Psoc Res IV Year
Psoc Res IV Year
Psoc Res IV Year
Hêmin Golpîra
University of Kurdistan
Kurdistan, Iran
Arturo Román-Messina
The Center for Research and Advanced Studies of IPN
Guadalajara, Mexico
Hassan Bevrani
University of Kurdistan
Kurdistan, Iran
This edition first published 2021
© 2021 John Wiley & Sons, Inc.
All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or
transmitted, in any form or by any means, electronic, mechanical, photocopying, recording or
otherwise, except as permitted by law. Advice on how to obtain permission to reuse material from
this title is available at http://www.wiley.com/go/permissions.
The right of Hêmin Golpîra, Arturo Román-Messina, and Hassan Bevrani to be identified as the
authors of this work has been asserted in accordance with law.
Registered Office
John Wiley & Sons, Inc., 111 River Street, Hoboken, NJ 07030, USA
Editorial Office
111 River Street, Hoboken, NJ 07030, USA
For details of our global editorial offices, customer services, and more information about Wiley
products visit us at www.wiley.com.
Wiley also publishes its books in a variety of electronic formats and by print-on-demand. Some
content that appears in standard print versions of this book may not be available in other formats.
Limit of Liability/Disclaimer of Warranty
While the publisher and authors have used their best efforts in preparing this work, they make
no representations or warranties with respect to the accuracy or completeness of the contents
of this work and specifically disclaim all warranties, including without limitation any implied
warranties of merchantability or fitness for a particular purpose. No warranty may be created or
extended by sales representatives, written sales materials or promotional statements for this work.
The fact that an organization, website, or product is referred to in this work as a citation and/or
potential source of further information does not mean that the publisher and authors endorse
the information or services the organization, website, or product may provide or recommendations
it may make. This work is sold with the understanding that the publisher is not engaged in
rendering professional services. The advice and strategies contained herein may not be suitable
for your situation. You should consult with a specialist where appropriate. Further, readers
should be aware that websites listed in this work may have changed or disappeared between when
this work was written and when it is read. Neither the publisher nor authors shall be liable for any
loss of profit or any other commercial damages, including but not limited to special,
incidental, consequential, or other damages.
10 9 8 7 6 5 4 3 2 1
To our mothers, Himêra, Diana and Ayesha
vii
Contents
Preface xiii
Acknowledgments xvii
Nomenclature xviii
List of Abbreviations and Acronyms xxi
1 Introduction 1
1.1 Power System Stability and Control 1
1.2 Current State of Power System Stability and Control 4
1.2.1 Frequency Control 5
1.2.2 Voltage Control 6
1.2.3 Oscillation Damping 7
1.3 Data-Driven Wide-Area Power System Monitoring and Control 8
1.4 Dynamics Modeling and Parameters Estimation 10
1.4.1 Modeling of Frequency, Voltage, and Angle Controls 11
1.4.2 Parameters Estimation 12
1.5 Summary 14
References 14
Index 293
xiii
Preface
universities, and power electric companies. The book also provides a thorough
understanding of the basic principles of the penetrated power system modeling,
model order reduction, and grid-connected MG equivalent model derivation.
One of the main parts of this book covers the modeling of power systems using
PMU data for the purpose of renewable integrated dynamics identification and
parameters estimation issues, as well as oscillation damping, voltage control,
and frequency control design problems. For this purpose, in addition to real net-
work data, several standard IEEE power system models are used as benchmarks
for generating data that are used in system identification. Furthermore, in addition
to dynamic stability analysis and controller synthesis, inertia challenge require-
ments and control levels are discussed, and recent advances in visualization of vir-
tual synchronous generators (VSGs) and the associated effects on system
performance are addressed.
This book could be useful for engineers and operators working on power systems
dynamic, control, and operation, as well as postgraduate students and academic
researchers. The book describes renewable integrated power system dynamics
modeling and control issues from introductory to the advanced steps. This book
is organized into eight chapters.
Chapter 1 discusses the term of power system stability and control with an
updated brief review on the areas of frequency, voltage, and angle controls, con-
cerning the penetration of RESs/DGs. In response to the existing challenges in
penetration of more RESs/DGs to the grid, the necessity of using data-driven mod-
eling, parameters estimation, and control synthesis in wide-area power systems is
emphasized; a general scheme for wide-area measurement system and wide-area
control is described.
Chapter 2 deals with dynamic equivalencing of penetrated power grid. Several
methods are introduced to model the host grid as well as the distribution network.
A center of gravity (COG)-based equivalent model is addressed to represent the
power system dynamic behavior in terms of slow power and frequency dynamics.
The relationship between the frequency of the COG and the motion of local cen-
ters of angle is analytically determined to compute local frequency deviations fol-
lowing major disturbances.
Chapter 3 addresses the power grid stability analysis from frequency, small sig-
nal and voltage points of view. Some analytical approaches have been discussed to
determine maximum penetration level of MGs concerning the upward system sta-
bility. The given methods explicitly rely on the basic power system equations
which, in turn, make the proposed indices completely independent of the test case;
and this helps to solve the associated difficulties with the system dimensions.
Chapter 4 explains the VSG concept and its applications in renewable integrated
power grids. The positive dynamic impacts of VSGs in a power system are dis-
cussed, and recent relevant achievements in the application of the advanced
Preface xv
Acknowledgments
Most of the contributions, outcomes, and insight presented in this book were
achieved through a long-term teaching and research cooperation on the renewable
integrated power systems over the last 15 years. The materials given in the present
book are mainly the research outcomes and original results of authors in Smart/
Micro Grids Research Center-SMGRC, University of Kurdistan (Sanandaj, Iran)
and in The Center for Research and Advanced Studies-CINVESTAV of the
National Polytechnic Institute of Mexico (Guadalajara, Mexico). It is a pleasure
to acknowledge the received supports from these sources, and the awards from
Iran Grid Management Company (IGMC), Iran National Science Foundation
(INSF), and Alexander von Humboldt (AvH) Foundation.
The authors would like to thank their colleagues Prof. J. Raisch and Prof. Bruno
Francois for their kind support. Finally, the authors offer their deepest personal
gratitude to their families for their patience during the preparation of this book.
xviii
Nomenclature
A amplitude
AGC automatic generation control
AQR automatic reactive power regulator
AVR automatic voltage regulator
B residue
BFV best fitness value
COG center of gravity
COI center of inertia
DER distributed energy resources
DFIG doubly-fed induction generator
DG distributed generation
DM diffusion map
DMD dynamic mode decomposition
E energy
EMT electromagnetic transient
ESS energy storage system
EV electrical vehicle
GA genetic algorithm
HVDC high-voltage direct current
KT Kumaresan–Tuft
LCC line commutated converter
MCL Markov clustering
MG microgrid
MMGs multi-MGs
NERC North American Electric Reliability Corporation
NYNE New York New England
PC principal component
PCTVAR percentage of variation
PF participation factor
xxii List of Abbreviations and Acronyms
PI proportional-integral
PLL phase-locked loop
PLS partial least squares regression
PLSC partial least squares correlation
PMSG permanent magnet synchronous generators
PMU phasor measurement unit
POIS point of interconnection with the system
PS pseudo spectrum
PSS power system stabilizers
PV photovoltaic
RES renewable energy source
RoCoF rate of change of frequency
SC synchronous condenser
SCADA supervisory control and data acquisition
SG synchronous generator
SLB static load bank
SOC state of charge
SS static switch
SVC static VAR compensator
SVD singular value decomposition
T-D time domain
TSO transmission system operator
UCTE Union for the Coordination of the Transmission of Electricity
UFLS underfrequency load shedding
ULTC under load tap changer
VAR volt–ampere reactive
VSC voltage source converter
VSG virtual synchronous generator
V2G vehicle-to-grid
Vvand Vandermonde matrix
WAMS wide-area measurement system
WF wind farm
WT wind turbine
z complex amplitude
1
Introduction
The term power system stability and control is used to define the application of con-
trol theorems and relevant technologies to analyze and enhance the power system
functions during normal and abnormal operations. Power system stability and
control refers to keep desired performance and stabilizing power system following
various disturbances, such as short circuits, loss of generation, and load.
The capacity of installed inverter-based distributed generators (DGs) and renewa-
ble energy sources (RESs) individually or through the microgrids (MGs) in power sys-
tems is rapidly growing, and a high penetration level is targeted for the next few
decades. In most countries including developing countries, significant targets are con-
sidered for using the distributed microsources and MGs in their power systems for
near future. The increase of DGs/RESs in power systems has a significant impact
on CO2 reduction; however, recent studies have shown that relatively high DGs/RESs
integration will have some negative impacts on power system dynamics, frequency
and voltage regulation, as well as other control and operational issues. Decreasing
system inertia and highly variable dynamic nature of DGs/RESs/MGs are known
as the main reasons. These impacts may increase for the dynamically weak power
systems at the penetration rates that are expected over the next several years.
In this chapter, a brief discussion on the power system stability and control in
modern renewable integrated power systems and the current state of this topic are
given. Data-driven wide-area power system monitoring and control is emphasized,
and the significance of measurement-based dynamic modeling and parameter esti-
mation is shown.
Power system stability and control was first recognized as an important problem in
1920s [1]. Over the years, numerous modeling/simulation programs, synthesis/
Renewable Integrated Power System Stability and Control, First Edition.
Hêmin Golpîra, Arturo Román-Messina, and Hassan Bevrani.
© 2021 John Wiley & Sons, Inc. Published 2021 by John Wiley & Sons, Inc.
2 1 Introduction
analysis methodologies, and protection schemes have been developed. Power grid
control must provide the ability of an electric power to regain a state of operating
equilibrium after being subjected to a physical disturbance, with most system vari-
ables, i.e., frequency, voltage, and angle, bounded so that practically the entire sys-
tem remains intact. Thus, the main control loops are known as frequency control,
voltage control, and rotor angle (power oscillation damping) control [2].
In many power systems, advanced measurement devices such as phasor meas-
urement units (PMUs) and modern communication devices are already being
installed. Using these facilities, the parameters of existing power system control-
lers can be adjusted by an online data-driven control mechanism [3]. The PMU
data after filtering are used to estimate some important parameters in the system
(scheduling parameters). These parameters are then used in the control tuning
algorithm that will adapt the controller parameters in frequency control, voltage
control, and power oscillation control. Therefore, the controller’s parameters are
adapted according to the current status of the system.
One of the important steps of reliable and performant control system design
is defining the performance specifications. It depends on the features of the con-
troller design method, the constraints on the controller structure, the achievable
performance that is limited by the physical constraints, the industrial standards on
the limit of the variables, the limits of the actuators, etc. Finding the control spe-
cifications and making them compatible with the controller design approach
require a deeper understanding of the physical system to be controlled.
The characteristics of three main control loops, i.e., frequency control, voltage
control, and angle control, should be studied to enable the definition of achievable
performance specifications and designing an effective control system.
tuning algorithm must adjust the LFC parameters to restore the frequency and
tie-line powers to the specified values.
• Angle control: Rotor angle stability is the ability of the power system to maintain
synchronization after being subjected to a disturbance. Angle stability refers to
damping of power oscillations inside subsystems and between subsystems on an
interconnected grid during variation beyond specified threshold levels. The risk
of losing angle stability can be significantly reduced by using proper control
devices inserted into the power grid to find a smooth shape for the system
dynamic response.
The power oscillation damping has been mainly guaranteed by power system
stabilizers (PSSs). A PSS is a controller, which, beside the turbine-governing sys-
tem, performs an additional supplementary control loop to the AVR system of a
generating unit. Depending on the type of PSS, the input signal could be the
rotor speed/frequency deviation, the generator active power deviation, or a com-
bination feedback of rotor speed/frequency and active power changes. This sig-
nal to be passed through a combination of a lead-lag compensators. The PSS
output signal is amplified to provide an effective output signal.
In order to damp the inter-area oscillations, which have smaller oscillation
frequency than the local oscillatory modes, a wide-area control (WAC) system
is required. The WAC system is a centralized controller that uses the PMU sig-
nals and produces auxiliary control signals for the PSSs.
This VSG provides a promising solution to improve power grid stability and per-
formance in the presence of a high penetration of DGs/RESs/MGs. The VSG is not
only applicable for improving of frequency regulation and oscillations damping,
particularly during the transient state following a disturbance, but also it is useful
to support the voltage stability. The VSG system can use the available DGs/RESs,
as primary sources to participate in power oscillation damping by adjusting their
active and reactive power generations. The VSG is more discussed in Chapter 4.
Power system stability and control can take different forms, which are influenced
by the type of instability phenomena. A survey on the basics of power system con-
trols, literature, and achievements is given in [6, 7].
PMUs are sophisticated digital recording devices that communicate global
positioning system (GPS) synchronized high sampling rate dynamic power sys-
tem’s data to the central control and monitoring stations. The recorded data by
PMUs provide valuable information about the dynamic of the power system that
can be used for data-driven modeling. An overview of system identification tech-
niques for modeling of power systems using PMU data is given in [8]. In [9], a sub-
space identification method is used to identify a reduced order model for power
oscillation control. The PMU data are used for the calibration of the parameters
of the reduced-order model of a power generator in [10]. The feasibility of
multi-input multi-output (MIMO) identification of power systems using low-level
Power Grid
Primary DC/DC
Source Pin DC/AC Pout
Inverter
AC/DC
VSG Control
probing signal is shown in [11]. An online algorithm is used in [12] to identify the
frequency response of power system dynamics, while it is combined with a
selective modal analysis. The transfer function and state-space model identifica-
tions using PMU data are compared in [13] for electromechanical oscillation
damping estimation. Several identification methods are compared for analysis
of inter-area oscillatory modes of power systems [14].
The data from PMUs have already been used for estimation of some important
power system parameters. The electromechanical modes of a power system and
their confidence intervals are estimated using PMUs operational data in [15,
16]. Amplitude, frequency, and damping of power system oscillations are esti-
mated using PMU measurements in [17, 18]. The PMU data are used in [19–21]
to identify the topology (or change in topology) of a power system. Recently, some
system identification methods have been employed to estimate the power system
inertia using the operational PMU data (with no external excitation signal)
[22, 23].
control supports in the power grids [62–69]. Providing frequency control support
via controllable loads and smart load technologies using the concepts of demand
response (DR) is discussed in [41, 42, 70–76]. Two recent works in this area are [77,
78], that discuss the impact of a high integration of MGs on the frequency control
of power systems, and propose a decentralized stochastic frequency control
of MGs.
PMU-based/data-driven online tuning frequency control approach is not
addressed in the abovementioned worldwide published works. In most cases,
the secondary frequency control is designed using conventional frequency
response model, which is very difficult to realize in a modern power grid with a
highly variable structure and penetration of DGs/RESs.
using PMU remote feedback signals was discussed. References [100, 111] proposed
the design of wide-area damping controllers that provide supplementary damping
control to synchronous generators (SGs). A networked control system model for
wide-area closed-loop power systems is applied in [109]. A power oscillation
damping controller is introduced in [112] based on a modal linear quadratic Gaus-
sian methodology. A combination of controlling SGs and renewable sources in
order to increase the overall damping capability of the system is shown in [101,
107, 108, 113]. Few LPV control solutions to power oscillation damping are pro-
posed that use either a low-order first principle model of the system [114] or a
reduced-order parametric LPV identified model [115].
In comparison of frequency and voltage control, a higher number of reports have
been published in PMU-based oscillation damping (rotor angle control) field.
However, most of the reported approaches require the detailed and accurate
knowledge of the complete network model (both topology and parameter values),
that is unavailable or corrupted in practice as a result of communication failures,
bad data in state estimation etc. In addition, the impact of disturbances on the
inter-area oscillations cannot be well captured by these methods.
Power grids modeling and control has become a more challenging issue due to
the increasing penetration of RESs, changing system structure and the integration
of new storage systems, controllable loads and power electronics technologies,
and reduction of system inertia. Conventional modeling and control designs
may not be any more effective to satisfy all specified objectives in various operation
modes of modern power grids. These challenging issues set new demand for
the development of more flexible, rapid, effective, precise, and adaptive
approaches for power system dynamic monitoring, stability/security analysis,
and control problems. Thanks to recent advances in control, communication,
and computing technologies, it is possible to tackle mentioned challenges by
implementing a data-driven-based modeling and control framework as shown
in Figure 1.2.
The system data are collected from the distributed PMUs in the grid through
a secure communication network. The development of information and commu-
nication technology (ICT) enables more flexibility in wide-area monitoring of
power system with fast and large data transmission. Especially, the wide-area
measurement system (WAMS) with PMUs is a promising technique as one of
the smart grid technologies in the bulk power grid.
1.3 Data-Driven Wide-Area Power System Monitoring and Control 9
Data
Storage
Parameters Estimation
and Dynamic Modeling
Figure 1.2 An overall data-driven control framework for renewable integrated power
systems.
The measured data are locally saved and then collected by phasor data concen-
trators (PDCs) for the post analysis or sent to a remote location via a standard
data format. These data with the time stamp of the synchronized GPS in real time
may applied for parameter and state estimations and finally used for the system
protection and/or real-time control. Figure 1.3 shows how a PMU-based WAMS
can provide data for the power system control center to generate continuous (in
normal states) and discontinuous control (in off-normal states) commands.
Before any application, the collected PMU data need to be cleaned and de-noised
and employed by the data processors for estimation, modeling, and control pur-
poses. The proposed de-noising method may use a rolling-averaging window with
pre-specified length to remove noise from the recorded data. The block of para-
meters estimation algorithms contains high fast and precise algorithms for estima-
tion of some important parameters and transient characteristics that are required
to use in control tuning algorithm or to detect a contingency and triggering the
emergency control and protection schemes. In case of crossing the assigned thresh-
olds showing an off-normal and emergency condition, the recorded data and some
estimated parameters are used to detect the amount of mismatch (size of distur-
bance) for the emergency control and protection schemes such as load shedding
10 1 Introduction
GPS
Continuous control
Data Storage
Data Analysis,
Parameter Estimation
and Control Center
Synchrophasor
Communication PDC
System (SPCS)
Emergency control
Power Grid
Discontinuous control and protection
From a system dynamic point of view, the bulk generating units, due to their high
inertia, provide a long time constant; such that the rotor speed and thus the grid
frequency cannot alter suddenly, while the load changes. Hence, the total rotating
mass enhances the dynamic stability. In future, a significant share of DGs/RESs/
MGs in the electric power grids is expected. This increases the total system gener-
ation power, while does not contribute to the system rotational inertia. System
dynamics are faster in power systems with low rotational inertia, making control
and power system operation more challenging [32].
1.4 Dynamics Modeling and Parameters Estimation 11
actual time. The proposed scheme relies on fast, yet iterative, estimation of fre-
quency nadir, and time of minimum frequency occurrence. Accordingly, the iner-
tia constant as well as the size of power mismatch are estimated which, in turn,
compares with the maximum size of imbalance, satisfying the pre-specified thresh-
olds, to determine the amount of shed load.
1.5 Summary
Modern power grids face new technical challenges arising from the increasing
penetration of power-electronic-connected RESs/DGs. Increasing MGs/DGs pen-
etration level may adversely affect frequency response and voltage and system con-
trol and lead to degraded performance of traditional control schemes. This, in turn,
may result in large deviations and, potentially, system instability.
This chapter provides the pre-requirement terminology and general background
for the next chapters of this book. The term power system stability and control with
an updated brief review on the areas of frequency, voltage, and angle controls, con-
cerning the penetration of RESs/DGs, is discussed. In response to the existing chal-
lenges in penetration of more RESs/DGs to the grid, the necessity of using data-
driven modeling, parameters estimation, and control synthesis in wide-area power
systems is emphasized.
References
1. Steinmetz, C.P. (1920). Power control and stability of electric generating stations.
Transactions of the American Institute of Electrical Engineers XXXIX (2):
1215–1287.
2. Smith, J.R., Andersson, G., and Taylor, C.W. (1996). Annotated bibliography on
power system stability controls: 1986–1994. IEEE Transactions on Power Systems
11 (2): 794–800.
3. Bevrani, H., Watanabe, M., and Mitani, Y. (2014). Power System Monitoring and
Control. Wiley.
4. Bevrani, H. and Hiyama, T. (2009). On load-frequency regulation with time
delays: design and real time implementation, IEEE Transactions on Energy
Conversion 24 (1): 292–300.
5. Bevrani, H. and Hiyama, T. (2011). Intelligent Automatic Generation Control.
CRC Press.
