost per installed kilowatt for a thermoelectric plant in Brazil
was set at US$ 1,073.50/kW, according to Braciani [9]. Several generator sizes
were considered, with the technical minimum load ratio set at 30%, according to
Kaldellis et al. [18] for heavy oil and diesel engines.
A connection to the grid was included, allowing the purchase of energy when
there is not enough energy production to meet the consumers, and the sale of energy,
when there is excess energy. The connection to the grid considered in the simulations
has dimensions comparable to the possible installed powers of the diesel generator
sets, allowing eventually the optimization process to choose one over the other.
Figure 10.5 shows monthly averaged load profile considered in this study.
A PV wind hydro hybrid system with pumped storage capacity 211

10.4 Simulations with HOMER


HOMER [14] is a software for optimization of hybrid energy systems. It was ori-
ginally developed by National Renewable Energy Laboratory and a version called
‘‘Legacy’’ is now available for universal access. HOMER simulates a system for
power generation over a period of 25 years at standard intervals of 60 min [15,16].
The HOMER software performs simulations of hybrid systems aiming to build
optimization spaces according to different sensitivity variables, allowing a com-
plete characterization of performance.
The HOMER software simulates hydroelectric power plants operating as
‘‘run-of-the-river’’ power plants. The simulation of hydroelectric power plant with
storage capacity and hydroelectric for operation as a reversible plant can be
performed as explained by Canales and Beluco [7]. The DC bus must contain only
the hydroelectric power plant and a battery adjusted to simulate a pumped storage
power plant. The operation of the two reservoirs is simulated with the battery, while
the supply of electricity to be transferred to the hybrid system is simulated by
the hydroelectric plant. The converter has a single direction of operation.
Simulations with the system of Figure 10.6 were performed. The optimization
variables considered were the following: 0, 10, 20, 25, 30, 35, 40, 45, 50, 55, 60,
65, 70, 75, 80, 85, and 90 wind turbines; 0, 10, 20, 30, 40, 50, and 60 MW for the
installed power of the diesel gen set; 0 and 1 battery modeled as pumped storage
plant; 0 and 722 kW for the converter capacity. The sensitivity inputs were the
following: 100, 200, 300, 400, and 500 MWh/d for AC load; US$ 1/L, US$ 2/L,
US$ 3/L, US$ 4/L, and US$ 5/L for the cost of diesel oil; 6, 8, 10, and 12 m/s for
the wind speed. Simulations with the system of Figure 10.6 were repeated with all

Figure 10.6 Wind hydro hybrid system with water storage capacity
212 Modeling and dynamic behaviour of hydropower plants

Figure 10.7 PV wind hydro hybrid system with water storage capacity considered
in this study

these variables and a fixed value of 10 MW for the installed capacity of photo-
voltaic plant.
Simulations with the system of Figure 10.7, with the PV modules assembled on
floating structures installed over the flooded surface of the reservoir, were performed.
The optimization variables considered were the following: 0, 10, 20, 25, 30, 35, 40,
45, 50, 55, 60, 65, 70, 75, 80, 85, and 90 wind turbines; 0, 1,200, 2,400, 4,800, 9,600,
and 19,200 MW for the installed power of the diesel gen set; 0, 100, 200, 400, and
800 kW for the capacity of PV modules; 0 and 1 battery modeled as pumped storage
plant; 0 and 722 kW for the converter capacity. The sensitivity inputs were the fol-
lowing: 100, 200, 300, 400, and 500 MWh/d for AC load; US$ 0.50/L, US$ 0.70/L,
US$ 0.90/L, and US$ 1.10/L for the cost of diesel oil; 6, 8, 10, and 12 m/s for the wind
speed; 0.0%, 2.5%, 5.0%, and 10.0% for the maximum capacity shortage.
A constraint of 95% of energy supplies must be obtained from renewable
resources limits the grid purchases. The values for AC load are adopted to deter-
mine the dimensions of the main components of the hybrid system. PV costs
multipliers were chosen to assess the impact of floating structures, adding 30% to
the costs and to evaluate possible cost reductions obtained through some kind of
financial or economic incentives on the price of PV modules.

10.5 Results and discussion

Figures 10.8–10.11 show the results obtained with the first stage of the simulation,
whereas Figures 10.12–10.15 show the results obtained with the second phase.
A very important result is that HOMER did not indicate any optimal solutions in
A PV wind hydro hybrid system with pumped storage capacity 213

Optimal system type


5

4
Diesel price ($/L)

1
100,000 200,000 300,000 400,000 500,000
RS typical load (kWh/d)

System types Fixed


Hydro/Wind/GEN1/Battery Inverter efficiency = 100%
Wind/GEN1/Battery Rectifier efficiency = 100%
OR Wind = 25%

Figure 10.8 Results for the optimization space obtained for diesel price as a
function of local typical load for the system of Figure 10.6

Optimal system type


5

4
Diesel price ($/L)

1
6 7 8 9 10 11 12
Wind speed (m/s)

System types Fixed


Hydro/Wind/GEN1/Battery PV capital multiplier = 1
Wind/GEN1/Battery

Figure 10.9 Results for the optimization space obtained for diesel price as a
function of wind speed for the system of Figure 10.6
214 Modeling and dynamic behaviour of hydropower plants

Figure 10.10 Optimization results constituting the optimization space shown


in Figure 10.12

Monthly statistics
100
Max
80 Daily high
Mean
SOC (%)

60
Daily low
40 Min

20

0
(a) Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
Battery bank state of charge %
24 100
Hour of day

18 80
60
12 40
20
6 0
0
(b) Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Figure 10.11 Annual change in state of charge of the reservoirs for a hybrid
system shown in Figure 10.10 with COE equal to US$ 0.609/kWh

the different optimization spaces shown in these figures (and even others not
shown) that contained PV modules. However, in several cases, as discussed below,
some combinations containing photovoltaic modules were discarded by very small
differences in relation to optimal solutions.
Figures 10.8 and 10.9 show the optimization space obtained for the system of
Figure 10.6, respectively showing diesel price as a function of the local typical load
and showing diesel price as a function of wind speed. The value currently practiced
for diesel oil and the average wind speed for the area indicate that the optimal
solution includes wind turbines and the pumped storage plant, in addition to sup-
porting diesel generators. This system, considered as a starting point for this study,
was the subject of a recent article [4].
A PV wind hydro hybrid system with pumped storage capacity 215

Optimal system type


1.1

1.0
Diesel price ($/L)

0.9

0.8

0.7

0.6

0.5
50,000 100,000 150,000 200,000
RS typical load (kWh/d)

System types Fixed


Hydro/GEN1/Battery PV capital multiplier = 1
Hydro/Wind/GEN1/Battery Hydro capital = $270,000
Wind/GEN1/Battery Max. annual capacity shortage = 0%

Figure 10.12 Results for the optimization space obtained for diesel price as a
function of local typical load, for the system of Figure 10.7 with
different values for diesel price and consumers load

Optimal system type


1.1

1.0
Diesel price ($/L)

0.9

0.8

0.7

0.6

0.5
6 7 8 9 10 11 12
Wind speed (m/s)

System types Fixed


Hydro/GEN1/Battery RS typical load = 200,000 KWh/d
Hydro/Wind/GEN1/Battery PV capital multiplier = 1
Wind/GEN1/Battery Max. annual capacity shortage = 0%

Figure 10.13 Results for the optimization space obtained for diesel price as a
function of wind speed, for the system of Figure 10.7 with different
values for diesel price and consumers load
216 Modeling and dynamic behaviour of hydropower plants

Figure 10.14 Optimization results constituting the optimization space shown in


Figure 10.12

Monthly statistics
100
Max
80 Daily high
Mean
SOC (%)

60 Daily low
40 Min

20
0
(a) Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Battery bank state of charge %


24 100
Hour of day

18 80
60
12 40
6 20
0
0
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
(b)

Figure 10.15 Annual change in state of charge of the reservoirs for a hybrid
system shown in Figure 10.12 with COE equal to US$ 0.495/kWh

Figure 10.10 shows the simulation results of this system with a photovoltaic
plant with capacity of 10 MW. The optimal result indicates an energy cost of US$
0.469/kWh operating with 20 wind turbines and diesel support system with 30 MW.
The seventh system in this list operate without wind turbines and a higher cost,
equal to US$ 0.609/kWh, with a variation of the charge state of the reservoirs
shown in Figure 10.11. The behavior of the curve, with energy at the beginning of
the period identical to the energy in the end, indicates an acceptable performance.
Figures 10.12 and 10.13 show the optimization space obtained for the system of
Figure 10.6, respectively, showing diesel price as a function of the local typical load
A PV wind hydro hybrid system with pumped storage capacity 217

and showing diesel price as a function of wind speed. The lower load consumers,
compared with the preceding figures, have given rise to areas on the bottom left of
these optimization spaces, corresponding to combinations not including wind tur-
bines. A lot of points of these optimization spaces show optimal results that led to
discard the combinations including PV modules for very small differences.
The complete output provided by HOMER for each feasible option allows
estimating and optimizing the capacity of pumped storage system for recovering
rejected renewable energy. This can also be used for calculating the effective
capacity factor of the wind farm with and without the pumped storage capacity.
According to the simulation results and based on the wind resource inputs of this
case study, the maximum capacity factor of the wind turbines reported by HOMER
is 35.6%, including excess electricity. The Osório Wind Park, used as model for
creating the wind turbines of this work, reports on its website a capacity factor
equal to 32.3%. These values are within the range of values reported by Boccard
[17], who gather global results reported by transmission system operators or
available in academic literature related to wind farm capacity factors.
On the basis of the result simulations, Figure 10.16 presents the estimated
duration curves for rejected power. As explained by Kaldellis et al. [18], large
amounts of rejected energy also mean severe financial losses that discourage future
investments in renewable energy projects. Without the pumped storage plant, the
extremely variable wind profile would require more turbines at the wind farm along
with a diesel generator of greater capacity, thus increasing the generation cost. As
shown in Figure 10.16(a), a system without storage capacity would reject power
about 80% of the time, with 25% of the time rejecting more than 50 MW. On the
contrary, with pumped storage capacity and using the same 50% of the wind farm
capacity as benchmark, Figure 10.16(b) and (c) shows that this energy storage
technology improves the wind energy absorption, limiting the occurrence of this
value to less than 10% of the time and reducing the cost of energy for the system. In
HOMER, the cost of energy is the average cost per kWh of useful electrical energy
produced by the system, which in this case is just the energy used to serve the primary
AC load (no grid sales, DC or deferrable loads are considered in the example).
Figure 10.14 shows results presented in Figure 10.12 and corresponding to the
consumer load equal to 200 kWh/d and diesel sold at US$ 0.90/L. The first system
of this list is what defines the green color at the corresponding point in the opti-
mization space shown in Figure 10.12. This first system presents cost of energy
equal to US$ 0.407/kWh, very close to the third system in the list that includes
PV modules and provides energy at cost of US$ 0.408/kWh. This system includes
100 kW in PV modules, the pumped storage plant, 20 wind turbines and diesel
support system with 19.2 MW.
Most systems in this list have small monthly variations of the state of charge
of the reservoirs over a year. The tenth system, however, in the list shown in
Figure 10.14, presents a more pronounced change in the state of charge of the reser-
voirs. This system includes 100 kW in PV modules, the pumped storage plant and
diesel support system without wind turbines. Figure 10.15 details this change in the
state of charge, indicating minimum values during the month of June. A further
Diesel price: US$1/L Diesel price: US$S/L Diesel price: US$S/L
Wind farm capacity: 100.0 MW Wind farm capacity: 90.0 MW Wind farm capacity: 80.0 MW
Size of AC diesel generator: 50.0 MW Size of AC diesel generator: 0 MW Size of AC diesel generator: 40.0 MW
Hydropower capacity: 0 MW Hydropower capacity: 46.1 MW Hydropower capacity: 46.1 MW
Cost of energy: US$0.561/kWh Cost of energy: US$0.192/kWh Cost of energy: US$0.181/kWh

100 100 100

80 80 80

60 60 60

40 40 40

20 20 20

Rejected power (MW)


Rejected power (MW)
Rejected power (MW)

0 0 0
0% 50% 100% 0% 5% 10% 15% 20% 25% 30% 0% 5% 10% 15% 20% 25% 30%
(a) Cumulative time (%) (b) Cumulative time (%) (c) Cumulative time (%)

Figure 10.16 Rejected power duration curves for an average daily load ¼ 500 MWh/d for three different conditions
A PV wind hydro hybrid system with pumped storage capacity 219

reduction in the past few months shows that the energy available at the end of the year
will be less than the energy in the beginning of the year, indicating an unsustainable
situation.