6. Kundur, P. (1994). Power System Stability and Control, vol. 7. New York:
McGraw-Hill.
References 15
7. Kundur, P., Paserba, J., Ajjarapu, V. et al. (2004). Definition and classification of
power system stability ieee/cigre joint task force on stability terms and definitions.
IEEE Transactions on Power Systems 19 (3): 1387–1401.
8. Pierre, J.W., Trudnowski, D., Donnelly, M. et al. (2012). Overview of system
identification for power systems from measured responses. IFAC Proceedings
Volumes 45 (16): 989–1000.
9. Eriksson, R. and Soder, L. (2011). Wide-area measurement system-based subspace
identification for obtaining linear models to centrally coordinate controllable
devices. IEEE Transactions on Power Delivery 26 (2): 988–997.
10. Zhou, N., Lu, S., Singh, R., and Elizondo, M.A. (2011). Calibration of reduced
dynamic models of power systems using phasor measurement unit (PMU) data.
2011 North American Power Symposium, Boston, MA (2011), pp. 1–7. doi: https://
doi.org/10.1109/NAPS.2011.6024873.
11. Zhang, J., Lu, C., and Han, Y. (2013). MIMO identification of power system with
low level probing tests: applicability comparison of subspace methods. IEEE
Transactions on Power Systems 28 (3): 2907–2917.
12. Wiseman, B.P., Chen, Y., Xie, L., and Kumar, P. (2016). PMU-based reduced-order
modeling of power system dynamics via selective modal analysis. In: 2016
IEEE/PES Transmission and Distribution Conference and Exposition (T&D),
1–5. IEEE.
13. Liu, H., Zhu, L., Pan, Z. et al. (2016). Comparison of MIMO system identification
methods for electromechanical oscillation damping estimation. 2016 IEEE Power
and Energy Society General Meeting (PESGM), Boston, MA (2016), pp. 1–5.
doi: https://doi.org/10.1109/PESGM.2016.7741834.
14. Tuttelberg, K., Kilter, J., and Uhlen, K. (2017). Comparison of system
identification methods applied to analysis of inter-area modes. Proceedings of
International Power Systems Transients Conference 2017, Seoul, South Korea
(26–29 June 2017).
15. Ghasemi, H. and Canizares, C.A. (2008). Confidence intervals estimation in the
identification of electromechanical modes from ambient noise. IEEE Transactions
on Power Systems 23 (2): 641–648.
16. Dosiek, L., Pierre, J.W., and Follum, J. (2013). A recursive maximum likelihood
estimator for the online estimation of electromechanical modes with error
bounds. IEEE Transactions on Power Systems 28 (1): 441–451.
17. Uhlen, K., Warland, L., Gjerde, J.O. et al. (2008). Monitoring amplitude,
frequency and damping of power system oscillations with PMU measurements.
2008 IEEE Power and Energy Society General Meeting – Conversion and Delivery of
Electrical Energy in the 21st Century, Pittsburgh, PA (2008), pp. 1–7, doi: https://
doi.org/10.1109/PES.2008.4596661.
18. Tripathy, P., Srivastava, S.C., and Singh, S.N. (2011). A modified TLS-ESPRIT-
based method for low frequency mode identification in power systems utilizing
16 1 Introduction
31. Ibraheem, P., Kumar, and Kothari, D.P. (2005). Recent philosophies of automatic
generation control strategies in power systems. IEEE Transactions on Power
Systems 20 (1): 346–357.
32. Bevrani, H. (2014). Robust Power System Frequency Control, 2e. Gewerbestrasse,
Switzerland: Springer.
33. Ulbig, A., Borsche, T.S., and Andersson, G. (2014). Impact of low rotational inertia
on power system stability and operation. IFAC Proceedings Volumes 47 (3):
7290–7297.
34. Jaleeli, N. and VanSlyck, L.S. (1999). NERC’s new control performance standards.
IEEE Transactions on Power Systems 14 (3): 1092–1099.
35. Hain, Y., Kulessky, R., and Nudelman, G. (2000). Identification-based power unit
model for load-frequency control purposes. IEEE Transactions on Power Systems
15 (4): 1313–1321.
36. Chang-Chien, L.R., Hoonchareon, N.-B., Ong, C.-M., and Kramer, R.A. (2003).
Estimation of /spl beta/ for adaptive frequency bias setting in load frequency
control. IEEE Transactions on Power Systems 18 (2): 904–911.
37. Wilches-Bernal, F., Concepcion, R., Neely, J.C. et al. (2018). Communication
enabled fast acting imbalance reserve (CE-FAIR). IEEE Transactions on Power
Systems 33 (1): 1101–1103.
38. Zhang, G. and McCalley, J.D. (2018). Estimation of regulation reserve
requirement based on control performance standard. IEEE Transactions on Power
Systems 33 (2): 1173–1183.
39. Polajzer, B., Brezovnik, R., and Ritonja, J. (2017). Evaluation of load frequency
control performance based on standard deviational ellipses. IEEE Transactions on
Power Systems 32 (3): 2296–2304.
40. Avila, T., Gutierrez, E., and Chavez, H. (2017). Performance standard-compliant
secondary control: the case of Chile. IEEE Latin America Transactions 15 (7):
1257–1262.
41. Douglas, L.D., Green, T.A., and Kramer, R.A. (1994). New approaches to the AGC
nonconforming load problem. IEEE Transactions on Power Systems 9 (2): 619–628.
https://doi.org/10.1109/59.317682.
42. Trovato, V., Sanz, I.M., Chaudhuri, B., and Strbac, G. (2017). Advanced control of
thermostatic loads for rapid frequency response in Great Britain. IEEE
Transactions on Power Systems 32 (3): 2106–2117.
43. Delavari, A. and Kamwa, I. (2018). Improved optimal decentralized load
modulation for power system primary frequency regulation. IEEE Transactions
on Power Systems 33 (1): 1013–1025.
44. Pan, C. and Liaw, C. (1989). An adaptive controller for power system load-
frequency control. IEEE Transactions on Power Systems 4 (1): 122–128.
45. Vajk, I., Vajta, M., Keviczky, L. et al. (1985). Adaptive load-frequency control of
the Hungarian power system. Automatica 21 (2): 129–137.
18 1 Introduction
46. Wang, W., Li, Y., Cao, Y. et al. (2018). Adaptive droop control of VSC-MTDC
system for frequency support and power sharing. IEEE Transactions on Power
Systems 33 (2): 1264–1274.
47. Prostejovsky, A.M., Marinelli, M., Rezkalla, M. et al. (2018). Tuningless load
frequency control through active engagement of distributed resources. IEEE
Transactions on Power Systems 33 (3): 2929–2939.
48. Stankovic, A.M., Tadmor, G., and Sakharuk, T.A. (1998). On robust control
analysis and design for load frequency regulation. IEEE Transactions on Power
Systems 13 (2): 449–455.
49. Rerkpreedapong, D., Hasanovic, A., and Feliachi, A. (2003). Robust load
frequency control using genetic algorithms and linear matrix inequalities. IEEE
Transactions on Power Systems 18 (2): 855–861.
50. Ojaghi, P. and Rahmani, M. (2017). LMI-based robust predictive load frequency
control for power systems with communication delays. IEEE Transactions on
Power Systems 32 (5): 4091–4100.
51. Zhang, C., Jiang, L., Wu, Q.H. et al. (2013). Delay-dependent robust load
frequency control for time delay power systems. IEEE Transactions on Power
Systems 28 (3): 2192–2201.
52. Zhao, J., Mili, L., and Milano, F. (2018). Robust frequency divider for power
system online monitoring and control. IEEE Transactions on Power Systems 33 (4):
4414–4423.
53. Aliabadi, S.F., Taher, S.A., and Shahidehpour, M. (2018). Smart deregulated grid
frequency control in presence of renewable energy resources by EVs charging
control. IEEE Transactions on Smart Grid 9 (2): 1073–1085.
54. Wang, D., Liang, L., Hu, J. et al. (2018). Analysis of low-frequency stability in grid
tied DFIGs by non-minimum phase zero identification. IEEE Transactions on
Energy Conversion 33 (2): 716–729.
55. Liu, Y., Jiang, L., Wu, Q.H., and Zhou, X. (2017). Frequency control of DFIG-
based wind power penetrated power systems using switching angle controller and
AGC. IEEE Transactions on Power Systems 32 (2): 1553–1567.
56. Pradhan, C. and Bhende, C.N. (2017). Frequency sensitivity analysis of load
damping coefficient in wind farm-integrated power system. IEEE Transactions on
Power Systems 32 (2): 1016–1029.
57. Golpira, H., Seifi, H., Messina, A.R., and Haghifam, M. (2016). Maximum
penetration level of microgrids in large-scale power systems: frequency stability
viewpoint. IEEE Transactions on Power Systems 31 (6): 5163–5171.
58. Leon, A.E. (2018). Short-term frequency regulation and inertia emulation using
an MMC-based MTDC system. IEEE Transactions on Power Systems 33 (3):
2854–2863.
References 19
59. Rakhshani, E., Remon, D., Cantarellas, A.M. et al. (2017). Virtual synchronous
power strategy for multiple HVDC interconnections of multi-area AGC power
systems. IEEE Transactions on Power Systems 32 (3): 1665–1677.
60. Li, D., Zhu, Q., Lin, S., and Bian, X.Y. (2017). A self-adaptive inertia and damping
combination control of VSG to support frequency stability. IEEE Transactions on
Energy Conversion 32 (1): 397–398.
61. Wu, Y., Yang, W., Hu, Y., and Dzung, P.Q. (2019). Frequency regulation at a wind
farm using time varying inertia and droop controls. IEEE Transactions on Industry
Applications 55 (1): 213–224.
62. Fang, J., Li, H., Tang, Y., and Blaabjerg, F. (2018). Distributed power system
virtual inertia implemented by grid-connected power converters. IEEE
Transactions on Power Electronics 33 (10): 8488–8499.
63. Li, Y., Xu, Z., Ostergaard, J., and Hill, D.J. (2017). Coordinated control strategies
for offshore wind farm integration via VSC-HVDC for system frequency support.
IEEE Transactions on Energy Conversion 32 (3): 843–856.
64. Ahmadyar, A.S. and Verbic, G. (2017). Coordinated operation strategy of wind
farms for frequency control by exploring wake interaction. IEEE Transactions on
Sustainable Energy 8 (1): 230–238.
65. Izadkhast, S., Garcia-Gonzalez, P., Frias, P., and Bauer, P. (2017). Design of plug-
in electric vehicle’s frequency-droop controller for primary frequency control and
performance assessment. IEEE Transactions on Power Systems 32 (6): 4241–4254.
66. Hwang, M., Muljadi, E., Jang, G., and Kang, Y.C. (2017). Disturbance-adaptive
short-term frequency support of a DFIG associated with the variable gain based
on the ROCOF and rotor speed. IEEE Transactions on Power Systems 32 (3):
1873–1881.
67. Attya, A.B.T. and Dominguez-Garcia, J.L. (2018). Insights on the provision of
frequency support by wind power and the impact on energy systems. IEEE
Transactions on Sustainable Energy 9 (2): 719–728.
68. Tielens, P. and Van Hertem, D. (2017). Receding horizon control of wind power to
provide frequency regulation. IEEE Transactions on Power Systems 32 (4): 2663–2672.
69. Garmroodi, M., Verbic, G., and Hill, D.J. (2018). Frequency support from wind
turbine generators with a time-variable droop characteristic. IEEE Transactions
on Sustainable Energy 9 (2): 676–684.
70. Khooban, M., Dragicevic, T., Blaabjerg, F., and Delimar, M. (2018). Shipboard
microgrids: a novel approach to load frequency control. IEEE Transactions on
Sustainable Energy 9 (2): 843–852.
71. Benysek, G., Bojarski, J., Smolenski, R. et al. (2018). Application of stochastic
decentralized active demand response (DADR) system for load frequency control.
IEEE Transactions on Smart Grid 9 (2): 1055–1062.
20 1 Introduction
72. Vrettos, E., Ziras, C., and Andersson, G. (2017). Fast and reliable primary
frequency reserves from refrigerators with decentralized stochastic control. IEEE
Transactions on Power Systems 32 (4): 2924–2941.
73. Short, J.A., Infield, D.G., and Freris, L.L. (2007). Stabilization of grid frequency
through dynamic demand control. IEEE Transactions on Power Systems 22 (3):
1284–1293.
74. Molina-Garcia, A., Bouffard, F., and Kirschen, D.S. (2011). Decentralized
demand-side contribution to primary frequency control. IEEE Transactions on
Power Systems 26 (1): 411–419.
75. Zhao, H., Wu, Q., Huang, S. et al. (2018). Hierarchical control of thermostatically
controlled loads for primary frequency support. IEEE Transactions on Smart Grid
9 (4): 2986–2998.
76. Yao, E., Wong, V.W.S., and Schober, R. (2017). Robust frequency regulation
capacity scheduling algorithm for electric vehicles. IEEE Transactions on Smart
Grid 8 (2): 984–997.
77. Ferraro, P., Crisostomi, E., Raugi, M., and Milano, F. (2017). Analysis of the
impact of microgrid penetration on power system dynamics. IEEE Transactions
on Power Systems 32 (5): 4101–4109.
78. Ferraro, P., Crisostomi, E., Shorten, R., and Milano, F. (2018). Stochastic
frequency control of grid connected microgrids. IEEE Transactions on Power
Systems 33 (5): 5704–5713.
79. Larsen, E. and Sener, F. (1996). Facts Applications. Catalogue No. 96TP116-0.
80. IEEE (1990). Voltage Stability of Power Systems: Concepts, Analytical Tools and
Industry Experience. IEEE Technical Report 90YH0358-2-PWR. IEEE/PES.
81. Balu, C. and Maratukulam, D. (1994). Power System Voltage Stability.
McGraw-Hill.
82. Van Cutsem, T. and Vournas, C. (2007). Voltage Stability of Electric Power Systems.
Springer Science & Business Media.
83. Kamwa, I., Grondin, R., and Hebert, Y. (2001). Wide-area measurement based
stabilizing control of large power systems – a decentralized/hierarchical
approach. IEEE Transactions on Power Systems 16 (1): 136–153.
84. Taylor, C.W., Erickson, D.C., Martin, K.E. et al. (2005). WACS wide-area stability
and voltage control system: R & D and online demonstration. Proceedings of the
IEEE 93 (5): 892–906.
85. Andersson, G., Bel, C.A., and Canizares, C. (2009). Frequency and voltage control.
In: Electric Energy Systems: Analysis and Operation. CRC Press.
86. Ilic, M.D., Liu, X., Leung, G. et al. (1995). Improved secondary and new tertiary
voltage control. IEEE Transactions on Power Systems 10 (4): 1851–1862.
87. Corsi, S., Pozzi, M., Sabelli, C., and Serrani, A. (2004). The coordinated automatic
voltage control of the Italian transmission grid-Part I: reasons of the choice and
References 21
100. Raoufat, M.E., Tomsovic, K., and Djouadi, S.M. (2016). Virtual actuators for wide-
area damping control of power systems. IEEE Transactions on Power Systems
31 (6): 4703–4711.
101. Mohagheghi, S., Venayagamoorthy, G.K., and Harley, R.G. (2007). Optimal wide
area controller and state predictor for a power system. IEEE Transactions on
Power Systems 22 (2): 693–705.
102. Mithulananthan, N., Canizares, C.A., Reeve, J., and Rogers, G.J. (2003).
Comparison of PSS, SVC, and STATCOM controllers for damping power system
oscillations. IEEE Transactions on Power Systems 18 (2): 786–792.
103. Bian, X.Y., Geng, Y., Lo, K.L. et al. (2016). Coordination of PSSs and SVC damping
controller to improve probabilistic small-signal stability \\of power system with
wind farm integration. IEEE Transactions on Power Systems 31 (3): 2371–2382.
104. Padhy, B.P., Srivastava, S.C., and Verma, N.K. (2017). A wide-area damping
controller considering network input and output delays and packet drop. IEEE
Transactions on Power Systems 32 (1): 166–176.
105. Giri, J. (2015). Proactive management of the future grid. IEEE Power and Energy
Technology Systems Journal 2 (2): 43–52.
106. Wu, X., Dorer, F., and Jovanovic, M.R. (2016). Input-output analysis and
decentralized optimal control of inter-area oscillations in power systems. IEEE
Transactions on Power Systems 31 (3): 2434–2444.
107. Zacharia, L., Hadjidemetriou, L., and Kyriakides, E. (2018). Integration of
renewables into the wide area control scheme for damping power oscillations.
IEEE Transactions on Power Systems 33 (5): 5778–5786.
108. Surinkaew, T. and Ngamroo, I. (2016). Hierarchical coordinated wide area and
local controls of DFIG wind turbine and PSS for robust power oscillation
damping. IEEE Transactions on Sustainable Energy 7 (3): 943–955.
109. Wang, S., Meng, X., and Chen, T. (2012). Wide-area control of power systems
through delayed network communication. IEEE Transactions on Control Systems
Technology 20 (2): 495–503.
110. Mokhtari, M. and Aminifar, F. (2014). Toward wide-area oscillation control
through doubly-fed induction generator wind farms. IEEE Transactions on Power
Systems 29 (6): 2985–2992.
111. Zhang, Y. and Bose, A. (2008). Design of wide-area damping controllers for inter-
area oscillations. IEEE Transactions on Power Systems 23 (3): 1136–1143.
112. Zenelis, I. and Wang, X. (2018). Wide-area damping control for interarea
oscillations in power grids based on PMU measurements. IEEE Control Systems
Letters 2 (4): 719–724.
113. Youseian, R., Bhattarai, R., and Kamalasadan, S. (2017). Transient stability
enhancement of power grid with integrated wide area control of wind farms and
synchronous generators. IEEE Transactions on Power Systems 32 (6): 4818–4831.
References 23
114. El-Guindy, A., Schaab, K., Schurmann, B. et al. (2017). Formal lpv control for
transient stability of power systems. In: 2017 IEEE Power & Energy Society General
Meeting, 1–5. IEEE.
115. Nogueira, F.G., Junior, W.B., da Costa Junior, C.T., and Lana, J.J. (2018).
LPV-based power system stabilizer: identification, control and field tests. Control
Engineering Practice 72: 53–67.
116. Toth, R. (2010). Modeling and Identification of Linear Parameter-Varying Systems,
vol. 403. Springer.
117. Golpira, H. and Bevrani, H. (2020). Frequency Analysis Based Centralized Load
Shedding and Island Detection Using PMU Data. Technical Report, Iran Grid
Management Company, IGMC1927, (In Persian).
25
2.1 Introduction
difficulties associated with the control and coordination of DERs [3, 4]. MGs, act-
ing as single controlled entities on the grid, can overcome some of the aforemen-
tioned limitations through local control of DERs/DGs, separation of generation
and corresponding loads from the distribution system in the presence of distur-
bances, and providing high local reliability for loads. In addition to the DERs/
DGs’ potential benefits, the conversion of only one-third of fuel energy into elec-
trical power and the poor efficiency of bulk power generation and high voltage
transmission systems make DERs/DGs in distribution network more pervasive
[3]. Moreover, the falling investment cost of small-scale power plants, develop-
ment of data communications and control technologies, the emerging potential
of DERs/DGs, and short installation time are the main incentives toward integrat-
ing significant amounts of DERs’ generation [5, 6]. It is envisaged that a large num-
ber of DERs/DGs will be connected to the host grid shortly, which will increase the
power system dimension and complexity.
In this chapter, a systematic methodology for modeling distributed MGs for inte-
gration studies is presented. This framework can be used to investigate the inte-
gration of high penetrations of DERs/MGs energy into the system.
d2 δCOI i t
M COI i = T mCOI i t − T eCOI i t , i = 1, …, n 21
dt 2
where δCOI i = j Ai M j δ j j Ai M j is the position of the COI, expressed in terms
of the individual rotor angles δj, M COI i = j Ai M j ωo is the equivalent inertia,
and T mCOI i t , T eCOI i t are the time-varying mechanical input power and electrical
output power, respectively, which include the effects of turbine dynamics and
other controllers [12].
With the COI description (2.1), used to represent the area response to external
forces, valuable information about local (bus) frequency behavior and inter-unit
synchronizing oscillations between generators are eliminated. Ideally, fast syn-
chronizing oscillations are suppressed when areas are chosen based on the
coherency identification techniques (uniform frequency for an area), such that
f j f Ai = f COI i , where f COI i = j Ai M j f j j Ai M j .
It is convenient to define fictitious transmission lines between key generators
(buses) and the COI to study the key generators’ motion, as illustrated in
2.3 Power Grid Modeling 29
Area n
Area 1
G1 Gj
δ1, M1 δj, Mj
δCOI , MCOI
1 1
XCOI1,I
COI XCOI1,j
XCOI1,m
Gm
δm, Mm
Figure 2.1 Area i illustrating the notion of the local COI. Dashed lines indicate virtual
connections between the motion of the area i COI and key system buses.
Figure 2.1. Conceptually, this is equivalent to use weighted strings or plumb line
techniques to find the COG in physical objects or systems.
Once the location of the COI is determined in Figure 2.1, expressions relating the
local frequencies fj, to the COI behavior can be obtained as
∂f j
fj= Δ f COI i , j Ai 22
∂ f COI i
where the sensitivities are determined from the parameters of the fictitious tie-
lines of Figure 2.1. By knowing these sensitivities, the propagation of slow fre-
quency oscillations and tie-line power exchanges may be evaluated.
Drawing on the above framework, Figure 2.2 schematically represents the posi-
tion of the COG relative to the local centers of inertia. Here, in analogy with the
single-area representation, Ptie tie
COI i ,COG and P COI i ,j represent virtual power flows across
the fictitious tie-lines interconnecting the ith local COI to the COG and the jth bus
30 2 MG Penetrated Power Grid Modeling
Area 2
Local COI
dynamics
δCOI1, MCOI1
COG
tie
Area k PCOI j,COG δCOG , MCOG
tie
PCOI n,COG
Figure 2.2 Interconnected power system divided into areas illustrating the notion of the
COG. Dotted lines indicate virtual connections between the motions of centers of inertia and
the system’s COG.
to the associated local COI that are of interest for calculation. Two issues are of key
interest here:
n
Ptie
n
COI i ,COG
T tie
COI i ,COG = =0 2 3b
i=1 i=1
2π f COG
in which, power transfers from the ith local COI to the COG are assumed to be of
the general form [14]:
V COI i V COG
Ptie
COI i ,COG = sin δCOI i − δCOG 24
X tie
COI i ,COG
where X tie
COI i ,COG is the virtual reactance between area i and the COG. On the other
hand, the condition for the equilibrium of the COI relative to the COG and local
buses requires that
n n Ptie
COI i ,j Ptie
COI i ,COG
T tie tie
COI i ,j + T COI i ,COG = + =0 25
j=1
2π f COI i 2π f COI i
j=1
i n
26
and similarly, for area Ai,
n
1 V j V COI i Ptie
COI i ,COG
T tie tie
COI i ,j + T COI i ,COG = sin δ j −δ COI i + =0
j
2π f COI i X tie
COI i ,j 2π f COI i
j=1
i n
27
In interpreting this model, note that the first term of (2.7) on the right-hand side
explains the virtual transferred power between bus j and the local COI, while the
second one describes the interactions between the local COI and the COG. The set
of equations (2.1)–(2.7) describes the multi-area dynamic energy balance and is
well suited for the efficient analysis of power and frequency transients of large
interconnected power systems. In the new equivalent system of Figure 2.2, the
equations are expressed in terms of variables that have the potential to approxi-
mate frequency behavior using a simplified model from which both local and
global properties can be analyzed. In what follows, the relationship between the
global frequency and the local frequencies is investigated. First, some basic
assumptions are discussed to derive the equivalent model of Figure 2.2.
32 2 MG Penetrated Power Grid Modeling
• Two different time horizons are assumed for voltage consideration in the
procedure of calculation of virtual impedances: (i) the period from when the
disturbance occurs to the time when the first under load tap changer (ULTC)
tap movement takes place and (ii) the time in which the ULTCs operate and bus
voltages increase to the reference values [15]. During the early period, bus voltage
magnitudes gradually come down to values lower than the pre-disturbance values.
The sensitivity of changes in the virtual impedances to the changes in bus voltage
magnitude, in these two periods, is assumed to be negligible. This means that the
ratio of impedances remains approximately constant over the time. Therefore, bus
voltage magnitudes are set to 1 pu in the process of calculation of impedances.
and
1 V j V COI i
min sin δ j − δCOI i + T tie
COI i ,COG 29
tie
X COI
i ,j j
2π f COI i X tie
COI i ,j
Since the minimum value of an absolute function is zero, the solutions of (2.8)
and (2.9) lead to the same results as solving (2.6) and (2.7). As the values of the
virtual impedances in (2.8) and (2.9) are used to weight the participation of each
area or bus in the multi-area dynamic energy balance, the optimization problem is
2.3 Power Grid Modeling 33
• Step 1. Given rotor angle positions δj, j = 1, …, ng and bus voltage magnitudes
Vj, minimize the absolute resultant torque in (2.8) and (2.9) using a GA. An
initial population of 200 chromosomes characterizes the GA. For a system
including n areas, each chromosome consists of 10 × n genes, representing
the X tie
COI i ,COG; each unknown variable, i.e. the fictitious reactance may be repre-
sented by 10 binary bits. Minimization of (2.9) requires defining 10 × (m + 1)
genes, where m is the number of generator buses in the associated area. This
means that m + 1 fictitious reactances, including m X tie tie
COI i ,j and one X COI i ,COG ,
interact with each other to make that the resultant torque vanishes for the local
COIs. As a result, only m X tie COI i ,j are considered as unknown variables.
•• tie
Step 2. Set X COI i ,COG to the value obtained from (2.8).
Step 3. Compute frequency (and power) sensitivities in (2.2).
• Step 4. Calculate local and global frequency responses and tie-line net flows.
where f COI i and Di are the COI frequency deviation for the area i following distur-
bance ΔPi and load damping parameter, respectively. The equivalent damping
coefficient D can be calculated using Prony analysis (refer to Section 7.2 for details
of Prony analysis). The last term in (2.11) has a physical interpretation of interest.
The net tie-line power exchange PtieCOI i ,COG can be rewritten in the alternative form
of (2.4). By knowing the virtual impedances, the term Ptie
COI i ,COG and (2.11) can be
evaluated. Dividing (2.11) by (2.10) gives
34 2 MG Penetrated Power Grid Modeling
Δ f COI i = Δ f COG 2 13
M COI i ΔPmech − ΔPelec
∂ f COI i
∂ f COG
df COI i
ΔPi in (2.13) is nonzero only for the disturbed area. Moreover, the term
df COG
describes the sensitivity of the COI frequency to the COG frequency. Multiplying
df COI i
by the COG frequency variations for a given energy mismatch ΔfCOG gives
df COG
area’s i COI frequency changes.
Rearranging terms in (2.13) results in:
Dividing (2.16) by (2.11) and following the same procedure as in (2.15) gives:
Mj ΔP j − Ptie
ij
Δf j = Δ f COI i 2 17
M COI i 2πD
1+ Δ f COI i
M j ΔP j − Ptie
COI i ,j
While (2.15) ties each COI frequency to the COG frequency dynamics, (2.17)
connects each bus frequency (local frequency) to the COI dynamics. This means
that the frequency dynamics of each bus can be tied to the COG frequency using
simple algebraic equations. In the other words, (2.15) and (2.17) illustrate how the
frequency dynamics propagate in the system.
Two-Area System
Figure 2.3 shows a single-line diagram of this system. All the generating units are
modeled with sixth-order synchronous machine models with excitation systems
[18]. Each generator is equipped with a simple turbine-governor model of
Figure 2.4. The total system load is 2734 MW; the disturbance considered is the
shedding of 1400 MW load, i.e. 14 pu at bus 14 in Area 2.
Area 1 Area 2
1 10 20 3 101 13 120 11
Gen3
Gen1
110
Gen2 Gen4
4
2 12
14
Speed Pmech
1 1 1 + ST3 1 + ST4
r 1 + STs 1 + STc 1 + ST5
Figure 2.4 Simple Turbine Governor model (r = 25, Ts = 0.1, T3 = 0.0, T4 = 1.25, T5 = 5.0,
Tc = 0.5, Tmax = 1.0).
Using (2.15) and (2.17), the frequency deviations at bus 1 in Area 1 can be
expressed in terms of the COG frequency as
1 06
Δf1 = Δ f COG 2 18
1 + 0 06Δ f COI 1
Similarly, for bus 12 in Area 2, one can find that
1 78
Δ f 12 = Δ f COG 2 19
1 + 0 93Δ f COI 2
The significant difference between (2.18) and (2.19) stems from the fact that the
fault takes place in Area 1. Figure 2.5 shows the transient behavior of synchronous
machines, obtained by a conventional washout filter, a frequency divider [8], and
the frequency propagation approach in (2.15) and (2.17). For comparison pur-
poses, the time constant of the washout filter Tf is set to the default value 0.01 [8].
Additional insight into the ability of the method to characterize the slow system
dynamics can be obtained from the modal analysis of the reduced-order COG rep-
resentation. Table 2.1 compares the frequency of the two slowest modes of the full
1.02
f1 [pu]
1.01
1
0 5 10 15
Frequency propagation
1.01 Frequency divider
1
0 5 10 15
Time [s]
Figure 2.5 Frequency responses of two-area system following the loss of 14 pu of load at
bus 14.
2.3 Power Grid Modeling 37
Table 2.1 Comparison of two slowest modes for original model and equivalent model.
model with those of the COG-based equivalent model. As shown, the maximum
error is less than 5%. This example suggests that simple algebraic equations of
(2.15) and (2.17) can estimate local frequencies from the COG-based model of
Figure 2.2 with high accuracy.
14 Area 3 Area 1
66 8
41 40 48 47 60
1
25 26 29
53
28
10 2 61
27
62 9
30
31 3
1
63 18 17
24
46 38 32 16
33 15 22
Area 5 9 21
34 4
14 19 58
49 35 6
42 36 5 20
8 6 12 56 23
67 64 57 4
15 45 12
51 5
59
50 44 37 11 13 7
7 54
2
65 10
52 43 13
39
55
68
3
Area 4 16 Area 2
Figure 2.6 Single line diagram of the NYNE system showing coherent areas and their
interconnections.
38 2 MG Penetrated Power Grid Modeling
Numerical results comparing the estimated frequency are presented below. Four
contingency scenarios (CSs) are considered in this analysis:
•• CS1:
CS2:
Loss of 6000 MW at bus 37 in Area 2
Three-phase fault at line 31–62, cleared in 12 cycles
•• CS3:
CS4:
Three-phase fault at line 52–68 followed by the outage of generator G16
Outage of line 50–52 resulting in system instability
Figures 2.7 and 2.8 illustrate the frequency and tie-line power dynamics at key
system locations for contingency scenarios CS1 and CS2. The frequency of buses
59 (Area 1), 1, 63 (Area 2), and 47 (Area 3) and the net power flows to Areas 2 and 4
are selected for analysis.
In each case, the exact frequency behavior, obtained by a time-domain (T-D)
simulation program, is used as a basis for evaluating the accuracy of the estimation
methods. In the studies described below, (2.4) is employed to estimate the net
power flow to areas in the equivalent system; the net power flow to Area 1 and
Area 4 in the physical system is calculated by
and
where Ptie
k,m represents the power flow across the tie-line connecting bus k, located
in the study area, to bus m, located in neighboring areas.
Careful analysis of the numerical results in Figure 2.8 reveals that the frequency
dynamics of bus 63 in response to contingency CS2 shows some discrepancy with
the actual frequency response. Such discrepancies may result from system topol-
ogy changing conditions modifying the estimated virtual reactances. In the basic
framework, the inputs to the COG system are selected as bus voltage magnitudes,
angles, and frequencies, obtained using a T-D simulation program. As a result, the
COG dynamics could be updated or renewed using various strategies. In the below
given results, three main strategies (STs) to compute the virtual reactances in (2.8)
and (2.9) are compared:
Table 2.2 compares the accuracy of the COG-based model in compliance with
the aforementioned strategies for estimating first-swing oscillations. As shown
in the table, by recalculating the virtual reactances, the accuracy of the proposed
framework can be enhanced even for system topology changing disturbances.
2.3 Power Grid Modeling 39
(a)
1.015
1.01
f1 [pu]
Actual value
Washout filter
1.005
Frequency divider
Frequency propagation
1
0 5 10 15
1.015
1.01
f59 [pu]
1.005
1
0 5 10 15
Time [s]
(b)
1.012
1.01
1.008
f59 [pu]
1
0 1 2 3 4 5 6
Time [s]
(c)
–4 Net power flow to Area 4 in the system
Net power flow to Area 4 in the equivalent system
Ptie [pu]
–4.5
–5
0 1 2 3 4 5 6 7 8 9 10
Time [s]
Figure 2.7 System response for contingency scenario CS1; (a) frequency responses of
NYNE system, and (b) detail of frequency response showing the COI frequency; (c) net power
flows to Area 4.
Figure 2.9 compares the frequency responses for the aforementioned strategies
with those of actual response and frequency divider approach [8]. The results con-
firm the high accuracy of the frequency propagation approach to estimate both the
transient and mid-term dynamics.
40 2 MG Penetrated Power Grid Modeling
(a)
1.01
f63 [pu]
0.99
0 5 10 15
1.002
Actual frequency
f47 [pu]
Frequency propagation
1 Washout filter
0.998
0 5 10 15
Time [s]
(b)
1.5
1
Ptie [pu]
0.5
Net power flow to Area 1 in the system
0 Net power flow to Area 1 in the equivalent system
–0.5
0 1 2 3 4 5 6 7 8 9 10
Time [s]
Figure 2.8 (a) Frequency responses of NYNE system for contingency scenario CS2, (b) net
power flows to Area 3.
To further verify the theoretical basis behind the computation of fictitious reac-
tances, Table 2.3 compares the ratio of the calculated impedances for different volt-
age magnitudes, including 1 pu voltage and the exact voltage magnitude after one,
two, and three cycles, following the inception of fault. The results suggest that volt-
age magnitudes may not significantly affect the ratio of reactances.
Numerical experience shows that during a fault, when bus voltage magnitudes
deviate from nominal values, the COG-based method can still exhibit satisfactory
performance. This could be justified by the fact that the COG-based equivalent
2.3 Power Grid Modeling 41
1.005
1
f63 [pu]
Figure 2.9 Frequency responses of different strategies for contingency scenario CS2.
X tie
COI X tie X tie X tie
2 ,COG COI 3 ,COG COI 4 ,COG COI 5 ,COG
X tie
COI X tie X tie X tie
1 ,COG COI 1 ,COG COI 1 ,COG COI 1 ,COG
1
f59 [pu]
0.99
0.98
0 5 10 15
1
Actual frequency
Frequency propagation
f63 [pu]
0.98
0 5 10 15
1.05
V27 [pu]
0.95
0 5 10 15
1
V63 [pu]
0.9
0.8
0 5 10 15
Time [s]
the COG-based equivalent model provides consistent results with eigen analysis
results of the original (unreduced) system. This result also suggests that the
approach can be effectively applied when characterizing long-term system
behavior.
To further confirm the validity of the model, Figures 2.10 and 2.11 show the
performance of the method for SC3 and SC4 resulting in system instability. The
method is seen to accurately characterize system instability for both voltage and
frequency signals.
During transients associated with cascading events, the stiffness of the COG-
based method mitigates the variations of the center of angles and bus voltages. This
is shown in Figure 2.11 which demonstrates that while such variations degrade the
performance and accuracy of the proposed method to characterize short-term
behavior, the overall trend of the unstable frequency oscillation is accurately
visualized.
2.3 Power Grid Modeling 43
1.01
1.005
1
f [pu]
0.995
0.99
0.985
0 0.2 0.4 0.6 0.8 1
Time [s]
1.002
Actual Local Frequency
1 Frequency Propagation
0.998
f [pu]
0.996
0.994
0.992
0.99
0 0.2 0.4 0.6 0.8 1
Time [s]
Table 2.5 Comparison of frequency dynamics using the washout filter and proposed
method.
2470 MW load shedding at bus 52, and tripping of 4000 MW generation at bus 68.
Numerical results demonstrate the accuracy of the COG-based method to capture
slow system motion.
1
0.998
0.996
0 5 10 15
1.002
f105 [pu]
1
0.998
0 5 10 15
1.005
f135 [pu]
0.995
0 5 10 15
Time [s]
Frequency propagation
1.002
f60 [pu]
T-D simulation
1
0.998
0.996
0 5 10 15
1.005
f93 [pu]
0.995
0 5 10 15
1.005
f105 [pu]
0.995
0 5 10 15
Time [s]
Figure 2.13 Frequency responses of 50-machine system including wind power facing
three-phase fault.
46 2 MG Penetrated Power Grid Modeling
the actual frequency obtained using T-D simulation. In this analysis, the best fit-
ness value (BFV), a measure of modeling accuracy, is calculated as
y−y
BFV = 1 − 2 22
y−y
where y is the frequency response obtained from the T-D simulation, ŷ is the fre-
quency response obtained from equations (2.15) and (2.17), and y is the mean of y.
Simulation results reveal that while the BFV for the COG-based equivalent model
is more than 85% in Figures 2.7–2.12, it is less than 65% for the system considering
wind farms. This illustrates the necessity of modifying the COG-based model
through stochastic approaches which are discussed in the next section.
It may also be noted that the time required to estimate local frequencies in the
frequency propagation algorithm in the presence of wind power is about 30 sec-
onds, while it is more than 500 seconds for close T-D simulation. Table 2.6 com-
pares the computational burden of frequency propagation paradigm and close T-
D simulation. The significant difference between the frequency propagation tech-
nique and T-D simulation can be justified by noting that the number of state vari-
ables in the COG-model based estimation is much less than that for T-D simulation.
Finally it should be emphasized that the results of Figures 2.12 and 2.13 suggest
that the tie-line power flows and the local frequencies converge to steady state in a
short-time horizon. This finding justifies the fact that the COG-based method
could be used to rapidly and accurately determine the post-fault stable equilibrium
points following perturbations as well as to assess the effect of remedial control
schemes on frequency dynamics.
suggest the need for modifying the deterministic COG-based model. The theory
developed in this section extends the considerations of the effect of variable gen-
erations and the associated uncertainties on the COG model formulation. Without
the loss of generality, DGs and ESSs included in a multi-MGs (MMGs) are assumed
as variable generations to realize penetrated power grids.
Using the COG concept, the given MGs integrated-power grid can be repre-
sented by a simplified equivalent model of Figure 2.14, obtained by solving a sim-
ple minimization problem of (2.8) and (2.9). Two limitations are inherent for the
representation of (2.8) and (2.9) to assess the impact of MMGs integration:
(i) uncertainties in the representation of MMGs cannot be properly incorporated
and (ii) different operating modes, i.e. grid-connected and islanded modes, cannot
be properly represented [9, 21].
60 Hz Area 2
Inertial
response
Nadir
Area 1 δCOI2,MCOI2
tie ,
PCOI 1 COG tie ,
PCOI 2 COG
δCOI1,MCOI1
COG
tie ,
PCOI
Area k j COG δCOG,MCOG
tie ,
PCOI n COG
δCOIk,MCOIk
Figure 2.14 Interconnected power system divided into areas illustrating the notion of the
COG considering MMGs and ESSs.
48 2 MG Penetrated Power Grid Modeling
COI i ,COG − ΔP i
Ptie
n n MMG
T tie
COI i ,COG = =0 2 23
i=1 i=1
ωCOG
1 V COI i V COG
min sin δCOI i − δCOG − ΔPMMG
ω
i = 1 COG X tie
COI ,COG
i
2 25
i
st 0< X tie
COI i ,COG ≤1
min ξ− − ξ +
X tie
COI ,COG
i
n n
1 V COI i V COG
st sin δCOI i − δCOG + ξ − − ξ + = ΔPMMG
ω
i = 1 COG X tie
COI ,COG i=1
i
i
−
ξ ,ξ +
≥0
COI ,COG ≤ 1
0 < X tie
i
2 26
2.3 Power Grid Modeling 49
where ξ− and ξ+ define deviations from the target value, ξ, in the negative and
positive directions. Goal programming assigns a goal value to each of the objective
measures, i.e. zero to (2.25). Undesired deviations from the target value ξ are
then minimized using (2.26). From linear programming theory, at least one of
the ξ− and ξ+ must be zero [22].