10.6 Final remarks


This chapter presented the results of an exploratory study to design a photovoltaic
wind hydro hybrid system with storage capacity. The optimal combinations
obtained from the simulations did not suggest the inclusion of PV modules, mainly
due to its high initial cost and the high cost of energy. Among the nonoptimal
solutions, it is possible to find solutions including PV modules that show perfor-
mance comparable to optimal solutions. Thus, this study suggested a hybrid system
constituted by 100 kW in photovoltaic modules, 20 wind turbines, and a diesel
support system with 19,200 kW, also with the pumped storage plant, providing
energy at a cost of US$ 0.408/kWh. This study also suggested a system with
10,000 kW in photovoltaic modules and a diesel support system with 30,000 kW,
without wind turbines and with the pumped storage plant, providing energy at a
cost of US$ 0.609/kWh.

Acknowledgments
This work was developed as a part of research activities on renewable energy
developed at the Instituto de Pesquisas Hidráulicas, at Universidade Federal do Rio
Grande do Sul, and Universidade Estadual do Rio Grande do Sul. The authors
acknowledge the support received by the institutions. The third author also
acknowledges the financial support received from CNPq for his research work.

References
[1] Beluco, A., Souza, P.K., Livi, F.P., Caux, J. (2012) Energetic com-
plementarity with hydropower and the possibility of storage in batteries and
water reservoirs, Chapter 7, In: Sørensen, B. (ed.) Solar Energy Storage,
Amsterdam, Netherlands: Academic Press.
[2] Beluco, A., Kroeff, P.K., Krenzinger, A. (2012) A method to evaluate the
effect of complementarity in time between hydro and solar energy on the
performance of hybrid hydro PV generating plants. Renewable Energy, vol. 45,
pp. 24–30.
[3] Beluco, A., Souza, P.K., Krenzinger, A. (2013) Influence of different
degrees of complementarity of solar and hydro availability on the perfor-
mance of hybrid hydro PV generating plants. Energy and Power Engineer-
ing, vol. 5, pp. 332–342.
[4] Canales, F.A., Beluco, A., Mendes, C.A.B. (2015) A comparative study of
a wind hydro hybrid system with water storage capacity: conventional reservoir
or pumped storage plant? Journal of Energy Storage, vol. 4, pp. 96–105.
220 Modeling and dynamic behaviour of hydropower plants

[5] Beluco, A. (2012) Three sites for implementation of reversible hydroelectric


power plant in the south of the Aparados da Serra, on the north coast of State
of Rio Grande do Sul (in Portuguese). Hidro & Hydro, no. 52, pp. 32–37.
Available at cerpch.unifei.edu.br/wp-content/uploads/revistas/revista-52.
pdf#page=32.
[6] Available at https://goo.gl/maps/FAgxVYLbF4u.
[7] Canales, F.A., Beluco, A. (2014) Modeling pumped hydro storage with the
micropower optimization model (Homer). Journal of Renewable and Sus-
tainable Energy, vol. 6, paper no. 043131.
[8] Benetti, A.D., Lanna, A.E., Cobalchini, M.S. (2003) Methods for the deter-
mination of residual flows in Rivers (in Portuguese). Revista Brasileira de
Recursos Hı́dricos, vol. 8, pp. 149–160. Available at www.abrh.org.br/
SGCv3/index.php?PUB=1.
[9] Braciani, U. (2011) Cost structure for Implementation of electric power
generation plants in Brazil (in Portuguese). BSc Thesis in Economic Sci-
ences, Universidade Federal de Santa Catarina (UFSC), Florianópolis, Brazil.
Available at tcc.bu.ufsc.br/Economia303023.pdf.
[10] Silva, J.S. (2012) Feasibility of electricity generation from ocean waves on
the northern coast of Rio Grande do Sul: a study of a hybrid system based on
renewable (in Portuguese). MSc Thesis in Water Resources, Universidade
Federal do Rio Grande do Sul (UFRGS), Porto Alegre, Brazil. Available at
www.lume.ufrgs.br/bitstream/handle/10183/78865/000900392.pdf.
[11] Ferrer-Gisbert, C., Ferran-Gonzalvez, J.J., Redon-Santafé, M., Ferrer-Gisbert,
P.S., Sanchez-Romero, F.J., Torregrosa-Soler, J.B. (2013) A new photovoltaic
floating cover system for water reservoirs. Renewable Energy, vol. 60,
pp. 63–70.
[12] Redon-Santafé, M., Ferrer-Gisbert, P.S., Sanchez-Romero, F.J., Torregrosa-
Soler, J.B., Ferran-Gonzalvez, J.J., Ferrer-Gisbert, P. (2013) Implementation
of a photovoltaic floating cover for irrigation reservoirs. Journal of Cleaner
Production, vol. 66, pp. 568–570.
[13] Feldman, D., Barbose, G., Margolis, R. et al. (2014) Photovoltaic system
pricing trends: historical, recent and near term projections. National
Renewable Energy Laboratory, US Department of Energy. Report no.
62558. Available at www.nrel.gov/docs/fy14osti/62558.pdf. Accessed on
February 23, 2016.
[14] Software HOMER, version 2.68 beta. The Micropower Optimization Model,
Homer Energy. Available at www.homerenergy.com.
[15] Lilienthal, P.D., Lambert, T.W., Gilman, P. (2004) Computer modeling of
renewable power systems. In: Cleveland, C.J. (ed.) Encyclopedia of Energy,
Elsevier, Boulder, Colorado, USA, vol. 1, pp. 633–647. NREL ReportCH-
710-36771.
[16] Lambert, T.W., Gilman, P., Lilienthal, P.D. (2005) Micropower system
modeling with Homer. In: Farret, F.A., Simões, M.G. (eds.) Integration of
A PV wind hydro hybrid system with pumped storage capacity 221

Alternative Sources of Energy, John Wiley & Sons, Hoboken, New Jersey,
USA, pp. 379–418. ISBN 0471712329.
[17] Boccard, N. (2009) Capacity factor of wind power realized values vs. esti-
mates. Energy Policy, vol. 37, no. 7, pp. 2679–2688.
[18] Kaldellis, J.K., Kapsali, M., Kavadias, K.A. (2010) Energy balance analysis
of wind-based pumped hydro storage systems in remote island electrical
networks. Applied Energy, vol. 87, no. 8, pp. 2427–2437.
Part IV
Small hydropower plants
Chapter 11
Modeling and simulation of a pico-hydropower
off-grid network
Sam J. Williamson1, Antonio Griffo2, Bernard H. Stark1
and Julian D. Booker1

11.1 Introduction
Nearly, 1 billion people who do not have access to electricity live in rural areas [1].
Extending the national power grid to many of these people is not feasible for
technical or economic reasons. These typically remote communities, therefore,
become reliant on local generation for their electricity supply. Diesel generators are
the most popular alternative with low capital expenditure and well-understood
technology, but with significant and fluctuating maintenance and running costs.
Therefore, renewable technologies – solar photovoltaics, wind, hydropower and
biomass – are attractive for off-grid communities, as they can provide locally
generated electricity from local resources.
Pico-hydropower is normally defined as electrical generation from a water
resource with the capacity of less than 5 kW [2]. Where this resource exists, pico-
hydropower is a suitable option for a community as it can operate over a range of
environmental conditions with different turbine designs, using simple and locally
manufactured technology [3]. The output power is constant over short time periods,
varying over longer periods due to seasonal changes so there is no requirement for
electricity storage. Pico-hydropower designs can be incorporated into already-
existing infrastructure, such as irrigation canals, with minimal environmental
impact [4,5]. Once the initial capital cost of a pico-hydropower system is covered,
the life-cycle cost is low and produces low-cost power with high availability [6].
Typical pico-hydropower systems are operated in stand-alone configuration, with
one unit feeding a number of houses. This system has no redundancy and is vul-
nerable to overload, such as the starting inrush current from an induction machine.
There is also no opportunity of expansion as demand increases.
Creating a pico-hydropower network, as shown in Figure 11.1, where identical
units across a geographic area are electrically connected together, is desirable as it
allows for a redundant, plug-and-play expandable system, with increased power

1
Faculty of Engineering, University of Bristol, Bristol BS8 1TR, UK
2
Department of Electrical and Electronic Engineering, University of Sheffield, Sheffield S1 3JD, UK
226 Modeling and dynamic behaviour of hydropower plants

Pico-hydro site
Power lines
Household/business

Map © CNES/Astrium, Google

Figure 11.1 Off-grid pico-hydropower network concept

generating capability allowing for both domestic and industrial loads. Each unit
must rely on local measurements for control and be able to be maintained and
serviced by local unskilled labour.
This chapter will cover the system arrangement, modeling and control of such
a system, with implementation of simulated example and expansion of the concept
to include solar PV and wind turbine sources.

11.2 System overview

The proposed pico-hydropower off-grid network is built up from a number of


identical pico-hydropower generator units (GUs), with one or more installed at
each turbine site dependent on the environmental conditions. As each site may have
different environmental conditions, each unit may have different power available at
the turbine, which may vary over time. The generator is interfaced onto the AC
network through a power electronic AC–DC–AC interface, which allows a dis-
tributed grid, shown in Figure 11.2. There is no communication between each unit,
increasing the system reliability, so each unit must use local measurements to
control the GU. This topology provides flexibility and redundancy in the network,
making it suitable for a scalable and expandable system.
Modeling and simulation of a pico-hydropower off-grid network 227

= = G
G = = = = = =
T T
= G
= = =
Local T
AC grid Load

Generator
G =
Shaft = = =
Rectifier DC–DC Inverter
T
Turbine Generator unit

Figure 11.2 Distributed AC grid topology with non-communicating inverter


front-ends

This type of network is sometimes referred to as a multi-master system [7] as


there is no central hub or single unit ensuring the network is properly controlled.
Instead, each unit within the network has responsibility for voltage and frequency
regulation, with the power sharing being a function of this control. This can lead to
small swings in the grid voltage and frequency, with associated deviations in
power sharing [8], but the major benefit is that there is no single point of failure,
and the grid can naturally evolve over time without any input required to the
existing installed units. The proposed network is made up of a number of GUs and
loads, all connected by an AC local grid. The loads and GUs are interspersed
geographically.