S
min
tie
ps ξs− − ξs+ 2 27
X COI s=1
i ,COG
s.t.
n n
1 V COI i V COG
sin δCOI i − δCOG + ξs− − ξs+ = ΔPMMG
ω
i = 1 COG X tie
COI i ,COG i=1
i,s
−
ξs , ξs+ ≥ 0
0 < X tie
COI i ,COG ≤1
2 28
where ps denotes the probability of each scenario and s denotes the scenario.
(2.27) subject to (2.28) reveals that the reduced equivalent model can be
characterized by
X tie
COI 1 ,COG = 0 2134; X tie
COI 2 ,COG = 0 3172 2 29
Area 2
Area 1 δCOI2,MCOI2
tie ,
PCOI1 COG tie ,
PCOI 2 COG
δCOI1,MCOI1
tie ,
COG
PCOI
Area k j COG δCOG,MCOG
tie ,
PCOIn COG
Area n
δCOIn ,MCOIn
δCOIk ,MCOIk
Figure 2.15 Interconnected power system divided into areas illustrating the notion of the
modified COG concept.
M × RoCoF = T m − T e 2 30
and hence, (2.6) would be rewritten for area i in Figure 2.15 as
k
T tie
COI i ,COG + T tie
COI i ,COI j = M × RoCoF 2 31
j=1
Considering (2.6) and (2.31), a set of n equations with n undefined fictitious reac-
tances would be derived.
XCOI1 ,COG
COG
δCOG , HCOG
Figure 2.17 compares the results of T-D simulation with those of obtained
using (2.32).
1
f1 [p.u.]
0.98
0 5 10 15
1
f2 [p.u.]
0.98
0 5 10 15
1 T-D simulation
f3 [p.u.]
COG-Based method
0.98
0.96
0 5 10 15
Time [s]
Figure 2.17 Frequency responses of two-area system for the outage of generator 4.
2.4 MG Equivalent Model 53
1.005
f1 [p.u.]
0 5 10 15
1.005
T-D simulation
f2 [p.u.]
COG-Based method
1
0 5 10 15
1.005
f3 [p.u.]
0 5 10 15
Time [s]
•• Scenario
Scenario
1:
2:
Outage of generator 16 (Figure 2.19)
Three-phase fault on line 49–52 (Figure 2.20)
•• Scenario
Scenario
3:
4:
Outage of generator 15 (Figure 2.21)
Tripping of load at bus 37 (Figure 2.22)
1
f1 [p.u.]
0.995
0 1 2 3 4 5 6 7 8
1
f5 [p.u.]
0.995
0 1 2 3 4 5 6 7 8
1
f12 [p.u.]
T-D simulation
COG-Based method
0.995
0 1 2 3 4 5 6 7 8
Time [s]
0.999
0 1 2 3 4 5 6 7
1.001
f7 [p.u.]
1 COG-Based method
T-D simulation
0.999
0 1 2 3 4 5 6 7
f13 [p.u.]
1
0.999
0.998
0 1 2 3 4 5 6 7
Time [s]
1
f1 [p.u.]
0.998
0.996
0.994
0 5 10 15
1
f9 [p.u.]
0.998
0.996
0.994
0 5 10 15
1
f13 [p.u.]
1.015
fCOI, 1 [p.u.]
1.01
1.005
1
0 1 2 3 4 5 6 7
1.015
fCOI, 2 [p.u.]
1.01
1.005
1
0 1 2 3 4 5 6 7
1.015
fCOI, 3 [p.u.]
1.01
COG-Based method
1.005 T-D simulation
1
0 1 2 3 4 5 6 7
Time [s]
frequency of Genset and the exciter that regulates the terminal voltage of the
Genset [27–30]. Details of the Genset controller are shown in Figure 2.24 [26].
Fuel command FCMD is the output of the fuel controller Figure 2.25. Using this
framework, the measured speed of the synchronous machine compares with the
reference signal to produce an error signal. The error signal is fed into a propor-
tional-integral (PI) controller to produce the torque signal [27, 28, 30].
56 2 MG Penetrated Power Grid Modeling
SS
SS SS
AC load
SS SS
ESS
DC bus SM
IC
SS SS SS
Controllable AC load
C
C
D
D
C
C
D
+
DC Load
–
WT ESS PV
Figure 2.23 Three-phase schematic representation of the UOK-MG; WT, wind turbine; PV,
photovoltaic; SS, static switch; IC, internal combustion; SM, synchronous machine.
The limiter often implements in the block diagram of Figure 2.25 to avoid unre-
alistic commands during large load transients. The resultant torque then converts
to the fuel command signal using the torque to fuel conversion ratio Ktf. The output
fuel command signal of Figure 2.25 then applies to the simplified IC engine model
of Figure 2.26 to produce mechanical power. In this way, first, the fuel command
signal converts into the torque signal. Furthermore, engine combustion delay
affects the resultant torque. Finally, the torque converts to the mechanical output
power while losses are removed from the resulting value [27, 28, 30].
The exciter mechanism of Figure 2.27 is simply represented by a first-order
transfer function. Furthermore, the output DC field voltage multiplies by the elec-
trical speed, producing the AC voltage magnitude. The implemented limiter in
Figure 2.27 represents the saturation of the DC exciter field. The output is the field
voltage of SM [27, 28, 30].
The input signal to the exciter of Figure 2.27 comes from the voltage regulator of
Figure 2.28. It could be observed from Figure 2.28 that the error signal combines
with the feedforward value, which is the expected value required for nominal volt-
age at the terminal. Feedforwarding this value allows for quicker initial conver-
gence without the integrator having to wind up [27–31].
2.4 MG Equivalent Model 57
Vreq
Ecmd +–
Controller –
Verr
mQ
ωbase
Fcmd ++ – Preq
Controller + mP +
ωerr – +
ωmeas
1
1 + τ ωs
ωmeas
0
Pmax + PI
–
–Δω
+
Pmeas
+
Δω
Pmin –
+ PI
0
Figure 2.24 Genset controller scheme; MQ, slope of Q − V droop; MP, slope of P − ω droop;
KPI, integral control gain; KPP, proportional control gain; FCMD, command fuel signal; ECMD,
exciter control signal; Pmeas, real-time value of real power; Qmeas, real-time value of reactive
power; I , line current; Δω, allowable frequency change.
●
ωref Pmeas
Figure 2.25 Fuel controller of Genset; Ktf, torque to fuel conversion ratio.
ω2meas Km
Fcmd Tm Pmec
ηthr Kcv Kfr Fcmd Time –
Delay ωmeas +
Figure 2.26 Simplified model of the IC engine; ηthr, thermal constant; Kcv, calorific value;
Kfr, fuel rate at rated speed; Km, mechanic losses constant.
Exciter Machine
Ecmd 1 Ef
Ecmd ωmeas
1 + τe s
PI 1
Vref Kvi
S
Figure 2.28 Voltage regulation diagram; Kvi, integral controller gain; Kvp, proportional
controller gain.
2.4.1.3 Inverter-Based DG
Figure 2.30 shows the block diagram representation of an inverter-based DG [31].
Owing to the turn on/off effects of high-frequency switches, output signals
have a fundamental component and higher harmonics. However, a high rate of
2.4 MG Equivalent Model 59
(a)
4500
Power [W]
3000
Simulated
1500 Experimental
500
(b)
50
frequency [Hz]
49
48
0 5 10 15
Time [s]
switching, in the range of 10 kHz, together with the implemented LC filters signif-
icantly mitigates harmonics [26, 29, 31]. Hence, the inverter, in the islanded oper-
ation mode, is represented as an ideal, balanced three-phase voltage source as
shown in Figure 2.31 [31]. Accordingly, the instantaneous three-phase bus vol-
tages can be expressed as
ab = mV DC cos ωt + θ
V ESS
bc = mV DC cos ωt + θ + 2π 3
V ESS 2 34
ca = mV DC cos ωt + θ − 2π 3
V ESS
In this model, the modulation index m in (2.34) is a scalar coefficient that con-
trols the voltage at the inverter terminals [26, 31]. Moreover, the same reasoning
that uses to represent the Genset controller in the islanded mode can be extended
to control inverter-based sources. However, the controller command signals are
PT CT
m LF3
Inverter
Controller θ(t) Firing System
+
VDC
CF3
Figure 2.30 Inverter-based DG block diagram; m, modulating index; θ(t), angle for the voltage at the inverter terminals.
2.4 MG Equivalent Model 61
inverter frequency and voltage modulation index (Figure 2.32) [32–34]. Further-
more, in Figure 2.32, conventional droop characteristics implement using fre-
quency and voltage droop coefficients [35].
• Step 3: Metering of the injected power of MG to the host grid, i.e. PMG, for each
operating point in Step 2
PI Vreq
Kvp
m + +
–
+ –
Kvi
S
ωbase
Preq
1 ++ –
θ (t) mP +
S +
0
Pmax + PI
–
–Δω
Pmeas +
+
Δω
Pmin –
+ PI
0
f ai f 1
2 35
PMG s + ci PMG 1 ci
s+
ai ai
where a, c are constant parameters which may be defined by Prony analysis; refer
to Section 7.2 for details of Prony analysis. Equation (2.35) follows the same char-
acteristics as the classical swing equation of the form
2.4 MG Equivalent Model 63
(a)
6
Experimental
P [KW]
Simulated
4
(b)
0.5
Q [KVar]
0
–0.5
0 3 6 9 12
Time [s]
Figure 2.33 Experimental and simulated waveforms for real and reactive power output for
inverter-based DG operation in the UOK-MG. (a) active power, (b) reactive power.
f 1
2 36
PMG M i s + Di
1 ci
Hence, (2.35) with a reinterpretation of ≜ M; ≜ D could be employed to
ai ai
calculate MG inertia constant. A conceptual scheme of the adopted model is
shown in Figure 2.34.
Generally stated, Figure 2.34 suggests that MG response, from an upward point
of view, would be mapped onto the conventional synchronous generator. Accord-
ingly, the classical swing equation (2.36) represents the MG dynamics.
• Step 5: Return to Step 2, repeat until the prespecified number of operating points
are considered
• Step 6: Extract relationship, using curve fitting tools, between inertia constant
and committed DGs.
Diesel Generators
Solar Photovoltaics
Power grid
While event 1 refers to the importing of power to the MG, event 2 refers to the
islanding of the MG from the grid. It is evident that, between events 1 and 2, the
host grid provides the power of the MG with inertia constant. Following (2.36) for
the injected power of the host grid to the MG in the transient period, in
Figure 2.36, gives
ωMG 0 37
2 37
ΔPMG 0 27 + 0 63s
and in the steady state, (2.37) would be rewritten as
ωMG
= 1000 2 38
ΔPMG
where ωMGs and ΔPMGs are the frequency at the point of common coupling and the
injected power of the MG, respectively. Equations (2.37) and (2.38) suggest that the
MG mimics the behavior of a synchronous generator and a constant power source in
the transient and steady-state periods, respectively. This brings a first-order circuit
response to a pulse function in mind. Therefore, the same reasoning that uses to
assess the first-order circuit in response to a pulse function could be adopted to
approximate the effects of MG. In this way, one could represent the classical swing
equation of MG as
2.4 MG Equivalent Model 65
Calculation of M by
using Modal analysis
N
i = k?
Y
M = f(PGenset)
End
df t
M MG = Tm t − Te t u ζ − u υ 2 39
dt
where ζ and υ (ζ < υ) are DGs redispatching and islanding times, respectively.
Also of interest, event 2 refers to the islanding of the MG where droop charac-
teristic affects frequency response. The ratio of the grid power variation to the
steady-state frequency deviation of the islanded MG would be defined as the droop
characteristic. One could write this for Figure 2.36 as
0 − 1000
ΔP 5000 pu
R= = = 1 17 2 40
Δf 49 83 − 50 Hz
Finally, a relationship between inertia constant of (2.37) and the MG capacity
should be derived to facilitate assessing of impact of penetration level on the
grid dynamics. For this purpose, the MG inertia, obtained by (2.37) for several
66 2 MG Penetrated Power Grid Modeling
1 2
4
2
P [KW]
1
0
–1
–2
Grid
50.2 MG
Wind
f [Hz]
50
49.8
49.6
0 6 12 18 24 30
t [s]
Figure 2.36 Experimental waveforms for real power and frequency output for MG, wind
turbine, and host grid.
operating points, maybe plotted against the ratio of the Genset to the MG capacity
Sn in Figure 2.37. In Figure 2.37, Sn, is defined as
0.8
0.7
0.6
0.5
H [s]
0.4
0.3
0.2
0.1
0.07 0.12 0.17 0.22 0.27 0.32 0.37 0.42 0.47
Sn [pu]
PGenset PGenset
Sn = = 2 41
PMG PGenset + PESS + PPV + PWT
Actually, Sn in (2.41) changes by the recommitment of the constituent Gensets.
For this purpose, two Gensets with the rated capacities of 5 and 10 kW and with,
respectively, 0.15-second and 0.21-second inertia constants are employed. To fur-
ther generate data sets for model derivation, an analog simulator, with inertia con-
stant of 0.1 seconds, is utilized. The figure reveals that there is a direct relationship
of form
H MG = 1 6109Sn − 0 0644 2 42
with root square (R2) of 0.954.
2.5 Summary
This chapter deals with the dynamic equivalencing of MGs-integrated power sys-
tems. Several methods are introduced to model the host grid and the distribution
network. The host grid modeling relies on the notion of COG dynamics in mechan-
ics to estimate local and global frequency behavior. Using this framework, the
power system dynamic behavior is represented by an equivalent model in which
the geographical areas interact with the COG through fictitious interconnectors
from which the inherent dynamics of power frequency transients are explained
using fundamental physics principles. The relationship between the frequency
of the COG and the motion of local centers of angle is determined and expressions
to compute local frequency deviations following major disturbances are derived.
Afterward, experimental-based models of constitute DGs of the UOK-MG in
islanded mode are derived. The basic description of the elements, as well as exper-
imental validation of the results, makes the finding of the chapter comprehensive.
A realistic model of the MG operating in grid-connected mode is derived to realize
a high penetrated grid in Chapters 3, 4, and 6.
References
1. Resende, F. and Peas Lopes, J. (2007). Development of dynamic equivalents for
microgrids using system identification theory. In: Power Tech, 2007 IEEE
Lausanne, 1033–1038. IEEE.
2. Bevrani, H., Ghosh, A., and Ledwich, G. (2010). Renewable energy sources and
frequency regulation: survey and new perspectives. IET Renewable Power
Generation 4 (5): 438–457.
68 2 MG Penetrated Power Grid Modeling
32. Gao, Y. and Ai, Q. (2018). A distributed coordinated economic droop control
scheme for islanded AC microgrid considering communication system. Electric
Power Systems Research 160: 109–118.
33. Karami, M., Seifi, H., and Mohammadian, M. (2016). Seamless control scheme for
distributed energy resources in microgrids. IET Generation, Transmission &
Distribution 10 (11): 2756–2763.
34. Venkataramanan, G., Illindala, M., Houle, C., and Lasseter, R. (2002). Hardware
Development of a Laboratory-Scale Microgrid Phase 1: Single Inverter in Island
Mode Operation. NREL Report No. SR-560-32527 Golden, CO. National
Renewable Energy Laboratory.
35. Chang, C.-C., Gorinevsky, D., and Lall, S. (2014). Dynamical and voltage profile
stability of inverter-connected distributed power generation. IEEE Transactions on
Smart Grid 5 (4): 2093–2105.
71
The increasing complexity and operation of the electric power grids with high
levels of microgrids (MGs) penetration together with higher loading conditions
require the development of efficient control and analysis techniques with the abil-
ity to extract the relevant system behavior, identify abnormal behavior, and design
corrective measures. This need is due to the distributed nature of diverse renew-
able energy resources in the system and more complex dispatching strategies.
In this chapter, the analysis of the impact of high levels of MG penetration on
power system stability is discussed from various points of view such as frequency
stability, small-signal stability and voltage performance. Methods for interpreting
system dynamics in terms of simplified system representations are developed, and
criteria to determine maximum penetration levels are given using deterministic
and statistical sensitivity analyzes. Extensions to this framework to capture the
impact of large-scale wind and solar photovoltaic (PV) farms are discussed in later
sections of this book.
3.1 Introduction
3.1.1 Motivation
Modern power grids face new technical challenges arising from the increasing
penetration of distributed energy resources (DERs) and distributed generation
(DG) sources, visualized through the MG concept. While low penetration of
MGs has a negligible influence on the stability of the host power system, high pen-
etration levels of MGs may significantly affect system stability and raise several
reliability concerns at the transmission level [1–4]. The increased demand for elec-
trical power, as well as environmental concerns and pressing need to reduce
dependence on fossil fuel, on the other hand, has forced many power system uti-
lities to set an ambitious target for the deployment of DERs/DGs.
Renewable Integrated Power System Stability and Control, First Edition.
Hêmin Golpîra, Arturo Román-Messina, and Hassan Bevrani.
© 2021 John Wiley & Sons, Inc. Published 2021 by John Wiley & Sons, Inc.
72 3 Stability Assessment of Power Grids with High Microgrid Penetration
The integration of multiple MGs into the bulk power system requires the devel-
opment of new analysis tools and methodologies, to assess various aspects of
power system dynamic behavior and transmission limitations. These include
(i) determining the maximum levels of DGs/MGs generation in the system,
(ii) identifying the best system locations to deploy new DERs, and (iii) assessing
MGs’ control and dispatching strategies to improve system reliability and stability.
The first issue is addressed in this chapter. Discussion of issues two and three are
deferred until Chapters 4, 6, and 8.
Because modern DGs utilize inverter-based systems to interconnect with the
grid, their interactions with the bulk power system are different from those of con-
ventional synchronous generating units. One critical issue is the reduction of sys-
tem inertia, especially associated with inverter-based DERs [5–7]. Inertia
reduction resulting from significant penetration of utility-scale MGs and wind
and solar PV farms renders system dynamics faster and thus jeopardizes system
stability and reliability [8, 9].
In this chapter, a systematic methodology for the analysis of the impact of
distributed MGs on power system dynamic behavior is proposed. The analysis
of system dynamics is facilitated by the adoption of a simplified power system
model, in which frequency dynamics are represented explicitly through sensitivity
relationships.
First, a brief review of the subject is presented. Some important definitions asso-
ciated with the system frequency response are introduced.
• Step 3: Calculate the sensitivity of the MGs-integrated power grid dynamics with
respect to the base conventional system dynamics.
• Step 4: Represent the MGs-integrated power grid dynamics based on the base
conventional system behavior using sensitivity factors of Step 3. The above steps
are summarized in the flowchart in Figure 3.1.
Extracting of RoCoF,
frequency nadir,
and frequency evolution
Sensitivity factors
calculation
fpenetrated = g(foriginal)
End
2M COI d2 θ
= ωCOI T mech − T elec 31
ω0 dt 2
in which the rotor angle, θ, is defined as
θ = ωCOI t − ω0 t + θ0 32
and, therefore,
d Δ f COI
2M COI = 2π f COI T mech − T elec 33
dt
Defining now
d Δ f COI
RoCoF = 34
dt
and
MiΔ f i
i
Δ f COI = 35
Mi
the COI frequency response (3.1) can be represented in the form
76 3 Stability Assessment of Power Grids with High Microgrid Penetration
1
fi = ΔPi − 2πDi f i − Ptie,ij ; i, j = 1, …, n, i j 3 11
2πM i j
for area i [9]. In (3.11), Ptie,ij and n are the tie-line power between areas i and j and
the number of areas, respectively.
According to the NERC guideline, the RoCoF is defined as the rate of frequency
deviation during 500 ms after the inception of a fault. Using this guideline and
assuming nominal frequency as 50 Hz, one can rewrite (3.11) as
f i − 50 1
= ΔPi − 2πDi f i − Ptie,ij 3 12
05 2πM i j
3.2 Frequency Stability Assessment 77
and hence,
05
f i = 50 + ΔPi − 2πDi f i − Ptie,ij 3 13
2πM i j
More precisely, Ptie,ij in (3.13) deviates from the pre-fault value and can be
explained using Taylor approximations, namely
rewrite (3.14) as
Fi f 1 , f 2 , …, f j + C i = 0; j = 1, 2, …, n, i = 1, 2, …, n − 1 3 16
The nth equation to complete the set of (3.16) is formulated based on the COI
response (3.1). The frequencies of the areas can then be computed by solving (3.16)
together with (3.1) for the MG-integrated power system. Furthermore, the RoCoF
for the MG-integrated power system can be calculated using (3.10).