11.3 Component models


The GU can be broken down into six different components; a turbine, an output
drive shaft, a generator, a rectifier, a DC–DC converter and a voltage source
inverter, as shown in Figure 11.2. Each of these components have an internal model
which passes signals in the form of rotational speed, torque, voltage outputs or
current draw between them, as shown in Figure 11.3.
In modeling of the system, several assumptions are made. The water flow rate,
and therefore head, at the site is usually assumed to be constant over a short time
period (minutes) but will generally change over the longer term (hours), although an
intake blockage would be an exception to this. The water flow passes through the
penstock and nozzle to the turbine, which is directly connected to a three-phase
permanent magnet generator. The generator output is rectified to DC. As the turbine/
generator will rotate at varying speeds dependent on load and head, a DC–DC
228 Modeling and dynamic behaviour of hydropower plants

Generator Rectifier DC–DC Inverter Grid and load


Vg VDC Vinv vO

G
=
Ig = IDC = Iinv = iO
ωturb Telec
D Vinv v* vO iO vnet
Shaft
H
ωturb Tturb
Controller

Turbine

Figure 11.3 An overview of the complete model of a generator unit, with signal
and control interactions between each component

converter with buck and boost capability is required to keep the DC link voltage at
the inverter constant. A single-phase H-bridge inverter, with an LC filter to attenuate
harmonics, connects the GU onto the AC grid (single phase, 50 Hz, 240 VRMS).
The grid lines use standard Aluminium Conductor Steel Reinforced (ACSR)
lines. The main load on a typical rural grid is lighting, with some power used for
entertainment such as radios, televisions, computers and mobile phone charging.
With a higher supply capacity, induction machines for agricultural processing and
workshops are also used. Therefore, resistive-inductive and rectified, non-linear
loads are used in the modeling.

11.3.1 Turbine
The turbine model has two requirements. First, it must output the turbine torque for
a given head and speed input, and second, from it, the maximum power available at
the turbine must be able to be calculated. There are several models available in
literature for hydro turbines [9–12], or alternatively if there is detailed experimental
data, this could be used to determine the turbine performance.
For this work, a low-head Turgo turbine model, theoretically derived and
experimentally validated in [9] is used. Using this, the maximum currently avail-
able turbine power can be calculated for the measured turbine head. The available
torque and power characteristics as a function of the available head and turbine
rotational speed used are shown in Figure 11.4.
The rotational speed is derived from the shaft model and the head is specified
by the environmental conditions as an external input. The calculated turbine output
torque is fed forward into the shaft model.

11.3.2 Shaft assembly


The shaft assembly model describes the inertia of the turbine, generator rotor and
the shaft itself. It receives the mechanical torque generated in the turbine and the
electrical retarding torque from the generator, adds the damping torque from the
Modeling and simulation of a pico-hydropower off-grid network 229

1,500

1,000
Power (W)

500

0
0
100
200 3.5
3
300 2.5
2
400 1.5
1
Speed (RPM) Head (m)
(a)

150

100
Torque (Nm)

50

0
0
100
200 3.5
3
300 2.5
2
1.5
400 1
Speed (RPM) Head (m)
(b)

Figure 11.4 Theoretical (a) power and (b) torque models for a low-head Turgo
turbine with the head ranging from 1.0 to 3.5 m [9]
drive train bearings and other resistances, to calculate the acceleration of the shaft
according to the following equation:
 
dw 1 
¼ Tturb  Telec  Tdamp (11.1)
dt J
230 Modeling and dynamic behaviour of hydropower plants

where J is the rotational moment of inertia of the complete drive train, Telec is the
electrical torque from the generator, and Tdamp is the damping torque which is
proportional to the rotational speed. Using (11.1), the angular acceleration is inte-
grated to obtain the rotational speed.
The shaft inertia is assumed to be negligible, with a relatively small radius of
gyration compared to the turbine and generator; therefore, the inertia is assumed to
be the sum of the turbine and generator inertias. For the turbine, this inertia is
estimated to be 0.47 kg m2 [13].

11.3.3 Generator
The shaft model’s rotational speed output wturb is fed into the generator model
along with the current drawn Ig from the rectifier. The generator is modeled in the
dq rotating reference frame using the following equations [14]:
did
vd ¼ Ld þ Rs id  Lq pwturb iq (11.2)
dt
diq
vq ¼ Lq þ Rs iq  Ld pwturb id þ lpwturb (11.3)
dt
where vd and vq are the d- and q-axis voltages, Ld and Lq are the stator inductances
referred to the d and q axes, R is the stator resistance, id and iq are the d and q axis
currents, p is the number of pole pairs, and l is the permanent magnet flux. The
voltages are converted from the dq reference frame to the three-phase stationary
coordinate system, and the output voltage amplitude, VgRMS, is calculated and output
to the rectifier. The electromagnetic torque from the generator is calculated by [14]:
Telec ¼ 1:5pliq (11.4)
and is used in (11.1) to calculate the rotational speed of the shaft, wturb.
The machine to be modeled for this application is the PMGO-1,5K 1.5 kW
permanent magnet generator by DVE Technologies [15]. Table 11.1 summarizes
the pertinent technical details of the machine. When rotating between 5 and 42 rad/s
(50–400 RPM), the amplitude of the generator output voltage, |Vg|, ranges from
93 to 744 V.

Table 11.1 Generator technical data for simulation

Parameter Value
Output rated power 1.5 kW
Rated output speed 200 RPM
Number of pole pairs, p 9
Permanent magnet flux* l 1.14 V s
Resistance (phase) at 20  C* Rs 4.75 W
Inductance (d and q axes)* Ld, Lq 0.11 H
Inertia Jgen 0.535 kg m2

*Derived from datasheet.


Modeling and simulation of a pico-hydropower off-grid network 231

11.3.4 Rectifier
The three-phase rectifier is modeled as a lossless rectifier converting the AC output
from the generator to DC, and assuming unity power factor conversion. The rec-
tifier output voltage is calculated by [16]:
VDC ¼ 1:35  Vg;LL ¼ 2:34  Vg;RMS (11.5)
The current draw from the generator can be calculated by equating the input and
output power for the rectifier, such that:
3Vg;RMS Ig;RMS ¼ VDC IDC (11.6)
where VDC and IDC are the DC link voltage and current at the output of the rectifier,
respectively. The DC link voltage is applied to the DC–DC converter and the peak
current draw flows back to the generator.

11.3.5 DC–DC converter


The DC–DC converter is a buck–boost converter as the rectifier will produce an
output between 265 V and in excess of 1 kV, as a function of the speed of the
turbine and generator, and the inverter requires a constant DC link voltage of
400 V. A typical four-switch synchronous buck–boost converter topology is mod-
eled, as shown in Figure 11.5.
The converter is assumed to operate in continuous current mode. The modeling
equations for the converter have been derived using the state-space averaging
method [17]:
diL  
LDCDC ¼ D  2RON þ RLðDCDCÞ iL þ VDC
dt   (11.7)
þð1  DÞ  2RON þ RLðDCDCÞ iL  vC þ RC Iinv
dvC
CDCDC ¼ DðIinv Þ þ ð1  DÞðiL  Iinv Þ (11.8)
dt
Vinv ¼ DðvC  RC Iinv Þ þ ð1  DÞðRC iL þ vC  RC Iinv Þ (11.9)
IDC ¼ DðiL Þ (11.10)
where LDCDC is the inductance, iL is the current through the inductor, D is the duty
ratio, RON is the resistance of the switches in the on state, RL(DCDC) is the resistance

IDC Iinv

Q1 RON LDCDC RL(DCDC) Q4 RON

Q2 Q3 CDCDC
VDC Vinv
RON RON RC

Figure 11.5 Four-switch buck-boost converter


232 Modeling and dynamic behaviour of hydropower plants

Table 11.2 DC–DC converter technical data for simulation

Parameter Value
Inductance, LDCDC 10 mH
Inductor parasitic resistance, RL(DCDC) 0.05 W
Capacitance, CDCDC 100 mF
Capacitor equivalent series resistance, RC(DCDC) 0.1 W
Switch on-state resistance, RON 0.05 W

of the inductor, RC is the series resistance of the capacitor, CDCDC is the output
capacitor, vC is the voltage across the capacitor, and Vinv and Iinv are the output
voltage and current of the converter. The values for the components used in the
simulation are shown in Table 11.2.
It is assumed that there is no reverse recovery current or dead time requirement
in the switching. From this model, the output voltage of the DC–DC converter is
passed to the inverter, and the current IDC is passed back to the rectifier.

11.3.6 Inverter modeling


The inverter is modeled using an ideal controlled voltage source in SimPo-
werSystems. This assumes that the inverter has an infinite control bandwidth, so it
is able to follow the reference signal exactly, and any high frequency switching
noise is eliminated by the output filter. The inverter is assumed to have a maximum
output power of 1.5 kW, thus providing margin to allow for short power spikes. The
efficiency of the inverter is modeled using the equations and parameters derived in
[18], assuming a type 2 inverter which has a good efficiency characteristic over a
range of loads:
p
hinv ¼
p þ p0 þ kp2
where
Pinv;out

Pinv;rated
p0 ¼ 0:0072
k ¼ 0:0345 (11.11)
The voltage from the inverter, VRMS, is fed into the grid, whereas output cur-
rent demand, IRMS, is measured and divided by the inverter efficiency before being
applied to the DC–DC converter as the demand current Iinv.

11.3.7 Transmission line and load modeling


The ACSR grid lines are modeled as impedances. The chosen lines have a cross-
sectional area of 21 mm2, with a resistance of 1.41 W/km and reactance of 0.32 W/km
Modeling and simulation of a pico-hydropower off-grid network 233

when conductors are 0.3 m apart [19]. The linear load is modeled as an impedance
with a lagging power factor of 0.9, representing a loaded induction machine run-
ning with some resistive load [20]. The non-linear load is modeled with a diode
bridge rectifier and capacitor with a load resistor.

11.4 Control scheme design


11.4.1 Turbine and DC–DC converter controller design
From (11.1), it can be seen that excess turbine torque accelerates the drive train, and
conversely as more load is drawn from the GU, the drive train slows down. The power
curve for the turbine is similar to those shown in Figure 11.4(a). Beyond a certain
loading, where the turbine has slowed to its maximum power speed, wCRIT, further
load reduces the turbine power and the system becomes unstable and stalls. A control
system is required to restrict the turbine rotational speed exceeding wCRIT. This can be
implemented by measuring output power and speed from the generator and reacting if
both the power and rotational speed are decreasing simultaneously by limiting the
current output through the rectifier. The output voltage of the DC–DC converter, Vinv,
is regulated to 400 V by using a Proportional Integral (PI) compensator.

11.4.2 Inverter control design


The inverter control system is critical to ensure the operation of the off-grid network.
It must be able to regulate the voltage and frequency using only local measurements,
provide the plug-and-play capability for the system, allow good power sharing
between units based on the power available at the turbine and ensure a good power
quality of the output voltage. The following section describes the design of this
controller in detail, with an overview of the control system shown in Figure 11.6.
11.4.2.1 Voltage and frequency regulation via droop control
Droop control adjusts the output voltage and frequency of an inverter in order to
establish a desired relationship between supplied active and reactive power and
measured local voltage and frequency. The relationships depend on the line and
inverter output impedances [21]. Typically, transmission lines are considered to be
inductive, e.g., in [7]; however, the network considered here operates at low voltage
(240 VRMS), therefore resistive lines are assumed and the droop equations are [8]:
f ¼ f0 þ mQ (11.12)
VO ¼ VO;0  nP (11.13)
where f is the output frequency, f0 is the output frequency set point, m is the
reactive power droop coefficient, Q is the measured reactive output power, VO is
the output voltage reference, VO;0 is the output voltage set point, n is the active power
droop coefficient, and P is the measured active power. Although this relationship
is designed for a low-voltage, resistive-line network to be used with identical con-
trollers, it has also been shown to operate in conjunction with inductive- or capacitive-
line based droop control [22].
234 Modeling and dynamic behaviour of hydropower plants

P
θinv Phase
Droop P and Q
Head locked
function calculation
loop
Q
VO finv iOα iOβ vOα vOβ

Δf Reference Virtual αβ finv dq θnet


generator resistance
finv αβ
αβ
vref θinv iO vnetα vnetβ
+ vVR

vref* αβ fnet
vharm vO
V and I iLα
control αβ finv
iLβ vnet
loop

v* iL
DC Off-grid
link network
vinv

LC filter
Network switch

Figure 11.6 Control system structure

The droop coefficient is in the forward path of the controller; therefore, the
transient response of the control is dependent on droop coefficients, m and n [8].
Ideally, these should be large to have good transient response and also allow for
accurate power sharing, but this would cause a large variation in the regulated
voltage and frequency. The regulation is also typically defined in the system spe-
cification, fixing the droop coefficients. Therefore, to improve the transient
response of the control system, additional terms can be included. In this system, a
differential term is included in the droop equations, as described in [8], as integral
terms can cause instabilities with resistive line droop equations. Therefore, (11.12)
and (11.13) become:
dQ
f ¼ f0 þ mQ þ md (11.14)
dt
dP
VO ¼ VO;0  nP  nd (11.15)
dt
where md and nd are the differential constants. The reference voltage waveform is
then constructed from these values of output voltage and frequency.
Modeling and simulation of a pico-hydropower off-grid network 235

The impedance of the transmission lines in off-grid networks and the inverter
output impedance are neither purely inductive nor resistive which leads to cross
coupling in the droop function [21]. A virtual output impedance can be used to
force these to appear either inductive or resistive, dependent on the control scheme
selected and allow the droop function to be completely decoupled [8]. In this
scheme, as resistive lines are assumed, this is achieved by multiplying the funda-
mental component of the measured output current iO by a virtual resistance gain RV
and then subtracting this from the reference voltage calculated by the droop
function.

11.4.2.2 Power sharing


Standard droop control, utilizing steep droop coefficients, is able to share power
well with units that have equal rated power. However, when there are unequally
rated units on the system, then the power sharing is not proportional to the rated
output of the system, so smaller units will have a larger proportion of their output
supplied to the system. The droop coefficients and virtual resistance can be made
dependent upon the input power, which assists in achieving accurate power sharing
proportional to the unit’s available input power [23,24], which have been used for
UPS inverters. In the presented case, the input power is dependent on the head at
the turbine.
From this, a normalized turbine power output, the ratio of PTURB;MAX ðH Þ to the
maximum possible turbine output power PTURB;MAX , a fixed value, can be written as
PTURB;MAX ðHÞ
g¼ (11.16)
PTURB;MAX
Then, the droop coefficients and virtual resistance are set to:
mMAX
m¼ (11.17)
g
nMAX
n¼ (11.18)
g
RV ;MAX
RV ¼ (11.19)
g
where mMAX and nMAX are defined from the regulated range of frequency and
output voltage and the maximum output active and reactive powers, and RV ;MAX is
the maximum power virtual resistance. In this way, as g reduces the gradients of the
droop curves become steeper. So in the case of the P vs. VO droop curve, for the
same output voltage, the active power delivered is reduced.

11.4.2.3 Power measurement


The droop function, (11.14) and (11.15), needs the line-cycle-averaged active and
reactive powers to calculate the output voltage and frequency. In three-phase sys-
tems, this can be achieved by using the Clarke’s transform [25] to convert the line
236 Modeling and dynamic behaviour of hydropower plants

values into orthogonal a and b components. The instantaneous powers are calcu-
lated using the following equation:

vOa iOa þ vOb iOb
P¼ (11.20)
2

vOb iOa  vOa iOb
Q¼ (11.21)
2
which are used as inputs for the droop function. For single-phase systems, there are
several different methods to achieve the conversion between a single-phase sinu-
soidal signal and ab components, such as shifting the signal by 90 using a trans-
port delay [8], integrating the incoming signal [26] or using a resonant filter [27],
which is based on a second-order generalized integrator (SOGI) [28]. A modified
version of this SOGI-based filter is proposed in [29], where an additional gain is
included in the orthogonal ( b) path, and the gains are calculated using a Kalman
function. The method used in this approach has an identical structure to that pre-
sented in [29], but with constant gain values. The resonant frequency used in the
filter is calculated from the droop function, (11.3), and is fed into the filter,
allowing the filter to adapt to any variation in the grid frequency. The structure of
the SOGI-based filter is shown in Figure 11.7(a).
Additional resonant loops can be added to filter any harmonics in the input
[29], as shown in Figure 11.7(a), where a third harmonic loop is included.


vin + kα +– 1/s ++

ω vα
(From
Fundamental
droop
curve) kβ + 1/s
+ vβ
Fundamental SOGI vin vα
kα3 +– 1/s ++ αβ
ω vβ

3 vα3 (b)

3rd Harmonic
kβ3 ++ 1/s vβ3
3rd Harmonic SOGI
To other From other
harmonic loops harmonic loops

(a)

Figure 11.7 (a) Second-order generalized integrator (SOGI)–based filter with


additional harmonic loops. (b) Symbolic representation of SOGI-
based filter.
Modeling and simulation of a pico-hydropower off-grid network 237

The harmonics from each additional loop can be extracted and used if needed.
Figure 11.7(b) shows the symbol used for the SOGI-based filter in the following
sections.

11.4.2.4 Fundamental voltage controller


A second-order LC filter is assumed at the output of the inverter. Similar to the
control structure commonly employed in three-phase grid connected inverters and
active rectifiers [30,31], a synchronous reference frame controller is employed here
for both the outer output capacitor voltage and inner inductor current control loops.
A similar approach has been used in [32] where the outer voltage loop is trans-
formed into the synchronous frame, before being returned into the stationary frame
for the current control loop. This current loop is a proportional loop with a feed-
forward term, which reduces the need for a large gain to reduce the steady-state
error. Similar control strategies are proposed in [33–35], where the systems
described are either already in three-phase or use delays to create the orthogonal
component.
Here, the control scheme shown in Figure 11.8 is employed to control the
inverter output voltage. This is different to [32] as it uses the synchronous reference
frame throughout the control of voltage and current, allowing simple PI control to
be used in both and ensuring a zero steady state error. Karimi-Ghartemani [33] uses
the synchronous reference frame throughout the controller, but the presented con-
troller differs as the reference voltage generated from the droop function and
measured voltage and currents are transformed into the dq synchronous reference
frame using the SOGI-based filter. The ab to dq transform, and inverse, is syn-
chronized with the angle fed forward from the droop equations, (11.14), not
through a phase locked loop (PLL) loop as normally used. PI loops are then used to
force the measured d-axis voltage to its reference and the q-axis voltage to zero.

θinv θinv
vd,ref
vref* dq +– PI ++ +– PI ++ dq vα*
0v +– PI +– +– PI +–
q,ref

Cqω Lqω

Cdω Ldω
θinv θinv
vd,meas,fund id,meas,fund
dq dq
vq,meas,fund iq,meas,fund
Vmeas,fund Imeas,fund
Voltage loop Current loop
Key
αβ
dq = αβ dq

Figure 11.8 Proposed voltage and current controller using synchronous


reference frame
238 Modeling and dynamic behaviour of hydropower plants

0 + PI dq vαh*

0 + PI αβ

θinv,h

dq
vd,meas,h
αβ
vq,meas,h

vα,meas,h vβ,meas,h

Figure 11.9 Harmonic voltage controller in synchronous reference frame

11.4.2.5 Harmonic voltage controller


In single-phase applications, harmonic distortion compensation has been proposed
in [36] where band-pass virtual resistances are employed at the point of common
coupling, providing a power filter prior to the load. In three-phase grid connected
inverters and active power filters applications, harmonic compensation using a
series of synchronous frame harmonic voltage PI controllers, or proportional-
resonant controllers have been proposed in [30,31,37,38]. Similar to the three-
phase strategies, here the output voltage harmonics are suppressed using a series of
synchronous frame harmonic voltage PI controllers at the inverter output, as with
the fundamental voltage controller described above. As shown in Figure 11.9, the
measured harmonic output voltages, extracted from the SOGI-based filters in Fig-
ure 11.7, are transformed into the dq reference frame. Using a PI controller, the
measured harmonic voltages are forced to zero. These are transformed back into the
ab reference frame, and the a component becomes the reference waveform from
the harmonic. All the harmonic reference waveforms are summed to form the final
reference, which is output to the PWM generator.

11.4.2.6 Phase locked loop (PLL)


The standard way of synchronization is for the network voltage to be transformed
into the synchronous reference frame via the ab stationary reference frame, with
the network angle found by forcing the q-axis component to zero. Advanced PLL
strategies are presented in [28,39–43], using a synchronous reference frame or
digital non-linear methods. Here, a synchronous reference frame–based PLL is
used, with a single-phase input. The transformation of the voltage signal from a
single-phase signal to the ab reference frame can be achieved using the SOGI-
based filter again, as in [28] and analysed in [39], and shown in Figure 11.10.
The off-grid network angle qnet is then compared with the inverter angle qinv.
This error is fed to a proportional controller and then back into the droop function
to change the frequency of the inverter to match the grid frequency. Once the error
Modeling and simulation of a pico-hydropower off-grid network 239

θnet
vnet,α vnet,d ωnet,nom
vnet αβ
vnet,q ωnet θnet
αβ dq PI + 1/s
+
vnet,β
ωnet

Figure 11.10 SOGI-based phase-locked loop

S1 ZLINE1
S3
= ZLOAD1
= = =

Generator unit 1
S2
ZLINE2
S4
=
= = =

Generator unit 2 AC bus bar

ZLOAD2

Figure 11.11 Layout of two generator units used in simulations, with each unit
connected to the AC bus bar via an impedance, and a linear and
non-linear load connected to the bus

is below the critical value ecrit, the switch between the inverter and the grid is
closed with a latch to ensure that there is no chattering.