ΔPL − 2πT ij Δ f 1
e − p 2 t − e − p1 t
j
Δfi t = − 3 19
Mi p1 − p2
in the time domain. In (3.19), p1 and p2 are the poles of (3.18). Furthermore, at the
end of the time interval of interest, the frequency deviation is represented by
ΔPL − 2πT ij Δ f 1
e − p2 t + ΔT
− e − p1 t + ΔT
j
Δ f i t + ΔT = −
Mi p1 − p2
3 20
Delta frequency detection criterion imposes that the difference between (3.20)
and (3.19) for the main grid as well as the MG-integrated power grid must fulfill
ΔPL − 2πT ij Δ f j 1
e − p2 t − e − p1 t
Mi p1 − p2
ΔPL − 2πT ij Δ f j 1
− e − p2 t + ΔT − e − p1 t + ΔT < 0 3 3 21
Mi p1 − p2
Following the same procedure as that of (3.10) specifies the frequency deviation
over a given time window for the MG-integrated power grid. For this purpose, all
the variables have specific values except for βi. The steady-state frequency devia-
tions for both the base and reduced systems are employed to calculate βi. In this
way, one could write
Δ f steady-state
β2i = β1i × 1i
3 22
Δ f steady-state
2i
Equation (3.22) describes the frequency bias of the MG-integrated power system
as a function of the main system characteristics.
14 Area 3 Area 1
66 8
41 40 48 47 60
1
25 26 29
53
28
10 1 2 61
27
62 9
30
31 3
11
63 18 17
24
46 38 32 16
33 15 22
Area 5 9 21
34 4
19 58
49 14
35 6
42 36 5 20
8 6 12 56 23
67 64
12
57 4
15
51 45 5
59
50 44 37 11 13 7
7 54
2
65 10
52 43 13
39
55
68
3
Area 4 16 Area 2
Figure 3.2 Single line diagram of the 68-bus system showing coherent areas and their
interconnections.
simulation, for various penetration levels, including 5.6% and 8%. The MG-
integrated system is realized by utilizing the MG equivalent model (2.39), devel-
oped in Chapter 2, i.e.
df t
M MG = Tm t − Te t u ζ − u υ 3 23
dt
Also of interest, the center of gravity (COG) concept, introduced in Chapter 2,
can be employed to estimate the local frequencies shown in Figure 3.3.
1
1 5.6% MG penetration
Without MG
0.999 0.999
8% MG penetration
0.998
0.998
f [Hz]
0.997
0.996
0.995
0 5 10 15
Time [s]
According to (3.16) and for the penetration levels of 5.6%, one could write:
0 11 f 1 + 0 33 f 2 − 0 02 f 3 − 0 15 f 4 + 0 32 f 5 = 0 092 3 24
0 01 f 1 + 0 43 f 2 − 0 17 f 3 − 0 27 f 4 + 0 71 f 5 = 0 312 3 25
0 67 f 1 + 0 03 f 2 + 0 07 f 3 − 0 83 f 4 + 0 21 f 5 = 0 762 3 26
0 85 f 1 + 0 12 f 2 + 0 37 f 3 − 0 29 f 4 + 0 54 f 5 = 0 341 3 27
To complete the set of equations (3.24)–(3.27), the COI frequency dynamics
MiΔ f i
d
Mi
2M COI = 2π f COI T mech − T elec 3 28a
dt
is employed. Further manipulation of (3.28a) leads to
Mif i
2M COI = 2π f COI T mech − T elec 3 28b
i
Mi
i
Table 3.1 Frequency dynamics error for NYNE test system under 5.6% penetration level.
Scenario Disturbance Nadir error (%) RoCoF error (%) 15-seconds error (%)
1 G12 ~0 1.99 ~0
2 G13 0.3 1.15 0.11
3 L14 0.32 1.22 0.23
4 L15 0.68 0.46 0.18
5 G16 1.00 0.15 0.10
6 L37 ~0 ~0 0.17
7 L42 1.34 1.04 0.27
8 L52 1.18 0.46 0.12
Table 3.2 Frequency dynamics error for NYNE test system under 8% penetration level.
Scenario Disturbance Nadir error (%) RoCoF error (%) 15-seconds error (%)
1 G12 ~0 0.81 ~0
2 G13 0.13 ~0 ~0
3 L14 ~0 0.38 ~0
4 L15 0.68 0.69 0.41
5 G16 ~0 ~0 0.03
6 L37 ~0 0.09 0.07
7 L42 0.74 0.21 0.07
8 L52 1.98 0.12 0.72
are incorporated into (3.1)–(3.22) to establish a systematic methodology for the cal-
culation of the maximum allowable penetration levels of MGs. A schematic depic-
tion of the calculation process is given in Figure 3.4.
Given Fault
M MG = 1 6109Sn − 0 0644 3 31
PGensets
where Sn = . Here, MMG, PGenset, and PMGs are the MGs inertia, the gener-
PMGs
ation of Genset units, and the generation of the MGs, respectively. To realize high
penetration levels of MGs in the power system, multi-MGs (MMGs) structure is
considered. It is assumed that each MG in the MMGs configuration has the same
structure as that of Figure 2.23, but with the different commitments of the loads
and sources. Simulation results suggest that yet there is a direct relationship, with
Root Mean Square Error (RMSE) equal to 0.2165, between the MMGs inertia and
the ratio of the GENSets to the MMGs capacities. It should be noted that power
systems are robust enough so that small variation of the MMGs parameters, caused
by the aforementioned RMSE, may not significantly affect the system perfor-
mance. Therefore, the direct relationship of the form (3.31) can be employed to
quantify the penetration levels of MGs.
To introduce the proposed methodology, assume that the same fault is applied to
both, the main and the reduced systems. Accordingly, dividing (3.8) by (3.7) gives:
M2 d Δ f 2 f
= 2 3 32
M1 d Δ f 1 f1
Multiplying both sides of (3.32) by Δf1 results in
M2 d Δ f 2 Δf1
Δf1 = f 3 33
M1 d Δ f 1 f1 2
d Δf2
where the term describes the sensitivity of frequency deviation for the MG-
d Δf1
integrated power system to the frequency deviation for the original system.
The minimum permissible inertia can be calculated by utilizing (3.33). Formally,
d Δf2
Δ f 1 in (3.33), as the frequency change of the MG-integrated power system,
d Δf1
is set to the maximum permissible frequency deviation, i.e. 800 mHz. Further, Δf1
and f1 are the post-fault frequency deviation and frequency for the original system,
respectively. Moreover, the frequency f2 is set to 49.2 Hz. Initializing (3.33) by the
aforementioned values gives:
M2 Δf1
08 = × 49 2 3 34
M1 f1
and then
Δf1
M2 = M1 × 61 5 3 35
f1
Equation (3.35) calculates the minimum rotational inertia which beyond it, the
system frequency tends to deviate from the permissible value. It is noteworthy that
in the case of frequency increment, (3.35) can be rewritten as the form
M2 Δf1
08 = × 50 8 3 36
M1 f1
then
Δf1
M2 = M1 × 63 5 3 37
f1
In deriving (3.35) and (3.37), the amplitude of the disturbance and the kinetic
energy of the rotating masses, as two important factors affecting frequency dynam-
ics [35], are taken into account.
Here, it should be noted that (3.35) and (3.37) are derived for a single-area power
system or the aggregated inertia for all the areas. However, to calculate the min-
imum permissible rotational inertia in each area, (3.11) and (3.12) should be used.
84 3 Stability Assessment of Power Grids with High Microgrid Penetration
Under these considerations, all the parameters in (3.12) are set to the permissible
values. More precisely, (δi − δj − δi0 − δj0) is set to the maximum permissible phase
difference between areas, fi − fj is set to the maximum permissible frequency devi-
ation and f i − f j will be replaced by an aggregated model of noncoherent gen-
erators, namely
1 Mi + M j Mi Mj
fi −f j = Pmi − Pei − Pmj − Pej 3 38
2π 2M i M j 2M i M j 2M i M j
Afterward, the same procedure as those of (3.35) and (3.37) can be followed to
calculate the maximum permissible inertia reduction.
ΔPL −2πT ij Δ f
≤1
M 2i
j 1
p'1 −p'2 e −p'2 t −e −p'1 t − 1
p'1 −p'2 e −p'2 t + 15 −e −p'1 t + 15
3 41
To calculate the minimum rotational inertia, (3.41) is tied to the associated
standard. In this way, Δfj is set to the maximum allowable frequency deviation
in the viewpoint of the transient or the steady-state value (whichever is the larger).
Setting Δfj to the maximum permissible value is a pessimistic assumption that
guarantees the reliable and secure operation of the system for any operating con-
dition. Moreover, (3.41) suggests that except for M and D, the droop characteristic
3.3 Maximum Penetration Level: Frequency Stability 85
50
49.98
f2 (Hz)
49.96
49.94 COI = 4
COI = 3.90, with MG in area 2 only
COI = 3.80, with MG in all areas
49.92 COI = 3.70, with MG in all areas
0 1 2 3 4 5 6 7 8 9 10
Time (s)
about 49.96 Hz. Also of interest, the maximum phase differences between areas in
(3.12) are calculated using the DSA tool, as shown in Table 3.3.
Moreover, the term Pmj − Pej in (3.38) is set to zero which satisfies the acceptable
operation in the worst-case condition. On the other hand, fi − fj in the MG-
integrated power system is set to the maximum permissible value, i.e. 1 Hz.
Accordingly, one can write:
0 15M 1 M 2 + 0 07M 1 M 3 − 0 1M 2 M 4 − 0 03M 1 M 5 = 0 3 50
0 06M 2 M 3 + 0 01M 1 M 4 − 0 12M 3 M 4 − 0 05M 3 M 5 = 0 891 3 51
0 08M 3 M 1 + 0 06M 3 M 2 − 0 07M 1 M 4 − 0 09M 4 M 5 = 0 711 3 52
0 13M 4 M 2 + 0 16M 1 M 4 − 0 04M 2 M 5 − 0 03M 1 M 2 = 0 951 3 53
Furthermore, the minimum permissible COI constant, COIoverall, defines the last
equation, namely
50 − 49 59
COI overall = M 1 61 5 3 54
49 59
or in the final form of
COI overall = 0 51 M 1 = 0 51 3 66 = 1 86 3 55
Solving the set of (3.50)–(3.55), we obtain the minimum inertia on each area of
the system, as reported in Table 3.4.
The minimum permissible inertia in compliance with the 15-second rolling win-
dow, on the other hand, is calculated. Representing the MG-integrated power sys-
tem based on the original system characteristics is done using (3.22), namely
1
D2i +
R2i 0 58
= = 1 16 3 56
1 05
D1i +
R1i
then
1 1
D2i + = 1 16 D1i + 3 57
R2i R1i
where 0.58 and 0.50 are the steady-state frequency deviations and the maximum
allowable deviation, respectively. Substituting (3.57) in (3.41) leads to a time-
variant equation with one degree of freedom, M2. The extracted relationship
between the inertias of the original and the MG-integrated power systems in
the form of
M 2 = 0 94 ln M 1 − 0 23 3 58
is employed to calculate M2. Initializing (3.58) with the actual values gives
M 2 = 0 94 ln 4 − 0 23 = 1 07 3 59
The minimum permissible rotational inertia of area 2 is then defined using
(3.47), namely
M 2,Critical = max 1 89, 1 07, 0 4 = 1 89 3 60
Note that 0.4 in (3.60) comes from (3.46). Equation (3.60) suggests that the fre-
quency nadir imposes a lower limit on the rotational inertia. Making use of (3.60)
into (3.48) and (3.49) leads to the maximum penetration levels of 11.34% expressed
in percentage in terms of the overall system capacity.
The same procedure as those of (3.50)–(3.60) is redone for eight scenarios related
to different disturbances. The numerical results are reported in Table 3.5. The
3.3 Maximum Penetration Level: Frequency Stability 89
results suggest that the 15-second rolling window could not affect the maximum
penetration levels of MGs, as the frequency reaches steady state in less than
10 seconds.
To further confirm the validity of the model, the IEEE 50-machine test system is
used. The system is an approximate model of an actual power system. Table 3.6
reports the minimum permissible rotational inertia and the associated maximum
penetration levels of MGs for eight different scenarios. In Table 3.6, the main factor
that imposes a lower limit on the penetration levels of MGs is underlined.
Aφi = λi φi
3 61
ψ i A = λi ψ i
ωi
fi = 3 63
2π
σi
ςi = 3 64
σ 2i + ω2i
In (3.62), a positive real part, i.e. σ i > 0, corresponds to an oscillation with the
increasing amplitude. As the complex pole (3.62) moves towards the right half-
plane (RHP), the damping of the system worsens [41]. Here, the damping ratio
determines the decline rate of a specific mode.
The problem of interest is to study how the critical modes of a system are affected
by penetration levels of MGs. Critical modes of the system are those within the
frequency range of 0.01–2 Hz and damping of less than 10% [41]. For analysis pur-
poses, the simple yet efficient method of [41] is applied to the MG-penetrated
system.
The analysis framework can be summarized as follows:
• Step 1: Calculate the critical modes of the conventional power grid, with no pen-
etration of MGs, using eigenvalue analysis;
• Step 2: Repeat the procedure as Step 1 for the system with different MGs pene-
tration level;
• Step 3: Compare the results to assess the impact of MGs penetration level on the
small-signal stability;
Penetration level (%) Real part Imaginary part Frequency Damping ratio
30% MGs penetration levels. By displacing those generators, the overall system
inertia dramatically reduces, and hence the damping ratio of the adversely
impacted mode significantly decreases [41].
To further confirm the validity of the framework, sensitivity analysis is carried
out corresponding to the reported mode in Table 3.8. The sensitivity analysis is
used to analytically demonstrate the detrimental impacts of the MGs penetration
levels on the small-signal stability. The ith eigenvalue sensitivity analysis with
respect to the inertia of the jth generator is expressed as
∂A T
∂λi ψ i ∂H j
ϕi
= 3 65
∂H j ψ i ϕi
T
Table 3.9 reports a summary of the sensitivity of the critical mode to the inertia
variations of the conventional generators that are being displaced with the MGs in
the 20% penetration levels. As the damping of the system modes is determined by
the real part of the eigenvalues, the real part is presented in Table 3.9.
The negative real part sensitivity of the eigenvalues in Table 3.9 shows the
adverse impact of high MGs penetration levels on system damping. The sensitivity
analysis results corroborate the results derived from the full eigenvalue analysis of
the system under various MGs penetration levels.
3.5 Maximum Penetration Level: Small-Signal Stability 93
55 4.96 −0.0213
56 4.16 −0.0189
59 4.32 −0.0200
62 2.91 −0.0103
63 2.00 0.0065
64 1.17 0.0094
• Step 1: Extract oscillation modes of the conventional power grid with no pene-
tration of MGs, using eigenvalue analysis.
•• Step 4: Determine the damping ratios for different MGs penetration levels;
Step 5: Extract a relationship between the MGs penetration levels and the damp-
ing ratio of the form
ςi = g PMGs 3 66
• Step 6: Set the left-hand side of (3.66) equal to zero and solving for PMGs, one can
find the maximum penetration levels of MGs.
10
9
Damping ratio [%]
8
7
6
5
4
3
0 5 10 15 20 25 30
Penetration level [%]
Figure 3.6 Extracted relationship between damping ratio and penetration level.
It is well known that MGs are realized in the low and medium voltage levels.
Therefore, following the calculation of maximum penetration levels of MGs in
3.6 Voltage-Based Realization of the MG-Integrated Power Grid 95
Sections 3.5 and 3.6, this section tries to determine the maximum active power that
MGs can inject into each bus of the distribution system without causing steady-
state voltage violations.
V = V 0 + ΔV P + ΔV Q 3 70
where ΔVP and ΔVQ are voltage changes in response to the variations of the
injected active and reactive powers, respectively. The first step to calculate ΔVP
and ΔVQ in (3.70) is to derive the voltage sensitivities related to the active and reac-
tive power injections.
ΔP J Pθ J PV Δθ
= 3 71
ΔQ J Qθ J QV ΔV
The elements of the Jacobian matrix (J) represent the sensitivities of the power
variations, i.e. ΔP, ΔQ, to the voltage variations, i.e. ΔV, Δθ.
By rearranging the terms in (3.72) with the use of the definition of the reduced
Jacobian matrix JRPV, one obtains
−1
ΔV P = J RPV ΔP 3 73
96 3 Stability Assessment of Power Grids with High Microgrid Penetration
It should be noted that inversion of JQθ is feasible only if all the buses are mod-
eled as PQ buses. This is the case for distribution systems, where the slack bus is the
only voltage control bus [44]. Moreover, MGs in the steady-state studies are for-
mally represented by PQ buses since they do not contribute to the voltage control
of the system.
where JRQV is the reduced Jacobian matrix, which represents the sensitivity of
voltage magnitude with respect to the reactive power injection variations. As
the penetration levels of MGs express in the term of active power, one could
substitute
ΔQ = ΔP × tan cos − 1 pf 3 76
in (3.75) to derive
−1
ΔV Q = J RQV ΔP × tan cos − 1 pf 3 77
and then
−1 −1
V = V 0 + J RPV + J RQV × tan cos − 1 pf ΔP 3 79
By defining
−1 −1
SPQ = J RPV + J RQV × tan cos − 1 pf 3 80
V = V 0 + SPQ ΔP 3 81
Considering the acceptable steady-state voltage, the maximum permissible MGs
penetration level in each bus of the distribution system is defined by
3.6 Voltage-Based Realization of the MG-Integrated Power Grid 97
ΔV
ΔP = 3 82
SPQ
Indeed (3.82) specifies the capacity of MGs which could be installed in the stud-
ied bus without violating the steady-state voltage.
29 30 31 32 33 34 35 36
48 49 50 51
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28
67 68 69 70
54 55 56 57 58 59 60 61 62 63 64 65 66
52 53
37 38 39 40 41 42 43 44 45 46 47
(a)
1
Maximum active power injection per bus [pu]
0.8
0.6
0.4
0.2
0
0 10 20 30 40 50 60 70
Bus number
(b)
0.98
0.96
V [pu]
0.94
0.92
0.9
0 10 20 30 40 50 60 70
Bus number
Figure 3.8 (a) Maximum active power injection per bus and (b) voltage magnitude per bus.
3.7 Summary 99
pf = 1
pf = 0.95
1 pf = 0.90
Maximum active power injection per bus [pu]
0.8
0.6
0.4
0.2
0
0 10 20 30 40 50 60 70
Bus number
Figure 3.9 Maximum active power injection per bus for various power factor.
maintain the voltage magnitude of every bus within the allowable range of
0.95–1.05 pu, for the case without DGs.
Considering the unity power factor, Figure 3.8a demonstrates the maximum
active power injection per bus in the system. For this purpose, the minimum per-
missible voltage magnitude in (3.80) considers being 0.9 which is an acceptable
value for an emergency condition. Figure 3.8b shows the voltage magnitude per
bus after installing the maximum capacity on each bus.
Also of interest, Figure 3.9 depicts the effect of power factor on the maximum
active power injection per bus. It appropriately demonstrates the effectiveness
of (3.80).
3.7 Summary
system equations which, in turn, make the proposed frameworks completely inde-
pendent of the test cases.
References
1. Hatziargyriou, N., Asano, H., Iravani, R., and Marnay, C. (2007). Microgrids: an
overview of ongoing research, development, and demonstration projects. IEEE
Power Energy Magazine 5 (4): 78–94. https://doi.org/10.1109/MPAE.2007.376583.
2. Bevrani, H., François, B., and Ise, T. (2017). Microgrid Dynamics and Control.
New York: Wiley.
3. Lasseter, R.H. (2002). Microgrids. In: 2002 IEEE Power Engineering Society Winter
Meeting. Conference Proceedings (Cat. No. 02CH37309), vol. 1, 305–308. IEEE.
4. Lasseter, R.H. and Paigi, P. (2004). Microgrid: a conceptual solution. In: Power
Electronics Specialists Conference, 2004. PESC 04. 2004 IEEE 35th Annual, vol. 6,
4285–4290. IEEE.
5. Golpîra, H., Atarodi, A., Amini, S. et al. (2020). Optimal energy storage system-
based virtual inertia placement: a frequency stability point of view. IEEE
Transactions on Power Systems 35 (6): 4824–4835.
6. Golpîra, H. and Messina, A.R. (2018). A center-of-gravity-based approach to
estimate slow power and frequency variations. IEEE Transactions on Power Systems
33 (1): 1026–1035.