11.5 Simulation results

The simulations of a basic pico-hydropower off-grid network are carried out for a
pair of GUs connected to an AC bus bar, with linear and non-linear loads on the
bus, as shown in Figure 11.11, using the models and control described in the pre-
vious sections. These simulations will identify how the different systems react to
changes in load and environmental (head and flow) conditions.

11.5.1 Single generator unit with varying load


The first simulation is to show the response of the system to a changing load. Using
the layout in Figure 11.11, only GU 1 is used, with a load that changes from 500 to
750 VA and then back down to 250 VA, all at a power factor of 0.9. Initially, S1 and
S2 are closed. The results from this simulation are shown in Figure 11.12.
At t ¼ 0, there is no load on the inverter; therefore, the turbine is at free-
wheeling speed. When the load is added at t ¼ 0 the rotating components slow,
240 Modeling and dynamic behaviour of hydropower plants

50

speed (rad/s)
Rotational
40

30

20

410
output voltage (V)
DC–DC converter

400

390

750 Demand
power (W)

500 Actual
Active

250

500
power (Var)

Demand
Reactive

250 Actual

250
voltage (VRMS)
Inverter

225

200
0 2.5 5 7.5 10 12.5 15
Time (s)

Figure 11.12 Response of a single generator unit to a varying load (500, 750 and
250 VA) with turbine/generator rotational speed, DC–DC converter
output voltage, active power, reactive power and inverter output
voltage RMS

but the inertia of the turbine and generator cause the deceleration to last 2 s. This
constant slowing over 2 s means that the input voltage to the rectifier, and therefore
the DC–DC converter, is not constant and is slowly reducing, although it can be
seen that DC–DC converter is able to maintain a constant output voltage once it has
achieved its reference value, even during this input voltage change. As the power
demand increases at t ¼ 5 s, the voltage and frequency droop and the speed of
the rotating components drop further to match the turbine output power with the
Modeling and simulation of a pico-hydropower off-grid network 241

demand from the inverter. Although there is a small oscillation in the DC–DC
converter output voltage and inverter output voltage, these return to a steady state
quickly. At t ¼ 10 s, the power demand reduces again, so the rotational speed
increases, as does the output voltage from the inverter, while the DC–DC converter
output voltage remains constant after a small oscillation.
The output power is slightly lower than the demand during the periods of high
demand due to the droop from the control system, and the further reduction
in inverter output voltage due to the virtual resistance, with the power demand
calculated on the nominal power output from the inverter.

11.5.2 Performance with non-linear load


GU 1 is then connected to a non-linear load, a diode rectifier with a 3,300 mF
smoothing capacitor and a 100 W resistor, with S1 and S4 closed. The voltage
waveforms obtained for this simulation are shown in Figure 11.13.
There are some harmonics in the output voltage waveform, due to the current
harmonics drawn from the load. The error has a high frequency and irregular

500
Voltage (V)

–500
0 1 2 3 4 5
(a) Time (s)
400
Voltage (V)

–400
4.98 4.985 4.99 4.995 5
(b) Time (s)
15
Error (V)

–15
4.98 4.985 4.99 4.995 5
(c) Time (s)

Output voltage Voltage reference Error

Figure 11.13 (a) Inverter output voltage tracking with a non-linear load (diode
rectifier with 3,300 mF capacitor and 100 W resistor), complete
voltage profile over 5 s from inverter switch on. (b) Voltage
reference vs. inverter output voltage. (c) The error between the
reference and output voltage.
242 Modeling and dynamic behaviour of hydropower plants

oscillatory nature. The peak amplitude of this error is approximately 11 V which
is approximately 3.6% of the peak voltage.

11.5.3 Power sharing performance


The following two simulations are for two GUs connected in parallel. The first
simulation has two GUs with equal line impedance between the GU and the AC bus
bar, with switches S1  S3 in Figure 11.11 closed. The load starts at 500 VA,
increasing to 1,500 VA at t ¼ 5 s, and then reduces to 1,000 VA at t ¼ 10 s, all with
a power factor of 0.9. The results for this can be seen in Figure 11.14.
As can be seen, the two systems share the load equally, delivering an equal
current to the load and the turbine rotational speed identical. When the line impe-
dances are not equal, with ZLINE2 twice size of ZLINE1, the results of the simulation
are shown in Figure 11.15.
In this simulation, as GU 1 is closer to the load, it takes a slightly higher
proportion of the active power demand; therefore, the rotational speed of the tur-
bine and generator is slower. As the reactive power is dependent on the frequency,
both GUs share this equally. Increasing the load after 5 s causes the power sharing
difference between the two units to rise, increasing the imbalance in supplied load
current from each unit.

11.5.4 Change in input power (drop in head)


In this simulation, the head at GU 1 remains constant at 3 m, but the head at GU 2
starts at 3.5 m, before reducing linearly to 2 m over a 5 s period. The line impe-
dances to the AC bus bar are not equal, GU 1 has a line impedance half that of GU 2.
The load remains constant at 500 VA. The results for this simulation are shown in
Figure 11.16.
As the head reduces at GU 2, GU 1 provides more of the power demand. The
DC–DC converter is able to compensate for the drop in generator voltage, and its
control keeps the output at the regulated value. These results show that the system
remains stable and is able to cope with a change in head at one unit.

11.6 Modeling of implementation in Nepal


The system is envisioned to operate in many locations all over the world where
there is no electricity supply from the national power grid and suitable water
resources, in both developed and developing countries. An example implementa-
tion of the off-grid pico-hydro network is shown in the following simulation.
The example implementation environment is located in Bhanbhane district,
Gulmi, Western Nepal. A map of the area is shown in Figure 11.17 with the layout.
The Patan River has five potential pico-hydro sites along it, which are detailed in
Table 11.3.
At site 3, there is 5 m of available head. As the low head, Turgo turbine in [9]
is only rated to 3.5 m head, an additional GU is added to make use of the remaining
1.5 m of head. The model and control that has been developed in the previous
Modeling and simulation of a pico-hydropower off-grid network 243

50

speed (rad/s)
Rotational
40

30

20
output voltage (V)

410
DC–DC converter

400

390

1,500
power (W)
Active

1,000

500

750
power (Var)
Reactive

500

250

0
0 2.5 5 7.5 10 12.5 15
Time (s)
400 3
voltage (V)

current (A)
Inverter

Inverter

0 0

–400 –3
4.5 4.505 4.51 4.515 4.52 4.5 4.505 4.51 4.515 4.52
Time (s) Time (s)
Inverter 1 Inverter 2 Demand Output power total

Figure 11.14 Simulation of two generator units with equal line impedance to an
AC bus bar feeding a load varying from 500 VA at t ¼ 0 s to 1,500
VA at t ¼ 10 s, then reducing to 1,000 VA at t ¼ 10 s

sections is then used to develop the grid. For this modeling of the implementation
site therefore, six GUs are connected to the network with ACSR transmission lines
and two linear 1 kVA, 0.9 p.f. loads at the load centre. Initially, GU 2–6 are con-
nected to the network, with a 1 kVA load. After 10 s, GU 1 is connected, and after
15 s, the second 1 kVA load is also connected. After 20 s, the head at GU 2 reduces
from 3.5 to 1.5 m, representing an instantaneous change in environmental condi-
tions, such as a blockage of the intake. The results from this simulation are shown
in Figure 11.18.
244 Modeling and dynamic behaviour of hydropower plants

50
speed (rad/s)
Rotational
40

30

20
output voltage (V)
DC–DC converter

410

400

390

1,500
power (W)
Active

1,000

500

750
power (Var)
Reactive

500

250

0
0 2.5 5 7.5 10 12.5 15
Time (s)
400 5
voltage (V)

current (A)
Inverter

Inverter

0 0

–400 –5
7.5 7.505 7.51 7.515 7.52 7.5 7.505 7.51 7.515 7.52
Time (s) Time (s)
Inverter 1 Inverter 2 Demand Output power total

Figure 11.15 Simulation of two generator units with unequal line impedance
(ZLINE2 ¼ 2  ZLINE1) to an AC bus bar feeding a load varying from
500 VA at t ¼ 0 s to 1,500 VA at t ¼ 10 s, then reducing to 1,000 VA
at t ¼ 10 s

As can be seen in Figure 11.18, once the units have reached a steady state in
current and speed, GU 1 is added. This causes the load on all the other GUs to drop,
with the current decreasing, and therefore, the turbine rotational speed increases. At
this point, GU 2 is closest to the load and has the maximum head; therefore, it
supplies the most current to the load. Conversely, GU 4 has the lowest current output
as it has the lowest head. When the head at GU 2 drops from 3.5 to 1.5 m, the current
output in all the other GUs increases, whereas the GU 2 reduces from 2 to 1 ARMS.
Modeling and simulation of a pico-hydropower off-grid network 245

Head (m)
2

50
speed (rad/s)
Rotational

40

30

20

410
output voltage (V)
DC–DC converter

400

390

500
power (W)
Active

250

250
power (Var)
Reactive

125

0
0 2.5 5 7.5 10 12.5 15
Time (s)
Inverter 1 Inverter 2 Demand Output power total

Figure 11.16 Simulation of two generator units with a head change from 3.5 to
2 m on generator unit 2 with a varying load and unequal line
lengths

11.7 Hybrid renewable off-grid network


An extension of the off-grid pico-hydropower grid is to hybridize it with other
renewable technologies such as solar PV or small scale wind turbines, as shown in
Figure 11.19.
The same inverter interface can be used to control each of these sources, with
minor changes to the rest of the hardware and control algorithm.
246 Modeling and dynamic behaviour of hydropower plants

Pico-hydro site

Power lines

Household/business
400 m 1
300 m
2
3
200 m
400 m
4

600 m

Map © CNES/Astrium, Google

Figure 11.17 Potential pico-hydro off-grid network on the Patan Khola,


Bhanbhane District, Western Nepal

Table 11.3 Details of the possible site locations in Bhanbhane District, Western
Nepal (minimum flow and head data)

Site Name Min. Head Latitude Longitude Minimum


no. flow (l/s) (m) power
output (kVA)
1 Badachaur Pokherel 35 3.3 28 60 4700 83 60 1500 0.69
2 Hadhade 35 3.5 28 60 4200 83 60 3000 0.74
3 Badachaur 35 5.0 28 60 3900 83 60 3700 1.05
4 Bharji 35 2.5 28 60 4400 83 60 4900 0.53
5 Badachaur School 35 2.5 28 60 5500 83 70 900 0.53
Total power 3.54

11.7.1 Solar PV interface modifications


The solar PV array outputs DC power, so the rectifier stage is not required in the
power conversion. The array is assumed to have a 1 kW p output, with a maximum
power point tracking algorithm. The power ratio that was previously used to modify
the droop coefficients, g, for the solar PV is calculated by the following equation:
PPV;MPPT
gPV ¼ (11.22)
PPV;MAX
where gPV is the PV power ratio, PPV;MPPT is the measured maximum power from
the PV array, and PPV;MAX is the maximum rated power from the PV array.
Modeling and simulation of a pico-hydropower off-grid network 247

50

Turbine speed
(rad/s)
25

0
420
voltage (V)
DC link

400

380
3
current (A)
Grid RMS

0
260
voltage (V)
Grid RMS

240

220
0 5 10 15 20 25
Time (s)
GU 1 GU 3 GU 5
GU 2 GU 4 GU 6

Figure 11.18 A six unit pico-hydro off-grid network operating with different
heads. After 10 s, one unit is connected to the network, and after
15 s, an additional load is added.

11.7.2 Wind turbine interface modifications


The wind turbine is assumed to feed directly into the rectifier, as with the pico-
hydro turbine. An alternative would be to include some energy storage on the
rectified DC stage, which would smooth the power generated from gusts; however,
this is not included for this simulation. The turbine is assumed to have a maximum
248 Modeling and dynamic behaviour of hydropower plants

electrical power output of 1 kW. As with the solar PV array, the power ratio to
modify the droop coefficients becomes:
PWD;MEAS
gWD ¼ (11.23)
PWD;MAX
where gWD is the wind turbine power ratio, PWD;MEAS is the measured maximum
power from the wind turbine, and PWD;MAX is the maximum rated electrical power
from the turbine.

11.7.3 Hybrid grid simulation


The hybrid grid is assumed to be as drawn in Figure 11.19, with a single pico-hydro
turbine, wind turbine and solar PV array connected to it. Each source is assumed to
be 1 km radially from the load. The solar panels are assumed to be five parallel
connected PV modules and modeled using a standard PV model [44]. The wind
turbine is modeled as a locally manufactured Piggott Turbine [45] using the model
in [46]. The pico-hydro turbine has a constant head of 3.5 m. The irradiance of the
PV array starts at 500 W/m2 and increases to 1,000 W/m2 after 10 s, simulating the
panel moving out of shade. The wind turbine sees a Gaussian distribution of wind
speed of with a mean of 10 m/s and variance of 3 m/s. A 500 VA 0.9 p.f. load is
connected after 5 s, then after 15 s the load increases to 1,500 VA.
As can be seen from Figure 11.20, the inverter interfaces operate similarly to
the pico-hydropower network, with each system feeding in dependent on its
available power. Therefore, initially with the solar PV able to supply half its rated
power, it supplies approximately half the current of the hydro turbine. As the wind
turbine power is constantly varying, so does the current input into the network. The
solar PV and hydro are able to support this varying current. If increased energy

= = G
= = = = =
Solar PV panel Pico-hydro turbine T

Load

G =
= = =
Wind turbine

Figure 11.19 Hybrid renewable off-grid network with pico-hydropower, solar PV


and wind turbine
Modeling and simulation of a pico-hydropower off-grid network 249

50 40
40
Turbine speed
30

voltage (V)
PV array
30
(rad/s) 20
20
10 10

0 0
4

3
current (A)
Grid RMS

0
300
voltage (V)
Grid RMS

150

0
0 5 10 15 20
Time (s)
Pico-hydro Wind turbine Solar PV

Figure 11.20 Turbine speed, PV voltage and output currents and voltage from
hybrid off-grid network

storage was added to the wind turbine system, this could become more constant,
and smoothing out the supply from the wind turbine.

11.8 Summary
This chapter has introduced the concept of the pico-hydropower off-grid network,
shown the proposed control design for the power electronic interface and developed
the models of the system to show the performance. The system is simulated initially
with a single GU, demonstrating the voltage droop, and the performance of the
system when connected to a non-linear load. Following this, two GUs are simu-
lated, showing how the power sharing alters as the distance between the generators
and load becomes unequal and how the input power to the turbine, calculated from
the head, the power-sharing ratio, also changes. This simulation is then expanded to
cover an example implementation site with multiple turbines, showing similar
results in a changing environment. Finally, the pico-hydropower off-grid network
250 Modeling and dynamic behaviour of hydropower plants

concept is extended to include additional renewable resources of solar PV and wind


turbines, with minor control modifications, and a simulation demonstrates how this
could operate with changing loads and changing input powers.

References
[1] International Energy Agency, World Energy Outlook, London: Organization
for Economic Co-operation and Development, 2014.
[2] Smith N. P. A., ‘‘Induction Generators for Stand-Alone Micro-Hydro
Systems,’’ Proceedings of International Conference on Power Electronics,
Drives and Energy Systems for Industrial Growth, New Delhi, 1996,
pp. 669–673.
[3] Maher P., The Pico Power Pack: Fabrication and Assembly Instructions
[online], May 2001. Available from www.picohydro.org.uk.
[4] Taylor S. D., Fuentes M., Green J. and Rai K., Stimulating the Market for
Pico-hydro in Ecuador, Department for International Development Report
R8150, London, 2003.
[5] Williams A. A. and Simpson R., ‘‘Pico hydro – reducing technical risks for
rural electrification,’’ Renewable Energy, vol. 34, pp. 1985–1991, 2009.
[6] Energy Sector Management Assistance Program, Technical and Economic
Assessment of Off-grid, Mini-grid and Grid Electrification Technologies,
World Bank Report, New York, 2007.
[7] Strauss P. and Angler A., ‘‘AC Coupled PV Hybrid Systems and Microgrids –
State of the Art and Future Trends,’’ Proceedings of Third World Conference
on Photovoltaic Energy Conversion, Osaka, 2003, pp. 2129–2134.
[8] Guerrero J. M., Matas J., de Vicuna L. G., Castilla M. and Miret J.,
‘‘Decentralized control for parallel operation of distributed generation
inverters using resistive output impedance,’’ IEEE Transactions on Industrial
Electronics, vol. 54, no. 2, pp. 994–1004, 2007.
[9] Williamson S. J., Stark B. H. and Booker J. D., ‘‘Performance of a low-head
pico-hydro Turgo turbine,’’ Applied Energy, vol. 102, pp. 1114–1126, 2013.
[10] Márqueza J. L., Molinab M. G. and Pacasc J. M., ‘‘Dynamic modeling,
simulation and control design of an advanced micro-hydro power plant for
distributed generation applications,’’ International Journal of Hydrogen
Energy, vol. 35, pp. 5772–5777, 2010.
[11] Muller G. and Senior J., ‘‘Simplified theory of Archimedean screws,’’
Journal of Hydraulic Research, vol. 47, pp. 666–669, 2009.
[12] Williamson S. J., Stark B. H. and Booker J. D., ‘‘Low head pico hydro
turbine selection using a multi-criteria analysis,’’ Renewable Energy,
vol. 61, pp. 43–50, 2014.
[13] Williamson S. J., Stark B. H. and Booker J. D., ‘‘Modelling of a Multi-
Source Low-Head Pico Hydropower Off-Grid Network,’’ Proceedings
of IEEE International Conference on Sustainable Energy Technologies,
Kathmandu, 2012, pp. 369–374.
Modeling and simulation of a pico-hydropower off-grid network 251

[14] Herold T., Franck D., Lange E. and Hameyer K., ‘‘Extension of a D-Q
Model of a Permanent Magnet Excited Synchronous Machine by Including
Saturation, Cross-Coupling and Slotting Effects,’’ Proceedings of IEEE
International Electric Machines & Drives Conference, Niagara Falls, 2011,
pp. 1363–1367.
[15] DVE generator PMGO–1,5K–200. Available from http://www.dvetech.dk.
[16] Mohan N., Undeland T. M and Robbins W. P., Power Electronics: Con-
verters, Applications, and Design, New York: John Wiley & Sons, 2003.
[17] Erikson R. W. and Maksimovic D., Fundamentals of Power Electronics,
New York: Springer-Verlag, 1998.
[18] Notton G., Lazarov V., Stoyanov L. and Heraud N., ‘‘Grid–Connected
Photovoltaic System: Optimization of the Inverter Size Using an Energy
Approach,’’ Proceedings of International Symposium on Advanced Elec-
tromechanical Motion Systems & Electric Drives, Lille, 2009.
[19] ESMAP/UNDP, Mini Grid Design Manual, Washington, DC: ESMAP,
2000.
[20] Liang X. and Ilochonwu O., ‘‘Induction motor starting in practical industrial
applications,’’ IEEE Transactions on Industry Applications, vol. 47, no. 1,
pp. 271–280, 2011.
[21] Monfared M., Golestan S. and Guerrero J. M., ‘‘Analysis, design, and
experimental verification of a synchronous reference frame voltage control
for single-phase inverters,’’ IEEE Transactions on Industrial Electronics,
vol. 61, pp. 258–269, 2014.
[22] Zhong Q. C. and Zeng Y., ‘‘Parallel Operation of Inverters with Different
Types of Output Impedance,’’ Proceedings of IEEE Industrial Electronics
Conference, Vienna, 2013, pp. 1398–1403.
[23] Zhong Q. C., ‘‘Robust droop controller for accurate proportional load shar-
ing among inverters operated in parallel,’’ IEEE Transactions on Industrial
Electronics, vol. 60, pp. 1281–1290, 2013.
[24] de Brabandere K., Bolsens B., Van den Keybus J., Woyte A., Driesen J.
and Belmans R., ‘‘A voltage and frequency droop control method for
parallel inverters,’’ IEEE Transactions on Power Electronics, vol. 22,
pp. 1107–1115, 2007.
[25] E. Clarke, Circuit Analysis of AC Power Systems, vol. I, New York: Wiley,
1943.
[26] Roshan A., Burgos R., Baisden A. C., Wang F. and Boroyevich D., ‘‘A D-Q
Frame Controller for a Full-Bridge Single Phase Inverter Used in Small
Distributed Power Generation Systems,’’ Proceedings of IEEE Applied
Power Electronics Conference, Anaheim, 2007, pp. 641–647.
[27] Burger B. and Engler A., ‘‘Fast Signal Conditioning in Single Phase Sys-
tems,’’ Proceedings of European Power Electronics and Drives Conference,
Graz, 2001.
[28] Ciobotaru M., Teodorescu R. and Blaabjerg F., ‘‘A New Single-Phase PLL
Structure Based on Second Order Generalised Integrator,’’ Proceedings of
IEEE Power Electronics Specialists Conference, Jeju, 2006, pp. 1–6.
252 Modeling and dynamic behaviour of hydropower plants