7. Milano, F., Dörfler, F., Hug, G. et al. (2018). Foundations and challenges of
low-inertia systems. 2018 Power Systems Computation Conference (PSCC): IEEE,
pp. 1–25.
8. Golpîra, H., Seifi, H., Messina, A.R., and Haghifam, M.R. (2016). Maximum
penetration level of micro-grids in large-scale power systems: frequency stability
viewpoint. IEEE Transactions on Power Systems 31 (6): 5163–5171. https://doi.org/
10.1109/TPWRS.2016.2538083.
9. Ulbig, A., Borsche, T.S., and Andersson, G. (2014). Impact of low rotational inertia
on power system stability and operation. IFAC Proceedings 47 (3): 7290–7297.
10. Vittal, E., Keane, A., and O’Malley, M. (2008). Varying penetration ratios of wind
turbine technologies for voltage and frequency stability. In: Power and Energy
Society General Meeting-Conversion and Delivery of Electrical Energy in the
21st Century, 2008 IEEE, 1–6. IEEE.
11. Lalor, G., Mullane, A., and O’Malley, M. (2005). Frequency control and wind
turbine technologies. IEEE Transactions on Power Systems 20 (4): 1905–1913.
12. Folly, K. and Sheetekela, S. (2009). Impact of fixed and variable speed wind
generators on the transient stability of a power system network. In: Power Systems
Conference and Exposition, 2009. PSCE’09. IEEE/PES, 1–7. IEEE.
References 101
13. Mitra, A. and Chatterjee, D. (2013). A sensitivity based approach to assess the
impacts of integration of variable speed wind farms on the transient stability of
power systems. Renewable Energy 60: 662–671.
14. Fernandez-Bernal, F., Egido, I., and Lobato, E. (2014). Maximum wind power
generation in a power system imposed by system inertia and primary reserve
requirements. Wind Energy 18 (8): 1501–1514.
15. Doherty, R., Mullane, A., Nolan, G. et al. (2010). An assessment of the impact of
wind generation on system frequency control. IEEE Transactions on Power Systems
25 (1): 452–460.
16. Yan, R., Saha, T.K., Modi, N. et al. (2015). The combined effects of high penetration
of wind and PV on power system frequency response. Applied Energy 145: 320–330.
17. Ahmadyar, A.S., Riaz, S., Verbič, G. et al. (2018). A framework for assessing
renewable integration limits with respect to frequency performance. IEEE
Transactions on Power Systems 33 (4): 4444–4453.
18. Keyhani, A. and Chatterjee, A. (2012). Automatic generation control structure for
smart power grids. IEEE Transactions on Smart Grid 3 (3): 1310–1316.
19. Chen, X., Lin, J., Wan, C. et al. (2019). A unified frequency-domain model for
automatic generation control assessment under wind power uncertainty. IEEE
Transactions on Smart Grid 10 (3): 2936–2947. https://doi.org/10.1109/
TSG.2018.2815543.
20. U. O. Handbook (2009). P1–Policy 1: load-frequency control and performance [C],
Version: v3. 0 rev, vol. 15, no. 01.04.
21. Golpîra, H. (2019). Bulk power system frequency stability assessment in presence
of microgrids. Electric Power Systems Research 174: 105863.
22. Golpîra, H. and Bevrani, H. (2019). Microgrids impact on power system frequency
response. Energy Procedia 156: 417–424.
23. Karimi, H., Karimi-Ghartemani, M., and Iravani, M.R. (2004). Estimation of
frequency and its rate of change for applications in power systems. IEEE
Transactions on Power Delivery 19 (2): 472–480.
24. Golpîra, H., Messina, A.R., and Bevrani, H. (2019). Emulation of virtual inertia to
accommodate higher penetration levels of distributed generation in power grids.
IEEE Transactions on Power Systems 34 (5): 3384–3394.
25. Illian, H.F. (2011). Frequency Control Performance Measurement and Requirements.
Berkeley, CA: Lawrence Berkeley National Laboratory.
26. NERC (2013). Standard BAL-003-1 – frequency response and frequency bias
setting.
27. Bevrani, H. (2008). Robust Power System Frequency Control. New York: Springer.
28. Golpira, H. and Bevrani, H. (2011). Application of GA optimization for automatic
generation control design in an interconnected power system. Energy Conversion
and Management 52 (5): 2247–2255.
102 3 Stability Assessment of Power Grids with High Microgrid Penetration
29. Golpira, H., Bevrani, H., and Naghshbandy, A.H. (2012). An approach for
coordinated automatic voltage regulator–power system stabiliser design in large-
scale interconnected power systems considering wind power penetration. IET
Generation, Transmission & Distribution 6 (1): 39–49.
30. Bevrani, H. and Hiyama, T. (2017). Intelligent Automatic Generation Control. USA:
CRC Press.
31. Chow, J. and Rogers, G. (2000). Power system toolbox. Cherry Tree Scientific
Software. https://www.ecse.rpi.edu/~chowj/PST_2020_Aug_10.zip (accessed
November 2020).
32. Wilson, D.H., Hay, K., and Rogers, G.J. (2003). Dynamic model verification using a
continuous modal parameter estimator. In: Power Tech Conference Proceedings,
2003 IEEE Bologna, vol. 2, 6. IEEE.
33. Rogers, G. (2000). Power System Oscillations. Norwell, MA: Kluwer.
34. Miller, N., Shao, M., Venkataraman, S. et al. (2012). Frequency response of
California and WECC under high wind and solar conditions. In: Power and Energy
Society General Meeting, 2012 IEEE, 1–8. IEEE.
35. Baggini, A.B. (2008). Handbook of Power Quality. New York: Wiley Online Library.
36. Abraham Ellis, R.W., Zavadil, B., Jacobson, D., and Piwko, R. (2012). 2012 Special
Assessment Interconnection Requirements for Variable Generation. NERC Report,
pp. 6–17.
37. PLECS (n.d.). Web address. http://www.plexim.com (accessed November 2020).
38. MAPLE (2018). Web address. https://maplesoft.com (accessed November 2020).
39. DSA Tools (n.d.). Dynamic security assessment software. Powertech Labs Inc.
http://www.dsatools.com/ (accessed November 2020).
40. Kundur, P., Balu, N.J., and Lauby, M.G. (1994). Power System Stability and Control.
New York: McGraw-Hill.
41. Eftekharnejad, S., Vittal, V., Heydt, G.T. et al. (2013). Small signal stability
assessment of power systems with increased penetration of photovoltaic
generation: a case study. IEEE Transactions on Sustainable Energy 4 (4): 960–967.
42. Messina, A.R. (2009). Inter-area Oscillations in Power Systems: A Nonlinear and
Nonstationary Perspective. Germany: Springer Science & Business Media.
43. Ugwuanyi, N.S., Kestelyn, X., Thomas, O. et al. (2020). A new fast track to
nonlinear modal analysis of power system using normal form. IEEE Transactions
on Power Systems 35 (4): 3247–3257. https://doi.org/10.1109/
TPWRS.2020.2967452.
44. Ayres, H., Freitas, W., De Almeida, M., and Da Silva, L. (2010). Method for
determining the maximum allowable penetration level of distributed generation
without steady-state voltage violations. IET Generation, Transmission &
Distribution 4 (4): 495–508.
45. Baran, M.E. and Wu, F.F. (1989). Optimal capacitor placement on radial
distribution systems. IEEE Transactions on Power Delivery 4 (1): 725–734.
103
4.1 Introduction
Modern power grids face new technical challenges arising from the increasing
penetration of power-electronic-interfaced loads, DG, and microgrids (MGs). Par-
amount among these is the inertia requirement challenge, as inverter-connected
renewable sources are increasingly replacing SGs. Unlike the conventional SGs,
inverter-based DGs do not provide physical inertia to support the grid. Therefore,
Renewable Integrated Power System Stability and Control, First Edition.
Hêmin Golpîra, Arturo Román-Messina, and Hassan Bevrani.
© 2021 John Wiley & Sons, Inc. Published 2021 by John Wiley & Sons, Inc.
104 4 Advanced Virtual Inertia Control and Optimal Placement
since SGs are gradually replaced by DGs, transmission system operators (TSOs) are
faced with the issue of lack of inertia, which intrinsically leads to a large rate of
change of frequency (RoCoF) and power variations in the grid. As a result, the
power system is prone to frequency and power fluctuations, and the design of
relays, including RoCoF-based relays, should be reconfigured.
As suggested by analytical results in Chapter 3, rotational inertia reduction in
the grid may adversely affect frequency dynamics, small-signal stability, and sys-
tem control and leads to degrade the performance of traditional control schemes.
This, in turn, may result in large dynamic deviations and, potentially, load shed-
ding, and instability. Accordingly, the increasing penetration of inverter-
interfaced DGs motivates the need to develop additional ancillary control services
to improve system dynamics and stability.
To address this issue, the concept of the virtual synchronous generator (VSG),
virtual synchronous machine, or synchronverter has been proposed [1, 2]. It is
shown that by adding short-term energy storage to emulate the kinetic energy
of a rotating mass and mimic the swing equation of an SG in the control scheme,
inverters can also provide inertia support for the grid to restrain its frequency fluc-
tuation, in the same way as an SG. Since the principle of these concepts is similar,
for convenience sake, all these inverters are referred to as VSG in this chapter.
Manipulation of the inertia constant by the VSGs could enhance system dynam-
ics and stability. This could be realized using inertia-based control schemes.
Indeed, inertia provision plays an important role in designing advanced control
schemes in modern power grids. However, the successful implementation of
advanced control schemes not only depends on the nature of the designed scheme
but also the adequacy of the inertia provision sources. In this way, after a discus-
sion on the application of VSGs in power system stability and performance
improvement, optimal placement of dispatchable inertia besides the associated
realization mechanisms would be emphasized in this chapter and afterward,
the inertia-based advanced control schemes are addressed in Chapter 6. Without
the loss of generality, while DGs (such as wind farms), high-voltage direct current
(HVDC) systems, and energy storage systems (ESSs) could emulate virtual inertia,
the ESSs are considered as the source of virtual inertia in this chapter.
Power grid control must provide the ability of electric power systems to regain a
state of operating equilibrium after being subjected to a physical disturbance, with
most system variables, i.e. frequency, voltage, and angle, bounded so that
4.2 Virtual Synchronous Generator 105
practically the entire system remains intact. Thus, the main control loops are
known as frequency control, voltage control, and power oscillation damping
control.
Additional flexibility may be required from various control levels so that the sys-
tem operator can continue to balance supply and demand on the modern power
grids. The contribution of DGs and renewable energy sources (RESs) in regulation
task refers to the ability of these resources to regulate their power output, by appro-
priate control action. This can be regarded as adding virtual inertia to the grid and
considered a solution. Virtual inertia emulation system requires the inverter to be
able to store or release an amount of energy depending on the grid frequency’s
deviation from its nominal value, analogous to the inertia of a conventional gen-
erator. This system is known as VSG and will then operate to emulate desirable
dynamics, such as inertia and damping properties, by flexible shaping of its output
active and reactive powers.
d2 d
ΔPVSG = Pm − Pe = JΔω' + DΔω = J 2 δ t + D dt δ t 41
dt
where J is the inertia coefficient, D is the damping coefficient, and δ is the rotor
angular position difference from its reference position.
The swing Equation (4.1) describes the relative motion of the rotor in respect to
the stator field as a function of time. The generator inertia reacts to the disturbance
and plays a significant role in power system stability, which is the most important
property of a synchronous machine.
106 4 Advanced Virtual Inertia Control and Optimal Placement
Primary DC/DC
Source
DC/AC
Inverter
AC/DC
VSG Control
Supervisory command
Lls Lf Zline
Primary Storage Cf
Source Device Vout(abc) Iout(abc)
BUS
Q0
Qref + +
Vbus Q Droop PI
^
E
– +
Vbase Virtual Vpwm
Impedance PWM
P0 Control θpwm abc/αβ
Virtual
Governor and 1/s θm V out(αβ) Iout(αβ)
ωm
Inertia
Pout
Power
Qout
LPF Meter
Vbus
Estimator
type of VSG control regulates the output voltage through the reactive power con-
trol loop. Owing to the absence of an inner voltage loop, the active power control
loop becomes quite simple and robust, and bus voltage deviations become smaller
and insensitive to the output impedance [3].
In the power generation part, the “Virtual Governor and Inertia” block is the
core of VSG control. In the literature, various approaches using different damping
technologies are proposed for this part. A virtual voltage drop over virtual induc-
tance Lls is generated to adjust the equivalent output reactance X of the inverter as
shown in (4.2), given by
X = ω0 LLS + L f + Lline 42
The block “Vbus estimator” estimates the bus voltage from the measurement of
output voltage and current to provide a common reference for the block “Q
Droop,” which makes a simple linear function between voltage and reactive
power [1]. To emulate the steady-state operation of an SG, the governor model
is usually simulated in a dispatchable VSG-based power source, as shown in (4.3):
Pin = P0 − k p ωm − ω0 43
To mimic the dynamics of an SG, the swing equation should be emulated. If the
effect of damper windings is omitted, the swing equation of an SG can be
expressed as:
dωm
Pin − Pout = Jωm 44
dt
Combining (4.3) and (4.4) yields
dωm
P0 − Pout = Jω0 + k p ωm − ω0 45
dt
In interpreting the control structure in Figure 4.2, note that the VSG control
emulates Equation (4.5) through the “Virtual Governor and Inertia” block.
Apart from the VSG control topology shown in Figure 4.2, several different VSG
controls exist in the literature. The VSG control schemes can be categorized into
three main groups based on the nature of the output reference from the VSG [4].
be applied in this approach, but unlike the first group, the utilizing voltage
command allows this VSG to function in stand-alone mode.
3) Power references-based VSG control: This group emulates the inertia response
by tracking the grid frequency without implementing any SG model. The group
is called “power references from VSG,” as the current reference corresponding
to a given power reference is used to control the inverter.
Some applications of VSGs in power systems have been reported in [5–10]. For
example, the VSG concept was successfully implemented in [5] for multiterminal
HVDC systems to suppress low-frequency oscillation and enhance the power oscil-
lation damping performance of the AC/DC system. A dual VSG-based modular
multilevel matrix converter (M3C) control scheme for frequency regulation sup-
port of a remote AC grid via low-frequency AC transmission system is proposed
in [6]. In [7], a VSG is used to enhance the performance of a stand-alone gas engine
generator.
Furthermore, several studies considering the application of VSG in PV systems
[8], wind power generation [9], permanent magnet synchronous generators
(PMSGs) [10, 11], and doubly fed induction generators (DFIGs) have been
reported. Besides, the VSG control can be applied to other grid-tied inverters, such
as those in the ESSs [12], bidirectional battery chargers of electrical vehicles (EVs)
providing vehicle-to-grid (V2G) services [13], and voltage source converters (VSCs)
in high-voltage DC transmission system [5, 14–17]. Generally, the VSG-based DC–
AC converter becomes a standard interface for smart grid integration [18].
As emphasized above, the VSGs with emulated inertia capabilities may play a
critical role in reducing the RoCoF and the frequency nadir of the power system
[19]. Owing to this promising feature, it attracts a lot of attention in recent years.
These works are mainly targeted at improving the damping of a VSG via fixed
parameter or adaptive parameter methods [1], improving its parameter tuning
[20], fault ride-through ability, power quality, and applying this concept in various
types of grid-interfaced power electronics devices [6, 7, 21, 22].
Line5_7 L3 Line6_9 L4
5 6
L1 Line4_5 Line4_6 L2
4
T1
1
G1
steam turbines, and exciters are taken from [23], except minor changes mentioned
in [24]. Power system stabilizers (PSSs) are not included in the system, and the
rated power of G1, G2, and G3 are 512, 270, and 125 MVA, respectively.
Three case studies to evaluate the grid integration performance of VSG are con-
sidered. In Case 1, it is assumed that all three generators in Figure 4.3 are SGs. Case 2
is the same as Case 1, except that a cluster of coherent grid-following inverter-based
DGs with the same overall rated power replaces G3. For the clarity of illustration,
the DGs are aggregated into a single DG of 125 MVA. Finally, Case 3 is similar to
Case 2, except that the VSG control is used to control the inverter-based DGs.
In these studies, the VSG is designed to have the same inertia and droop coef-
ficient as G3 in Case 1, while the equivalent reactance from the transient emf to the
adjacent bus of the G3 node is 0.3 pu, and the damping ratio of VSG is set to 0.9.
377
376.8
Case 1
376.6 Case 2
Case 3
376.4
377
ωm2 (rad/s)
376.8
376.6
376.4
Po1 (MW)
78
76
74
72
Po2 (MW)
168
166
164
studied cases further verifies the conclusion obtained with Scenario 1. The VSG
improves frequency nadir and reduced transient RoCoF. On the other hand, Case
1 shows completely different dynamics, because changing the power command of
an SG is delayed by its slow governor and turbine responses.
377
ωm1 (rad/s)
376.5 Case 1
Case 2
376 Case 3
377
ωm2 (rad/s)
376.5
376
Po1 (MW)
80
Po2 (MW)
170
Figure 4.5 System response for a 20-MW step decrease in G3 power command.
which is then converted to HVDC by a wind farm converter station for transmission
to a load center converter station (Figure 4.6).
The VSC-HVDC is currently considered as the market leader for wind integra-
tion at distances greater than 60–80 km largely due to its established use in point-
to-point bulk power transfer [26]. The HVDC system is broadly applied for
Main Grid
Power
converter
HVDC System
Wind farms
Small power
stations Rural/remote
Load centers
point-to-point interconnection of two power systems. It is also well used for inter-
connecting the offshore wind farm with the onshore stations using a subma-
rine cable.
An HVDC link may improve the system stability of a wide area power system by
decoupling of some areas. For instance, consider two interconnected AC grids by
an HVDC link. This system exhibits a natural decoupling between the AC grids in
terms of both voltage and frequency (the AC grids are interconnected in an asyn-
chronous manner). These kinds of interconnections are primarily aimed at pre-
venting the excursions of oscillations between AC systems, for instance,
between a stronger and a weaker electric network [27]. Furthermore, the HVDC
transmission technology is currently deemed a desirable solution for long-distance
power transmission and offshore RESs.
In contrast, the VSC-HVDC systems have some weak points. The HVDC system
is sensitive to faults on the DC lines. When a fault occurs on the DC side of a
VSC-HVDC system, the insulated-gate bipolar transistor (IGBT) cannot control
and freewheeling diodes work as a bridge rectifier. It is not able to withstand large
surge currents and may be damaged before the fault is cleared. Some solutions are
proposed; however, additional control and switching devices are needed.
An HVDC system uses the VSC stations that, from a physical inertia point of
view, their dynamic behavior is quite different from that of the SGs. Thus, it makes
the frequency stability problem in the connected AC grid more serious than the
past. For this reason, the HVDC systems are expected to participate in the grid
frequency support. This issue is more significant when it is known that the inte-
gration of RESs instead of SGs reduces system inertia. Consequently, the system
frequency stability and dynamic performance are affected. We can consider this
issue from a different perspective. One typical application of HVDC transmission
systems is to supply a weak or passive/islanded grid far away from the main grid.
With constant power control, the output power from the HVDC-VSC station is
constant, and no frequency support is provided to the weak grid. When a signif-
icant part of the power in a local gird is supplied by this station, the system inertia
becomes lower, and changes in generation or load may cause large-frequency
deviations possibly leading to system instability [14]. In response to this challenge,
it could be a promising solution to use the VSC station to provide frequency sup-
port to the local grid.
There are several works investigating system frequency support through HVDC
systems [5, 14, 27–29]. The main idea is to use the VSG concept for grid-tied HVDC
converters to emulate the classical SG, thus providing virtual inertia and system
frequency support, as shown in Figure 4.7.
4.3 Dispatchable Inertia Placement 113
VDC
DC/AC
Inverter
VSG Control
Supervisory command
Figure 4.7 HVDC system as a primary source for grid frequency support.