[29] de Brabandere K., Loix T., Engelen K., et al., ‘‘Design and Operation of a
Phase-Locked Loop with Kalman Estimator-Based Filter for Single-Phase
Applications,’’ Proceedings of IEEE Conference on Industrial Electronics,
Paris, 2006, pp. 525–530.
[30] Lascu C., Asiminoaei L., Boldea I. and Blaabjerg F., ‘‘Frequency response
analysis of current controllers for selective harmonic compensation in
active power filters,’’ IEEE Transactions on Industrial Electronics, vol. 56,
pp. 337–347, 2009.
[31] Zhang R., Cardinal M., Szczesny P. and Dame M., ‘‘A Grid Simulator
with Control of Single-Phase Power Converters in D-Q Rotating Frame,’’
Proceedings of IEEE Annual Power Electronics Specialists Conference,
Cairns, 2002, pp. 1431–1436.
[32] Micallef A., Apap M., Spiteri-Staines C., Guerrero J. M. and Vasquez J. C.,
‘‘Reactive power sharing and voltage harmonic distortion compensation of
droop controlled single phase islanded microgrids,’’ IEEE Transactions on
Smart Grid, vol. 5, pp. 1149–1158, 2014.
[33] Karimi-Ghartemani M., ‘‘Universal integrated synchronization and control
for single phase DC/AC converters,’’ IEEE Transactions on Power Electro-
nics, vol. 30, pp. 1544–1557, 2015.
[34] Bahrani B., Saeedifard M., Karimi A. and Rufer A., ‘‘A multivariable design
methodology for voltage control of a single-DG-unit microgrid,’’ IEEE
Transactions on Industrial Informatics, vol. 9, pp. 589–599, 2013.
[35] Bahrani B., Rufer A., Kenzelmann S. and Lopes L., ‘‘Vector control of
single-phase voltage source converters based on fictive axis emulation,’’
IEEE Transactions on Industrial Applications, vol. 47, pp. 831–840, 2011.
[36] Marwali M. N., Jung J. and Keyhani A., ‘‘Control of distributed generation
systems – Part II: Load sharing control,’’ IEEE Transactions on Power
Electronics, vol. 19, pp. 1551–1531, 2004.
[37] Mattavelli P., ‘‘A closed-loop selective harmonic compensation for active
filters,’’ IEEE Transactions on Industrial Applications, vol. 37, pp. 81–89,
2001.
[38] Xiaoming Y., Merk W., Stemmler H. and Allmeling J., ‘‘Stationary-frame
generalized integrators for current control of active power filters with zero
steady-state error for current harmonics of concern under unbalanced and
distorted operating conditions,’’ IEEE Transactions on Industrial Applica-
tions, vol. 38, pp. 523–532, 2002.
[39] Golestan S., Monfared M., Freijedo F. D. and Guerrero J. M., ‘‘Dynamics
assessment of advanced single-phase PLL structures,’’ IEEE Transactions on
Industrial Electronics, vol. 60, pp. 2167–2177, 2013.
[40] Golestan S., Monfared M., Freijedo F. D. and Guerrero J. M., ‘‘Advantages
and challenges of a type-3 PLL,’’ IEEE Transactions on Power Electronics,
vol. 28, pp. 4985–4997, 2013.
[41] Geng H., Sun J., Xiao S. and Yang G., ‘‘Modeling and implementation of an
all digital phase-locked-loop for grid-voltage phase detection,’’ IEEE
Transactions on Industrial Informatics, vol. 9, pp. 772–780, 2013.
Modeling and simulation of a pico-hydropower off-grid network 253

[42] Rodriguez P., Luna A., Candela I., Mujal R., Teodorescu R. and Blaabjerg
F., ‘‘Multiresonant frequency-locked loop for grid synchronization of power
converters under distorted grid conditions,’’ IEEE Transactions on Industrial
Electronics, vol. 58, pp. 127–138, 2011.
[43] Guerrero J. M., Vasquez J. C., Matas J., de Vicuna L. G. and Castilla M.,
‘‘Hierarchical control of droop-controlled AC and DC microgrids – a general
approach toward standardization,’’ IEEE Transactions on Industrial Elec-
tronics, vol. 58, pp. 158–172, 2011.
[44] Villalva M. G., Gazoli J. R. and Filho E. R., ‘‘Comprehensive approach to
modeling and simulation of photovoltaic arrays,’’ IEEE Transactions on
Power Electronics, vol. 24, pp. 1198–1208, 2009.
[45] Monteiro J. P., Silvestre M. R., Piggott H. and Andre J. C., ‘‘Wind tunnel
testing of a horizontal axis wind turbine rotor and comparison with simula-
tions from two Blade Element Momentum codes,’’ Journal of Wind Engi-
neering and Industrial Aerodynamics, vol. 123, part A, pp. 99–106, 2013.
[46] Bolte E. and Landwehr M., ‘‘Mathematical Model of Small Wind Turbines,’’
Proceedings of Ninth International Conference on Ecological Vehicles and
Renewable Energies, pp. 1–6, 2014.

Further Reading
Jenkins N., Ekanayake J. B. and Strbac G., Distributed Generation, London: IET,
2010.
Zhong Q. C. and Hornik T., Control of Power Inverters in Renewable Energy and
Smart Grid Integration, Chichester: John Wiley & Sons Ltd, 2013.
Harvey A., Brown A., Hettiarachi P. and Inversin A., Micro Hydro Design Manual:
A Guide to Small Scale Water Power Schemes, Rugby: Practical Action
Publishing, 1993.
Practical Action, Poor People’s Energy Outlook, Rugby: Practical Action Pub-
lishing, 2010–2014.
Williamson S. J, Griffo A., Stark B. H. and Booker J. D., ‘‘A controller for single-
phase parallel inverters in a variable-head pico-hydropower off-grid net-
work,’’ Sustainable Energy, Grids and Networks, vol. 5, pp. 114–124, 2016.
Index

actuator solicitation rate 120 damped oscillation 82, 87–8, 92, 97


Agency for the Cooperation of Energy data acquisition (DAQ) hardware 4–7
Regulators (ACER) 106, 109 multiplexing 7
alternating current (AC) 11 range 7
Aluminium Conductor Steel resolution 7
Reinforced (ACSR) 228, 232, sampling rate 6–7
243 deflector movement 11
analog-to-digital converter (ADC) deterministic hydro generation
5–7, 136 scheduling (D-HGS) 161–2,
Aparados da Serra 205 173
Apple Macintosh 7 cascade D-HGS formulation 177–8
automatic generation control 37–40 deterministic mid-short term
scheduling planning 183
biconcave formulation 177–8 dewatering 120
Bode diagram 76, 122–3 diesel generators 208, 210, 214, 217,
branch-and-bound (B&B) algorithm 225
164, 180 digital-to-analog converter (DAC) 7
branch junction 22, 25–6 direct current (DC) 11
Brazil, Guarita Hydro Power Plant in DC–DC converter 231–2
187–202 discharge limits 175–6
distributed AC grid topology with
cascade D-HGS formulation 177–8 non-communicating inverter
Clarke’s transform 235 front-ends 227
closed-loop system, turbine governing distribution system operators (DSOs)
system as 81–2 109
‘commonly used block’ library 3 diverging oscillation 87–8, 92, 97
Compact PCI 4 downstream reservoir with constant
concatenation 68 water level 24–5
constant water level droop coefficients 233–5, 246, 248
downstream reservoir with 24–5 droop control 233, 235
upstream reservoir with 23–4 DVE Technologies 230
continuity equation 22, 25, 27, 82, 84 dynamic behaviour, analysis of 69
controlled plant 81 fast dynamics, decomposition
corporate simulator 119 of 72
Courant–Friederichs–Lewy criterion primary frequency control,
52 performance limitation for 74
256 Modeling and dynamic behaviour of hydropower plants

penstock water start time fluctuation stability 98–9


criterion 75–6 fourth-order polynomials 164, 169
surge tank cross-section criterion Francis turbine 27–8, 31, 53, 55, 70
76–7 hydropower unit with 27
slow dynamics, decomposition of governor system 28–31
72–4 turbine and generator 27–8
dynamic processes of HPP, case study model of 28
of 32 frequency containment reserves
emergency stop and load rejection (FCR) 110, 112
43–5 frequency control
grid-connected operation 35 grid codes requirements for
automatic generation control 106–12
37–40 and turbine governing systems
primary frequency control (PFC) specifications 114–18
35–7 frequency restoration reserves (FRR)
isolated operation 40–3 111
start-up and no-load operation 33–5 frequency sensitive mode (FSM)
control 112–13
electric equivalent circuit 57 friction constants, calculating 3
electricity balancing network code fundamental voltage controller 237
(NC EB) 110 Furnas hydropower plant, auxiliary
electro-hydraulic (EH) converter 11 functions for 169
electro-mechanical subsystem 67–8
ellipsoid algorithm 163 gas circuit breaker (GCB) 17, 150–1,
Emborcação hydropower plant, 153
auxiliary functions for 170 generator model 230
emergency stop and load rejection 43–5 geometric functions
equations of the model of hydropower physical properties of 165
units 28 convexity and concavity 166–7
European Network of Transmission increasing property 165
System Operators for special cases of 167–70
Electricity (ENTSO-E) 106–7 global optimization approach 177–8
energy net generation 109 computational results 180–3
members 108 governor 81, 99
net generating capacity 108 mathematical model of 86
network codes 109–10 governor system model 27–31
projections of intermittent installed block diagram 29
capacities 109 grid codes requirements, for frequency
RES generation 110 control 106–12
excitation system, elements of 11–14 grid connected hydropower plants,
reduced order models for 49
fast dynamics, decomposition of 72 complete state-space model for a
feedback effect 82 hydro plant connected to a
five-to-two-needle operation 156 grid 67
transition from 155 concatenation 68
Index 257

electro-mechanical subsystem harmonic voltage controller 238


67–8 headrace tunnel 80, 82–4, 99
hydro-mechanical subsystem 67 head sensitive discharge limits 176,
interconnected operation 68–9 178
isolated operation 69 high-frequency vibration 132, 139
dynamic behaviour, analysis of 69 high-order model of HPPs 21
fast dynamics, decomposition of hill chart 10, 12
72 HOMER 187, 193–6, 205, 207–8, 217
penstock water start time simulations with 211–12
criterion 75–6 Hurwitz criterion 90–1, 92, 94–5
slow dynamics, decomposition hybrid renewable off-grid network
of 72–4 245
surge tank cross-section criterion hybrid grid simulation 248–9
76–7 solar PV interface modifications
hydropower plant model 50 246
hydraulic circuit model 56–9 wind turbine interface
hydro-mechanical model of the modifications 247–8
power plant 61 hydraulic circuit model 56
mechanical model of the linearized model 58–9
generating unit 59–61 non-linear model in per unit 58
penstock and tunnel models 51–2 hydraulic component, mathematical
surge tank model 52–5 differential equations of 9
synchronous power system models hydraulic–electric analogy
61 for impulse turbine 50
general model 62–4 for reaction turbine 56
model for an interconnected grid hydraulic–mechanical–electrical
64–6 coupling system in HPPs 27
model for an isolated grid 66–7 hydraulic model of hydropower plant
grid-connected operation 35, 46 with two penstocks 13
automatic generation control 37–40 hydraulic turbine
primary frequency control (PFC) characteristic curves of 49–50, 53,
35–7 67
Guarita hydroelectric power plant 187 operating point of 10
in Brazil 188–90 static characteristic relationships of
components of PV hydro hybrid 10
system 191–3 hydroelectric power plant (HPP) 9, 49,
ecological flow rate in 189 144, 163, 187–8, 191, 193–4,
hydraulic turbine and electric 200, 202, 206, 211
machine 190 hydro generation scheduling 161
location (satellite view) of 192 cascade D-HGS formulation 177–8
parameters of 194 global optimization approach 178
solar energy availability 196 computational results 180–3
use of residual flow of 190–1 hydropower generation function
guide vane opening (GVO) 34, 57, 164
81, 84 mathematical properties 170–5
258 Modeling and dynamic behaviour of hydropower plants

physical properties of geometric instruction set architecture (ISA) 4


functions 165–7 integrated circuits (IC) sensors 5
special cases of geometric interconnected grid, model for 64–6
functions 167–70 inverter control design 233
water conservation and discharge fundamental voltage controller
limits 175 237
head sensitive discharge limits harmonic voltage controller 238
176 phase locked loop (PLL) 238–9
hydro generator stator’s steel structures power measurement 235–7
vibration sensors measuring power sharing 235
vibration of 133, 136–40 voltage and frequency regulation via
hydro generator testing 132 droop control 233–5
hydro-mechanical model of power inverter modeling 232
plant 61 island mode operation, in hydropower
hydro-mechanical subsystem 67 plant 149
hydropower plant (HPP), 3 measures to improve 157–8
dynamic simulation issues for performance 150–7
105–27 isolated grid, model for 66–7
European network example
grid codes requirements for Kalman function 236
frequency control and Kaplan turbine 53, 55
balancing 106–12 Karush–Kuhn–Tucker conditions 174,
French EDF experience 177
application for turbine governing Kirchhoff’s law 56, 63
systems 112–27
hydraulic model of 13 Laboratory Virtual Instrumentation
pipeline and power generating Engineering Workbench
system of 80 (LabVIEW) 4, 6–8
signal analysis methods for vibration Laplace equations 9
control at 131–44 Laplace transfer function 9
and turbine governing system laser timer 136
79–100 Legacy 193, 211
vibration control methodology linear simulation models 119–20
133–5 Linha Sete pumped storage power
vibration diagnostics case study plant 207–8
136–44 Linux 7
with/without surge tank 79–80 load frequency control (LFC) 39, 74,
see also run-off-type hydropower 79–80, 110
plant dynamic hierarchy of 111
hydro turbine governors 11 load frequency control and reserves
network code (LFCR NC)
IEEE1394 4 110–11
impulse turbine 53–5 Longyangxia Dam 206
hydraulic–electric analogy for 50 Loop Transfer 75–6, 118
indefinite Hessian matrix 163 low-frequency vibration 132
Index 259

Mac OS X 7 off-grid pico-hydropower network


Malgovert hydropower plant 74 concept 226
mathematical differential equations of opening feedback (OF) 30, 38
each hydraulic component 9 opening slew rate 115, 120
mathematical governing equations 8 open-loop þ closed-loop method 33
mechanical model of the generating operating point of hydraulic turbine
unit 59 10, 73
linearized model 60–1 orifice surge tank, model of 26–7
non-linear model in per unit 60 Osório Wind Park 209, 217
Microsoft Windows 7 overpressure 120
momentum equation 22, 26–7, 54–5,
82, 84 parallel proportional integral derivative
multi-machine system 62 (PID) 85
multi-penstock HPP 56 Particle Swarm optimization 163
multiplexing 6–7 PC Cards 4
PCI extensions for instrumentation
National Instruments 7 (PXI) 4
National Renewable Energy Pelton turbine 53–5, 57, 123, 126
Laboratory 193, 211 Pelton wheel 53
network codes (NCs) penstock 9–10, 82, 94
development of 109–10 discretisation 52
network flow algorithms 163 starting time in 10
Newton’s second law 54, 84 and tunnel models 51–2
non-linear simulation models 105, water start time criterion 75–6
119–22 peripheral component interconnect
non minimum phase system (PCI) personal computer
72, 75 bus 4
nozzle-Deflector Conjugation persistent oscillation 87–8, 92, 97
Function 11 Personal Computer Memory Card
numerical model of hydropower International Association
plants 22 (PCMCIA) 4
features of the model 31 personal computers (PCs) 4
hydropower unit with Francis phase locked loop (PLL) 237–9
turbine 27 photo-voltaic (PV) hydro hybrid
governor system 28–31 system 187
turbine and generator 27–8 components of 191–3
piping system 22 cost of 194
branch junction (pipeline floating structures 188, 191–2, 194,
junction) 25–6 196
downstream reservoir with parameters of 195
constant water level 24–5 results and discussion 197–202
series junction 25 schematic drawing of 193
surge tank 26–7 simulations with HOMER software
upstream reservoir with constant 193–6
water level 23–4 solar energy availability 196
260 Modeling and dynamic behaviour of hydropower plants

using residual flow of Guarita HPP, pipeline junction: see branch junction
in Brazil 187–202 piping system of HPPs 22
see also Guarita hydroelectric power branch junction (pipeline junction)
plant 25–6
photo-voltaic (PV) wind hydro hybrid downstream reservoir with constant
system with pumped storage water level 24–5
capacity 205 series junction 25
components of 208–10 surge tank 26–7
Linha Sete pumped storage power upstream reservoir with constant
plant 207–8 water level 23–4
results and discussion 212–19 poly-harmonic low-frequency
simulations with HOMER 211–12 vibration 132
pico-hydropower off-grid network 225 primary frequency control (PFC) 35–7
component models 227 performance limitation for 74
DC–DC converter 231–2 penstock water start time
generator 230 criterion 75–6
inverter modeling 232 surge tank cross-section criterion
rectifier 231 76–7
shaft assembly 228–30 power response in, for HPP 116,
transmission line and load 119, 121–2, 124
modeling 232–3 schematic block diagram 118
turbine 228 turbine governing system in 114,
control scheme design 233 117
inverter control design 233–9 Proportional Integral (PI)
turbine and DC–DC converter compensator 233
controller design 233 pseudoconcave functions 172
hybrid renewable off-grid network pumped storage hydroelectric power
245 plant, implementation of 206
hybrid grid simulation 248–9
solar PV interface modifications range, on DAQ board 7
246 reaction turbine 27, 53, 55
wind turbine interface hydraulic–electric analogy for 56
modifications 247–8 rectifier 227, 231
modeling of implementation in reduced order models for grid
Nepal 242–5 connected hydropower plants:
simulation results 239 see grid connected hydropower
change in input power 242 plants, reduced order models
performance with non-linear for
load 241–2 relaxed objective function value 179
power sharing performance 242 renewable energy sources (RES)
single generator unit with varying generation, intermittent 107–8
load 239–41 replacement reserves (RR) 111
system overview 226–7 requirements for grid connection of
Piggott Turbine 248 generators network code (RfG
pipeline, mathematical model of 82–4 NC) 110–12
Index 261

reservoir stair-like model 178 simulations with HOMER 193–6,


residual flow of Guarita HPP 187–202 211–12
resistance temperature detector Simulink library 3, 18
(RTDs) 5 slow dynamics, decomposition of
resolution, on DAQ board 7 72–4
rotary blade hydro turbine 136 Sobradinho hydropower plant,
Routh criterion 89–90 auxiliary functions for 171
run-off-the-river plant with variable spectral analysis, for vibration control
tailrace height 171–2 135, 137–8, 140
run-off-type hydropower plant 3 speed signal generator (SSG) 11
excitation system 11–13 speed track, simulation without 158
governor system 11 ST1A static excitation system model
measurements 4 15
data acquisition (DAQ) hardware stability 79, 115, 118
6–7 of turbine governing system 86–99
LabVIEW 7–8 basic concepts 86–9
signal conditioning 5–6 critical stable sectional area of
transducers 5 surge tank 98–9
modeling of the plant 8–11 dynamic system 86–91
model validation/simulations Routh and Hurwitz criterion
13–18 89–91
‘run-of-the-river’ power plants 211 with/without surge tank 91–8
stair-like reservoir with uniform
sampling rate 6–7 tailrace 173
second-order generalized integrator starting time in penstock 10
(SOGI) 236–9 start-up process of hydropower unit
sequential linear programming 163 33–5
series junction, model of 25 state-space model for hydro plant
servomotor 82 connected to a grid 67
shaft assembly model 228–30 concatenation 68
signal conditioning 5 electro-mechanical subsystem
amplification 5 67–8
excitation 6 hydro-mechanical subsystem 67
filtering 6 interconnected operation 68–9
isolation 5–6 isolated operation 69
linearization 6 static characteristic relationships of
multiplexing 6 hydraulic turbines 10
signal conditioning extension for static excitation systems (SES) 11
instrumentation (SCXI) 5–6 storage plant with constant tailrace
SimPowerSystems 232 height 172–3
simulation numerical studies storage plant with variable tailrace
for hydropower generation control height 174–5
general issues 119 surge tank model 26–7, 53, 79–80
principles 119–20 cross-section criterion 76–7
results for HPP case 120–3 impulse turbine 53–5
262 Modeling and dynamic behaviour of hydropower plants

mathematical model of 84 turbine


reaction turbine 55 and DC–DC converter controller
synchronous multi-machine system 61 design 233
synchronous power system models 61 and generator 27–8
general model 62–4 mathematical model of 84–5
interconnected grid, model for turbine control system 81–2
64–6 turbine governing system 79
isolated grid, model for 66–7 composite structure/control flow of
synchronous reference frame–based 81
PLL 238 electricity submodel 86
fluctuation stability of 98–9
telegraph equations 51 French EDF experience 112–27
Tennant method 208 hydraulic submodel
thermistors 5–6 mathematical model of pipeline
thermocouples 5–6 82–4
Third Energy Package, EU 106, 110 mathematical model of surge
Thoma assumption 99 tank 84
Thoma criterion 40 mechanic submodel
3-D hill chart 12 mathematical model of governor
time constants, calculating 3 86
time-domain analysis of regulation and mathematical model of turbine
operation of HPPs 84–5
case study of various dynamic modeling of 80–6
processes of HPP 32 stability analysis of 86–99
emergency stop and load critical stable sectional area of
rejection 43–5 surge tank 98–9
grid-connected operation 35–40 dynamic system 86–91
isolated operation 40–3 with surge tank 94–8
start-up and no-load operation without surge tank 91–4
33–5 Turgo turbine model 228
numerical model of hydropower two-to-five-needle operation 155, 157
plants 22 transition from 156
features of the model 31
hydropower unit with Francis universal serial bus (USB) 4
turbine 27–31 UNIX 7
piping system 22–7 upstream reservoir
practical engineering case 31–2 model of 24
TOPSYS 22, 31–2 with constant water level 23–4
transducers 5–6
transient process, of dynamic system variable voltage source 56
86–8 vertical hydro generator 136
transmission line and load modeling vibration control, at HPP
232–3 controlling object and measurement
TSOs (transmission system operators) equipment characteristics
106–7, 114 136–7
Index 263

data wavelet analysis 137–44 Virtual Instrument 8


implementation of wavelet virtual resistance 235
transform 133–5
signal analysis methods for water conservation and discharge
131–44 limits 175
vibration displacement, average value head sensitive discharge limits 176
of 141–3 water inertia 83–4, 97, 115, 120
vibration sensors wavelet coefficients 133, 140, 143–4
layout of absolute values of 142
on stator of hydraulic generator and vibration displacement 143
137 visualization of 135, 141
on supporting structures of wavelet transform analysis,
hydraulic unit 138 implementation of 133–5
measuring vibration of wind hydro hybrid system with water
hydro generator stator’s storage capacity 211
steel structures 133, 136–40 Work-Energy theorem 59

You might also like