Not only an insufficient level of inertia but also its heterogeneous distribution may
render system dynamics faster. This fact, along with the need to economically
keeps the system secure, and makes the optimal placement of virtual inertia as
a key factor [30]. Accordingly, optimal placement of virtual inertia can be studied
in two perspectives: (i) frequency stability and (ii) small-signal stability.
frequency response as well as steady-state metrics as the main metrics utilized for
system resilience analysis, neglecting the impact of inertial response from RESs on
system stability. In this direction, some recent studies have investigated the effect
of reduced inertia on frequency stability and transient stability of the power system
[35–42]. In parallel with these efforts, some recent works, [43–50], investigate ways
in which virtual inertia could be emulated, including appropriate control of wind
turbines and ESSs. In [51], a framework that addresses various aspects of inertia
emulation and control, including how virtual inertia emulation and its location in
the system impacts system stability, is proposed. Some questions about the heter-
ogeneous inertial profiles and how the associated negative impacts are reduced by
inertia emulation have been raised in [38]. Further, Poolla et al. [30] propose an
H2-based performance metric to determine the optimal placement of virtual
inertia. The determination of the optimum size of ESS to provide the primary
frequency control is addressed in [52, 53]. However, due to the lifetime concerns,
the ESSs cannot effectively participate in the primary regulation. In practice, the
ESSs are dispatched using an optimal control strategy, designed to optimize the
state-of-charge (SOC) range and the lifetime constraints. Some research works,
such as in [54], deal with the optimal placement of virtual inertia in power systems
considering network structure. These approaches utilize DC power flow to incor-
porate network structure into the model.
The main limitation of these approaches is their reliance on static considerations.
It is well known, however, that dynamic frequency indices, such as the RoCoF and
frequency nadir, are important parameters to assess frequency performance and the
development of protection schemes. Based on these considerations, the following
subsections formulate the problem of finding virtual inertia locations in terms of
dynamic as well as static metrics. The main step toward the optimal placement
of virtual inertia in a power grid is to analyze its effects on the frequency behavior;
and this could be realized by the appropriate modeling of the virtual inertia.
s
δ = ωs t − ω0
46
Mωs = T m t − T e t
s
where δ , ωs, and ω0 are the mechanical rotor angle, the mechanical rotor angular
speed, and the initial angular speed, respectively; M, Tm(t), and Te(t) are the inertia
constant, the mechanical input torque, and the electrical output torque, respec-
tively [58]. Taking the slow electromechanical behavior of the ESS into account,
the associated dynamics could be represented by (4.6). The problem of interest,
however, is to calculate the equivalent inertia constant and the mechanical input
torque.
In what follows, a data-driven approach in which the uncertain behavior of the
ESS is accounted for in the swing Eq. (4.6) is proposed.
where, ak, Φk, ωk, and w(n) are the magnitude and the initial phase angle, har-
monic frequency in radius, and additive white noise, respectively. In the model,
ak and ωk are assumed to be deterministic and unknown, and Φk is unknown
and assumed to be random and uniformly distributed in [−π, π]. Alternatively,
the model (4.7) can be expressed in the form of noisy complex exponentials as [59]:
K
P n = Ak e jnωk + w n 48
k=1
1
PMUSIC ejω = z 2 49
i = K+1 eU si
where si is the eigenvectors associated with the noise subspace that is orthogonal
to the signal eigenvector e = [1 ejω ej2ω ej(z−1)ω]T, and eU denotes the complex-
conjugate transpose; z is the dimension of space spanned by P(n). It is worth
emphasizing that PMUSIC(ejω) in (4.9) does not relate to any real power spectrum;
rather, the only purpose of this pseudo-spectrum is to generate peaks whose fre-
quencies correspond to those of the dominant frequency components. This feature
makes the MUSIC approach interesting to develop an equivalent model based on
dominant modes.
For a given signal of interest and using (4.7)–(4.9), the model eigenvalues can be
calculated. By knowing the eigenvalues and because the impulse response is the
inverse Laplace transform of eigenvalues, one could represent a signal of interest
with a predefined model of (4.6). For this purpose, suppose the impulse response of
system is I(t); for the input signal x(t), i.e. x(t) = Tm(t) in (4.6), one could write y(t),
i.e. ω, as:
∞
y t =x t I t ejωt dt 4 10a
t=0
where
I t = y t ∗PS 4 10b
and PS in (4.10b) is the pseudo-spectrum of the signal. Equation (4.10b) reveals
that I(t) is obtained from the convolution of y(t) and PS. In the modeling procedure
and by measuring the output response of the system y(t) and by knowing I(t), the
problem of interest is to calculate x(t) in (4.10a). By calculating Tm (t), the ESS
could be replaced by the SG model of (4.6). Using this framework, angular speed
ω in (4.10a) is defined based on the dominant frequency components of the
pseudo-spectrum in (4.9). Figure 4.8 gives a schematic illustration of this model.
In this plot, Figure 4.8a illustrates the process of virtual inertia emulation using the
battery ESS, while Figure 4.8b describes a simplified block diagram representation
of the equivalent model. The input to the control algorithm is the frequency at the
connection point of the inverter fcp, and Pin represents the grid injected power.
(a) Inverter
Pin
DC
Energy Storage
DC
Power grid
+ fCP
Control
–
Algorithm
(b)
y(t)
ejωt
Max
ω
(PS)
Figure 4.8 Block diagram representation of the proposed modeling process: (a) virtual
inertia emulation mechanism and (b) the proposed equivalent model; Pin, y(t), and fcp
represent the grid injected power, the frequency deviation of the ESS, and the frequency at
the point of connection of the ESS, respectively.
The main grid power deviation during the charging process in Figure 4.10 is uti-
lized to calculate the pseudo-spectrum of Figure 4.11 which, in turn, is used to
estimate the dominant frequency components in (4.9). Setting the frequency devi-
ation of the ESS in Figure 4.12 as y(t) and the pseudo-spectrum of Figure 4.11 as PS
for (4.10b) gives
s
δ = ωs − ω0
4 11
0 53ωs = 1 − e − 0 38t C − T e t
SS
SS SS
AC load
SS SS
ESS
DC bus SM
IC
SS SS SS
Controllable AC load
C
C
D
D
C
C
D
+
DC Load
–
WT ESS PV
1
Grid
3 ESS #1
P [KW]
1.5
0
–1.5
0 3 6 9 12 15
Time [s]
model (4.11) to approximate the inertial response behavior of the ESS. To exactly
mimic the frequency behavior of ESS using (4.11), the oscillatory behavior of the
adopted model can be removed using a 20-sample rolling-averaging window. This
approach averages the long-term oscillations, and hence, mitigates the oscillatory
behavior beyond the inertial response horizon. Results in Figure 4.12 show that the
dynamic behavior of the ESS, especially in the inertial response horizon, can be
4.3 Dispatchable Inertia Placement 119
10
Power [dB]
–10
–20
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Normalized Frequency
50
49.85
0 6 12 18 24 30
Time [s]
Figure 4.12 Comparison of frequency response computed from the experiment and the
equivalent frequency response models.
where tch is the charging/discharging time. The balance of the ESS, known as the
BOP, includes site wiring, interconnecting transformers, and other additional
ancillary equipment and is measured on a $/kW basis [62]. However, the LCC
is a more common metric to evaluate and compare different ESS technologies.
The annualized LCC is formulated according to (4.13) which considers operation
and maintenance costs (Co&M,a), replacement cost (CR,a), and annualized TCC:
CLCC,a = Ccap,a + COM,a + CR,a $ kW yr 4 13
in which:
T
i 1+i
CRF = T
4 14
1+i −1
Ccap,a = TCC × CRF $ kW yr 4 15
COM,a = CFOM,a + C VOM,a × ncycle × t ch $ kW yr 4 16
r − kt CR × t ch
CR,a = CRF × 1+i × $ kW yr 4 17
k=1 ηsys
where CRF, i, T, r, t, and ηsys are the capital recovery factor, interest rate and the
lifetime, the number of substitutions in a lifetime, the replacement period, and the
overall efficiency, respectively; CFOM,a and CVOM,a define the fixed and variable
operation and maintenance costs. The subscript “a” stands for “annualized” costs.
This equation gives the average hourly power that can be injected/absorbed to/
from the grid by the ESS. By substituting (4.19) into (4.13), one could write the
optimization problem as:
nESS
M ESS Sbase
minimize F M ESS
i = CLCC,a i × i
4 20a
M ESS
i i=1
3600 sec
st RoCoF i ≤ RoCoF max 4 20b
Δ f nadir i ≤ Δ f nadir max 4 20c
SOC min ≤ SOC i ≤ SOC max 4 20d
where nESS is the number of ESSs. Moreover, the SOC should remain within an
appropriate range which is addressed in (4.20d). The SOC can be calculated as fol-
lows [63, 64]:
Δt
0 ζp t dt
SOC Δt = SOC 0 − 4 21a
E ESS,rated
where
ζc p t <0
ζ= 1 4 21b
p t >0
ζd
and p(t) is battery power which gets negative values for the charging procedure
and positive values for the discharging period; EESS,rated, Δt, ζ c, and ζ d are the nom-
inal energy capacity, charge/discharge time, and charging and discharging effi-
ciencies of the battery, respectively.
measure the RoCoF [65, 66]. The ENTSO standard [65] explains that RoCoF is
allowed to get a value between 0.5 and 1 Hz/s.
To obtain dynamic frequency indices based on lower bounds of inequality con-
straints in (4.20b), the RoCoF can be defined in terms of the classical swing equa-
tion of (4.6) as [67]:
dΔf t
2M = ΔPm t − ΔPL t − ΔPtie t 4 22
dt
where ΔPm(t), ΔPL(t), and ΔPtie(t) represent mechanical power, electrical power,
and tie-line power changes, respectively. Considering the definition of RoCoF, one
could write
ΔPm t − ΔPL t − ΔPtie t
RoCoF = 4 23
2M
The Taylor series expansion of (4.23) about the independent variables of H, ΔPm,
ΔPL, and ΔPtie gives
∂RoCoF i ∂RoCoF i ∂RoCoF i ∂RoCoF i
ΔRoCoF i = ΔΔPmi + ΔΔPLi + ΔΔPtiei + ΔM i
∂ΔPmi ∂ΔPLi ∂ΔPtiei ∂ΔM i
1 −1 −1 − ΔPmi − ΔPLi − ΔPtiei
= ΔΔPmi + ΔΔPLi + ΔΔPtiei + ΔM i
2M i 2M i 2M i 2M i 2
4 24
Since the slow inherent dynamics of interest is only given by the last term in
(4.24), the other terms can be neglected. Inserting (4.23) into (4.24) gives
ΔM i
RoCoF i − = ΔRoCoF i 4 25
Mi
It follows that the minimum inertia which guarantees that the RoCoF remains
within the permitted range can be calculated as:
ΔRoCoF i, max
i, min = ΔM i, min = M i −
M 'ESS
RoCoF i
4 26
RoCoF i, max − RoCoF i
M 'ESS
i, min = Mi −
RoCoF i
where RoCoFi,max, is the maximum allowable RoCoF, and M ESSi,min represents the
minimum required inertia that should be emulated by the battery ESS in area i,
which complies with the RoCof. It equals to the difference between the desired
inertia to enforce the system to follow the standards and the present inertia con-
stant, i.e. ΔMi,min.
and the capability of the power resources to provide primary frequency response.
According to the NERC and the Union for the Coordination of the Transmission of
Electricity (UCTE) standards [68, 69], the allowed minimum frequency in a power
system during normal operation is 800 mHz. Taking the time dependence of the
governor response into account, one can approximate the frequency nadir after
a system event as [43]:
ΔPL + ΔPtie 2 T d
Δ f nadir = 4 27
4MR
where R is the extra power received from the governor and Td is the response time
of the governor.
In deriving (4.27), it is assumed that the mechanical power through the governor
increases as a linear function of time with the steady gradient R/Td [70, 71]. While
this is a conservative assumption, Great Britain and Ireland practices show that
this is the case for the power increment within 5 and 10 seconds (Td), respectively,
following a contingency [72]. Applying Taylor’s expansion to (4.27) gives
∂Δ f nadir,i ∂Δ f nadir,i ∂Δ f nadir,i
ΔΔ f nadir,i = ΔΔPLi + ΔΔPtie,i + ΔM i
∂ΔPLi ∂ΔPtie,i ∂ΔM i
ΔPLi + ΔPtiei T di ΔPLi + ΔPtiei T di
= ΔΔPLi + ΔΔPtiei
2M i Ri 2M i Ri
− ΔPLi + ΔPtiei 2 T di
+ ΔM i
4M 2i Ri
4 28
Following the same procedure as that in (4.26), one could rewrite (4.28) in
the form:
ΔM i
Δ f nadir,i − = ΔΔ f nadir,i = Δ f nadir,i − f 0 4 29
Mi
Moreover, the overall system inertia has a direct impact on the frequency indi-
ces. This means that some considerations should be made regarding overall system
inertia and, consequently (4.26) and (4.30) will be completed by adding a new
equality constraint. For this purpose, the frequency of the overall center of inertia
(COI), which should satisfy strict frequency standards, can be used to determine
the overall amount of inertia in the system as:
Δ f COI
M COI = Q = M 61 5 4 32
f COI
where ΔfCOI and fCOI represent the frequency deviation and frequency of the sys-
tem, without ESS, after the fault, respectively. Details on how to derive (4.32) are
given in Chapter 3. Formally, Equation (4.32) gives the required amount of inertia
constant which guarantees acceptable frequency dynamics of the COI. Of note that
Q would be realized by adding the emulated inertia of ESSs to the conventional
SGs inertia. Accordingly, the optimization problem (4.20) can be rewritten as:
nESS
M ESS Sbase
minimize F M ESS
i = C LCC,a i i
4 33a
M ESS
i i=1
3600 sec
st M COI = Q 4 33b
M ESS
i, min ≤ M ESS
i ≤ M ESS
i, max 4 33c
SOC min ≤ SOCi ≤ SOCmax 4 33d
where the dynamic inequality constraints (4.20b) and (4.20c) are reformulated as
the algebraic inequality constraint (4.33c) in terms of the inertia constant. This
increases the simplicity and speed of the calculations.
1 ΔPLi
βi
Ri
ΔPgi ΔPmi
+ – –
+ 1 1 + 1 Δfi
K (s)
+ ACEi ΔPCi 1 + sTgi 1 + sTti – Di + 2(Hi + hi)s
narea
Controller Governor Turbine Rotating mass
and load ∑ Tij
j=1
j≠i
ΔPtie,i 2π/
+
s
narea –
∑ TijΔfj
j=1
j≠i
Figure 4.13 Block diagram representation of control area i. βi, Ri, Tgi, Tti, and Di are
frequency bias, droop characteristic, governor time constant, turbine time constant, and
damping property, respectively.
nESS
M ESS Sbase
minimize F M ESS
i = CLCC,a i i
4 34a
M ESS
i i=1
3600 sec
st M COI = 0 053 4 34b
0 0129 ≤ M ESS
1 4 34c
0≤ M ESS
2 4 34d
0 0225 ≤ M ESS
3 4 34e
30 ≤ SOCi ≤ 80 4 34f
where for instance, the minimum inertia of area 1 in (4.34c) is calculated based on
(4.26), (4.30), and (4.31) as:
1 + 1 1870
M 'ESS
1, min = 0 08335 × − = 0 0129
− 1 1870
M ''ESS
1, min = 0 4 35
M ESS
1, min = max 0, 0 0129 = 0 0129
The parameters used in (4.34a) are summarized in Tables 4.1 and 4.2.
The results obtained, using a simple genetic algorithm (GA) with 0.05 and 0.8
mutation and crossover coefficients, respectively, from the optimization of
(4.34a) are shown in Table 4.3. To further assess the efficiency of the formulation,
Table 4.3 compares the results with those obtained in [30, 73]. Comparison results
justify the fact that the dynamic behavior of ESS can significantly affect the
optimal placement problem.
126 4 Advanced Virtual Inertia Control and Optimal Placement
Table 4.1 Economical parameters related to the optimization problem [42, 43].
Further, Figure 4.14 compares the frequency behavior and RoCoF of generators 1,
2, and 3 for three cases of interest: (i) with virtual inertia and according to the for-
mulation of (4.33a)–(4.33d), (ii) with virtual inertia and according to [73], and
(iii) with virtual inertia and according to Ref. [30]. It can be seen that while
frequency traces of Refs. [30, 73], and the formulation of (4.33a)–(4.33d) meet the
RoCoF and frequency nadir standards, (4.33a)–(4.33d) results in a less ESS capacity.
Also of interest, modal analysis of the results, as explained in Table 4.4, shows the
efficiency of (4.33a)–(4.33d) in comparison with that of Refs. [30, 73].
4.3 Dispatchable Inertia Placement 127
(a)
0
Δf1[Hz]
0
Δf2[Hz]
–0.2
–0.4
–0.6
0 2 4 6 8 10 12 14 16 18 20
0.2
0
Δf3[Hz]
–0.2
–0.4
–0.6
0 2 4 6 8 10 12 14 16 18 20
Time (s)
(b)
1
RoCoF1 [Hz/sec]
0.5
0
–0.5
–1
0 2 4 6 8 10 12 14 16 18 20
1
RoCoF2 [Hz/sec]
0.5
0
–0.5
–1
0 2 4 6 8 10 12 14 16 18 20
1
RoCoF3 [Hz/sec]
0.5
0
–0.5
–1
0 2 4 6 8 10 12 14 16 18 20
Time (s)
Figure 4.14 (a) Frequency behaviors and (b) RoCoF of generators 1 to 3 of three-area power
system for three cases: proposed formulation [30, 73].
128 4 Advanced Virtual Inertia Control and Optimal Placement
Results show that a lower emulated virtual inertia based on (4.33a)–(4.33d) not
only decreases the cost function but also provides better performance in terms of
enhanced damping.
Tables 4.3 and 4.4 and Figure 4.14 suggest that with virtual inertia, the results of
Refs. [30, 73] seem to fulfill the constraints and are almost the same as the results
in (4.33a)–(4.33d). This could be justified through the fact that the set of generator
buses for small systems includes a few members to be considered as candidates for
ESS installation. Therefore, different algorithms may differ a bit from the capacity
point of view rather than the location which in turn causes a negligible difference
in the results. To further assess the effects of virtual inertia on the frequency and
transient stabilities, two nonlinear systems are used in what follows.
Two-Area Power System In this section, the two-area power system, shown in
Figure 4.15, is considered to further demonstrate the efficiency of the optimal
placement formulation. Modeling considerations are essentially those described
in [74]; all the generating units are modeled with sixth-order synchronous
machine models with excitation systems.
Area 1 Area 2
1 10 20 3 101 13 120 11
Gen3
Gen1
110
Gen2 Gen4
4 12
2
14
The test scenario of interest is the outage of generator G4 in the first second of
the simulation. The lower bounds of virtual inertia are calculated according to
(4.34a)–(4.34f) and (4.35) as:
nESS
M ESS Sbase
minimize F M ESS
i = CLCC,a i i
4 36a
M ESS
i i=1
3600 sec
st M COI = 4 973 4 36b
0 3151 ≤ M ESS
1 4 36c
0 3881 ≤ M ESS
2 4 36d
0 ≤ M ESS
3 4 36e
30 ≤ SOCi ≤ 80 4 36f
in which the equality constraint (4.36b) reveals that
Table 4.6 Frequency indicators of two-area system before and after the application of
optimal inertia values.
With VI Proposed
Without VI method With VI [30] With VI [73]
Ref. [30]
1.002 Proposed method
Δf1[Hz]
Ref. [73]
0.998
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
1.002
Δf2[Hz]
0.998
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
1.004
1.002
Δf3[Hz]
1
0.998
0.996
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time [s]
Figure 4.16 Frequency response of generators 1, 2, and 3 of the two-area power system for
three cases: proposed formulation [30, 73].
where δmax is the maximum angle separation of any two generators in the system.
It is noteworthy that during severe faults, most ESSs, if remain connected and con-
tinue to inject active power, get saturated and cannot follow the frequency prop-
erly. It is noted here that (4.38) relies on a conservative assumption that the fault is
not a severe one, which causes ESSs to be disconnected from the grid. Moreover,
4.3 Dispatchable Inertia Placement 131
Method δmax η
the saturation of ESSs in response to sever faults is neglected. Table 4.7 demon-
strates the better performance of the proposed method in comparison with those
of Refs. [30, 73].
New York New England System The NYNE test system is used to further illustrate
the efficiency of the proposed algorithm for large-scale power systems. A single-
line diagram of the system, showing major coherent areas and their interconnec-
tions, is shown in Figure 4.17.
Five different contingency scenarios, including tripping of major generating
units and load rejection, are considered. Table 4.8 compares the results of the
optimization problem, given by (4.33a)–(4.33d), with those of Refs. [30] and
14 Area 3 Area 1
66 8
41 40 48 47 60
1
25 26 29
53
28
10 1 2 61
27
62 9
30
31 3
11
63 18 17
16 24
46 38 32
33 15 22
Area 5 34
9
4
21
19 58
49 14
35 6
42 36 5 20
8 6 12 56 23
67 64
12
57 4
15 45
51 5
59
50 44 37 11 13 7
7 54
2
65 10
52 43 13
39
55
68
3
Area 4 16 Area 2
Figure 4.17 Single-line diagram of NYNE test system showing coherent areas and their
interconnections.
132 4 Advanced Virtual Inertia Control and Optimal Placement
0.7
Scenario 1
0.6
Scenario 2
0.5 Scenario 3
Virtual Inertia
Scenario 4
0.4
Scenario 5
0.3
0.2
0.1
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Generator number
1
Ref. [30]
Proposed method
0.998 Ref. [73]
f[Hz]
0.996
0.994
0 5 10 15
Time [s]
Figure 4.19 Frequency responses of NYNE test system for different approaches.
(nominal) frequency in defining the system impedances which is far from realistic
for systems with high penetration of inverter-based ESSs. This point is successfully
addressed in (4.33a)–(4.33d) by explicitly representing the dynamic behavior of
ESSs in the problem formulation.
2 9.7516 9.5913
3 9.8617 9.9302
The effectiveness of (4.41) is now assessed for the NYNE system. For this pur-
pose, the outage of generator 1 in area 1, i.e. Scenario 1, is considered as the base
case for sensitivity analysis. The cost functions for the outage of generator 7 in area
1, i.e. Scenario 2, and generator 11 in area 2, i.e. Scenario 3, are calculated using
(4.41). Table 4.10 compares the exact results of (4.33a)–(4.33d) with those of (4.41)
which justify the effectiveness of the sensitivity analysis (4.41).
For uncertainty analysis, the equality constraint (4.33b) is represented in the
objective function (4.33a) as:
nESS
M ESS
i Sbase
minimize F M ESS
i = CLCC,a i + β M COI − Q 4 42
M ESS
i i=1
3600 sec
where β is arbitrary chosen high to enforce the results to follow the equality con-
straint (4.33b). Considering parametric uncertainty for the inertia constant MCOI,
one could write (4.42) as:
nESS
M ESS Sbase
minimize F M ESS
i = CLCC,a i i
+ β M COI + γ − Q
M ESS
i i=1
3600 sec
nESS
M ESS Sbase
minimize F M ESS
i = CLCC,a i i
+ β M COI − Q + βγ
M ESS
i i=1
3600 sec
4 43
F M ESS
i = 3 8298, 5 5112 4 45
nESS
M ESS
i Sbase
minimize F M ESS
i = CLCC,a i COI − K
+ α M ss 4 48
M ESS
i i=1
3600 sec
where α is arbitrary chosen high to enforce the results to follow the equality
constraint (4.46b). On the other hand, as the damping ratio of a specific mode
is influenced by the generating units according to their participation factors, it
seems that the participation factors should be reflected in the objective function
(4.48). More precisely, the machines with higher participation factors in the stud-
ied mode should contribute to the minimization problem (4.48) with lower
weights. In this way, one could rewrite (4.48) as:
nESS
M ESS Sbase 1 ESS
minimize F M ESS
i = C LCC,a i i
+α M+ M −K
M ESS
i i=1
3600 sec i
pf i i
4 49
where pfi is the participation factor of generator i in the critical mode of interest.
0.9
0.8
0.7
Participation Factor
0.6
0.5
0.4
0.3
0.2
0.1
0
0 2 4 6 8 10 12 14 16
Generator Number
Figure 4.20 Participation factor of the machines participating in the critical mode
detrimentally affected by high MGs penetration level.
According to (4.49) and Figure 4.20, it is expected that buses related to genera-
tors number 15 and 16 should require a larger ESSs capacity. Figure 4.21 shows the
allocation of virtual inertia among the generator buses of the system which in turn
justify the effects of the participation factor on the considered objective function.
0.35
0.3
0.25
Virtual Inertia
0.2
0.15
0.1
0.05
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Generator Number
4.4 Summary
The advanced control schemes in modern power grids rely on inertia manipulation
in the system to enhance stability. The successful development of advanced control
schemes highly depends on the adequacy of the inertia provision sources. In the
present chapter, first, the VSG concept and its applications in power grids are
explained. Afterward, dispatchable inertia is optimally placed in the system accord-
ing to dynamical metrics to enhance system stability and dynamic performance.
References
1. Bevrani, H., François, B., and Ise, T. (2017). Microgrid Dynamics and Control.
New York: Wiley.
2. Bevrani, H., Ise, T., and Miura, Y. (2014). Virtual synchronous generators: a survey
and new perspectives. International Journal of Electrical Power & Energy Systems
54: 244–254.
3. Liu, J., Miura, Y., Bevrani, H., and Ise, T. (2016). Enhanced virtual synchronous
generator control for parallel inverters in microgrids. IEEE Transactions on Smart
Grid 8 (5): 2268–2277.
4. D’Arco, S. and Suul, J.A. (2013). Virtual synchronous machines—classification of
implementations and analysis of equivalence to droop controllers for microgrids.
In: Proceedings of the 2013 IEEE Grenoble Conference, 1–7. IEEE.
5. Cao, Y. et al. (2017). A virtual synchronous generator control strategy for VSC-
MTDC systems. IEEE Transactions on Energy Conversion 33 (2): 750–761.
6. Al-Tameemi, M., Liu, J., Bevrani, H., and Ise, T. (2020). A dual VSG-based M3C
control scheme for frequency regulation support of a remote AC grid via low-
frequency AC transmission system. IEEE Access 8: 66085–66094.
7. Hlaing, H.S., Liu, J., Miura, Y. et al. (2019). Enhanced performance of a stand-alone
gas-engine generator using virtual synchronous generator and energy storage
system. IEEE Access 7: 176960–176970.
8. Mishra, S., Pullaguram, D., Buragappu, S.A., and Ramasubramanian, D. (2016).
Single-phase synchronverter for a grid-connected roof top photovoltaic system. IET
Renewable Power Generation 10 (8): 1187–1194.
9. Ma, Y., Cao, W., Yang, L. et al. (2017). Virtual synchronous generator control of full
converter wind turbines with short-term energy storage. IEEE Transactions on
Industrial Electronics 64 (11): 8821–8831.
10. Wang, Y., Meng, J., Zhang, X., and Xu, L. (2015). Control of PMSG-based wind
turbines for system inertial response and power oscillation damping. IEEE
Transactions on Sustainable Energy 6 (2): 565–574.
140 4 Advanced Virtual Inertia Control and Optimal Placement
11. Li, Y., Xu, Z., and Wong, K.P. (2016). Advanced control strategies of PMSG-based
wind turbines for system inertia support. IEEE Transactions on Power Systems
32 (4): 3027–3037.
12. Lopes, L.A. (2014). Self-tuning virtual synchronous machine: a control strategy for
energy storage systems to support dynamic frequency control. IEEE Transactions
on Energy Conversion 29 (4): 833–840.
13. Suul, J.A., D’Arco, S., and Guidi, G. (2016). Virtual synchronous machine-based
control of a single-phase bi-directional battery charger for providing vehicle-to-grid
services. IEEE Transactions on Industry Applications 52 (4): 3234–3244.
14. Guan, M., Pan, W., Zhang, J. et al. (2015). Synchronous generator emulation
control strategy for voltage source converter (VSC) stations. IEEE Transactions on
Power Systems 30 (6): 3093–3101.
15. Aouini, R., Marinescu, B., Kilani, K.B., and Elleuch, M. (2015). Synchronverter-
based emulation and control of HVDC transmission. IEEE Transactions on Power
Systems 31 (1): 278–286.
16. Dong, S., Chi, Y., and Li, Y. (2015). Active voltage feedback control for hybrid
multiterminal HVDC system adopting improved synchronverters. IEEE
Transactions on Power Delivery 31 (2): 445–455.
17. Li, C., Li, Y., Cao, Y. et al. (2018). Virtual synchronous generator control for
damping DC-side resonance of VSC-MTDC system. IEEE Journal of Emerging and
Selected Topics in Power Electronics 6 (3): 1054–1064.
18. Zhong, Q.-C. (2016). Virtual synchronous machines: a unified interface for grid
integration. IEEE Power Electronics Magazine 3 (4): 18–27.
19. Bevrani, H. (2009). Robust Power System Frequency Control. Springer.
20. Fathi, A., Shafiee, Q., and Bevrani, H. (2018). Robust frequency control of
microgrids using an extended virtual synchronous generator. IEEE Transactions on
Power Systems 33 (6): 6289–6297.
21. Jongudomkarn, J., Liu, J., Yanagisawa, Y. et al. (2020). Model predictive control for
indirect boost matrix converter based on virtual synchronous generator. IEEE
Access 8: 60364–60381.
22. Terazono, D., Liu, J., Miura, Y. et al. (2020). Grid frequency regulation support
from back-to-back motor drive system with virtual-synchronous-generator-based
coordinated control. IEEE Transactions on Power Electronics: 1. https://doi.org/
10.1109/TPEL.2020.3015806.
23. Demetriou, P., Asprou, M., Quiros-Tortos, J., and Kyriakides, E. (2017). IEEE 9-bus
modified test system. https://www2.kios.ucy.ac.cy/testsystems/index.php/ieee-9-
bus-modified-test-system/ (accessed November 2020).
24. Bevrani, H., Imanaka, M., and Kato, T. (2020). Grid dynamics shaping using
flexible controlled power converters. Energy Reports 6: 1490–1495.
References 141
25. Madariaga, A., Martín, J., Zamora, I. et al. (2013). Technological trends in electric
topologies for offshore wind power plants. Renewable and Sustainable Energy
Reviews 24: 32–44.
26. Rodrigues, S., Restrepo, C., Kontos, E. et al. (2015). Trends of offshore wind
projects. Renewable and Sustainable Energy Reviews 49: 1114–1135.
27. Castro, L.M. and Acha, E. (2015). On the provision of frequency regulation in low
inertia AC grids using HVDC systems. IEEE Transactions on Smart Grid 7 (6):
2680–2690.
28. Leon, A.E. (2017). Short-term frequency regulation and inertia emulation using an
MMC-based MTDC system. IEEE Transactions on Power Systems 33 (3): 2854–2863.
29. Rakhshani, E., Remon, D., Cantarellas, A.M. et al. (2016). Virtual synchronous
power strategy for multiple HVDC interconnections of multi-area AGC power
systems. IEEE Transactions on Power Systems 32 (3): 1665–1677.
30. Poolla, B.K., Bolognani, S., and Dörfler, F. (2017). Optimal placement of virtual
inertia in power grids. IEEE Transactions on Automatic Control 62 (12): 6209–6220.
31. Golpîra, H. and Bevrani, H. (2019). Microgrids impact on power system frequency
response. Energy Procedia 156: 417–424.
32. Lian, B., Sims, A., Yu, D. et al. (2016). Optimizing LiFePO4 battery energy storage
systems for frequency response in the UK system. IEEE Transactions on
Sustainable Energy 8 (1): 385–394.
33. Wu, Z., Gao, D.W., Zhang, H. et al. (2017). Coordinated control strategy of battery
energy storage system and PMSG-WTG to enhance system frequency regulation
capability. IEEE Transactions on Sustainable Energy 8 (3): 1330–1343.
34. Zhang, F., Hu, Z., Xie, X. et al. (2017). Assessment of the effectiveness of energy
storage resources in the frequency regulation of a single-area power system. IEEE
Transactions on Power Systems 32 (5): 3373–3380.
35. Ahmadyar, A.S., Riaz, S., Verbič, G. et al. (2018). A framework for assessing
renewable integration limits with respect to frequency performance. IEEE
Transactions on Power Systems 33 (4): 4444–4453.
36. Golpîra, H., Seifi, H., Messina, A.R., and Haghifam, M.-R. (2016). Maximum
penetration level of micro-grids in large-scale power systems: Frequency stability
viewpoint. IEEE Transactions on Power Systems 31 (6): 5163–5171.
37. Spahic, E., Varma, D., Beck, G. et al. (2016). Impact of reduced system inertia on
stable power system operation and an overview of possible solutions. In:
Proceedings of 2016 IEEE Power and Energy Society General Meeting (PESGM), 1–5.
New York: IEEE.
38. Ulbig, A., Borsche, T.S., and Andersson, G. (2014). Impact of low rotational inertia
on power system stability and operation. IFAC Proceedings Volumes 47 (3):
7290–7297.
142 4 Advanced Virtual Inertia Control and Optimal Placement
39. Wang, Y., Bayem, H., Giralt-Devant, M. et al. (2014). Methods for assessing
available wind primary power reserve. IEEE Transactions on Sustainable Energy
6 (1): 272–280.
40. Wang, Y., Delille, G., Bayem, H. et al. (2013). High wind power penetration in
isolated power systems—assessment of wind inertial and primary frequency
responses. IEEE Transactions on Power Systems 28 (3): 2412–2420.
41. Chu, Z., Markovic, U., Hug, G., and Teng, F. (2020). Towards optimal system
scheduling with synthetic inertia provision from wind turbines. IEEE Transactions
on Power Systems 99: 1-1.
42. Golpîra, H., Haghifam, M.R., and Seifi, H. (2015). Dynamic power system
equivalence considering distributed energy resources using Prony analysis.
International Transactions on Electrical Energy Systems 25 (8): 1539–1551.
43. Golpîra, H., Messina, A.R., and Bevrani, H. (2019). Emulation of virtual inertia to
accommodate higher penetration levels of distributed generation in power grids.
IEEE Transactions on Power Systems 34 (5): 3384–3394.
44. D’Arco, S. and Suul, J.A. (2014). Equivalence of virtual synchronous machines and
frequency-droops for converter-based microgrids. IEEE Transactions on Smart
Grid 5 (1): 394–395.
45. Rakhshani, E., Remon, D., Cantarellas, A.M., and Rodriguez, P. (2016). Analysis of
derivative control based virtual inertia in multi-area high-voltage direct current
interconnected power systems. IET Generation, Transmission & Distribution 10 (6):
1458–1469.
46. Hammad, E., Farraj, A., and Kundur, D. (2019). On effective virtual inertia of
storage-based distributed control for transient stability. IEEE Transactions on
Smart Grid 10 (1): 327–336.
47. Farmer, W.J. and Rix, A. (2019). Optimising power system frequency stability using
virtual inertia from inverter-based renewable energy generation. In: 2019
International Conference on Clean Electrical Power (ICCEP), 394–404. IEEE.
48. Attya, A., Anaya-Lara, O., and Leithead, W. (2018). Novel concept of renewables
association with synchronous generation for enhancing the provision of ancillary
services. Applied Energy 229: 1035–1047.
49. Poolla, B.K., Groß, D., and Dörfler, F. (2019). Placement and implementation of
grid-forming and grid-following virtual inertia and fast frequency response. IEEE
Transactions on Power Systems 34 (4): 3035–3046.
50. Esmaili, M., Ghamsari-Yazdel, M., Amjady, N., and Chung, C.Y. (2020). Convex
model for controlled islanding in transmission expansion planning to improve
frequency stability. IEEE Transactions on Power Systems https://doi.org/10.1109/
TPWRS.2020.3009435.
51. Borsche, T.S., Liu, T., and Hill, D.J. (2015). Effects of rotational inertia on power
system damping and frequency transients. In: 2015 54th IEEE Conference on
Decision and Control (CDC), 5940–5946. IEEE.
References 143
52. Fini, M.H. and Golshan, M.E.H. (2018). Determining optimal virtual inertia and
frequency control parameters to preserve the frequency stability in islanded
microgrids with high penetration of renewables. Electric Power Systems Research
154: 13–22.
53. Oudalov, A., Chartouni, D., and Ohler, C. (2007). Optimizing a battery energy
storage system for primary frequency control. IEEE Transactions on Power Systems
22 (3): 1259–1266.
54. Wogrin, S. and Gayme, D.F. (2015). Optimizing storage siting, sizing, and
technology portfolios in transmission-constrained networks. IEEE Transactions on
Power Systems 30 (6): 3304–3313.
55. Mo, O., D’Arco, S., and Suul, J.A. (2016). Evaluation of virtual synchronous
machines with dynamic or quasi-stationary machine models. IEEE Transactions on
Industrial Electronics 64 (7): 5952–5962.
56. Golpîra, H., Atarodi, A., Amini, S. et al. (2020). Optimal energy storage system-
based virtual inertia placement: a frequency stability point of view. IEEE
Transactions on Power Systems 35 (6): 4824–4835.
57. Golpîra, H., Seifi, H., and Haghifam, M.R. (2015). Dynamic equivalencing of
an active distribution network for large-scale power system frequency
stability studies. IET Generation, Transmission & Distribution 9 (15):
2245–2254.
58. Ajala, O., Dominguez-Garcia, A., Sauer, P., and Liberzon, D. (2018). A library of
second-order models for synchronous machines. arXiv preprint arXiv:1803.09707.
59. Bollen, M.H. and Gu, I.Y. (2006). Signal Processing of Power Quality
Disturbances. Wiley.
60. Manolakis, D.G., Ingle, V.K., and Kogon, S.M. (2000). Statistical and Adaptive
Signal Processing: Spectral Estimation, Signal Modeling, Adaptive Filtering, and
Array Processing. Boston: McGraw-Hill.
61. Zakeri, B. and Syri, S. (2015). Electrical energy storage systems: a comparative life
cycle cost analysis. Renewable and Sustainable Energy Reviews 42: 569–596.
62. Mongird, K., Viswanathan, V.V., Balducci, P.J. et al. (2019). Energy Storage
Technology and Cost Characterization Report. Pacific Northwest National Lab
(PNNL), Richland, WA (United States).
63. Teruo, I. (2005). State of charge calculation device and state of charge calculation
method. US Patents, US6845332B2.
64. Tang, Z.X. and Lim, Y.S. (2016). Frequency regulation mechanism of energy
storage system for the power grid. 4th IET Clean Energy and Technology
Conference (CEAT 2016), Kuala Lumpur, 1–8. doi: https://doi.org/10.1049/
cp.2016.1272.
65. European Network of Transmission System Operators for Electricity (ENTSOE)
(2016). Frequency Stability Evaluation Criteria for the Synchronous Zone of
Continental Europe – Requirements and Impacting Factors, March.
144 4 Advanced Virtual Inertia Control and Optimal Placement
66. Eto, J.H. (2011). Use of Frequency Response Metrics to Assess the Planning and
Operating Requirements for Reliable Integration of Variable Renewable Generation.
Berkeley, CA: Ernest Orlando Lawrence Berkeley National Laboratory Tech. Rep.
LBNL-4142E.
67. Golpîra, H. (2019). Bulk power system frequency stability assessment in presence
of microgrids. Electric Power Systems Research 174: 105863.
68. Ekwue, A. and Cory, B. (1984). Transmission system expansion planning by
interactive methods. IEEE Transactions on Power Apparatus and Systems 7:
1583–1591.
69. C. E. O. H. ENTSO-E (2009). P1-Policy 1: Load-Frequency Control and
Performance, ed: Tech. Rep. 2000-130-003, May 2000.
70. Chávez, H., Baldick, R., and Sharma, S. (2014). Governor rate-constrained OPF for
primary frequency control adequacy. IEEE Transactions on Power Systems 29 (3):
1473–1480.
71. Teng, F., Trovato, V., and Strbac, G. (2016). Stochastic scheduling with inertia-
dependent fast frequency response requirements. IEEE Transactions on Power
Systems 31 (2): 1557–1566.
72. Grid, N. (2019) Security and quality of supply standards. https://www.
nationalgrideso.com/industry-information/codes/security-and-quality-supply-
standards (accessed November 2020).
73. Borsche, T. and Dörfler, F. (2017). On placement of synthetic inertia with explicit
time-domain constraints. arXiv preprint arXiv:1705.03244.
74. Rogers, G. (2012). Power System Oscillations. Berlin: Springer Science &
Business Media.
75. Aien, M., Hajebrahimi, A., and Fotuhi-Firuzabad, M. (2016). A comprehensive
review on uncertainty modeling techniques in power system studies. Renewable
and Sustainable Energy Reviews 57: 1077–1089.
145
5.1 Introduction
Wide-area area voltage control of large interconnected systems has attracted con-
siderable interest in the last few decades [1–6]. Network voltage control at genera-
tors, and dynamic reactive power compensation devices placed at key system
locations, among other measures, can support voltage regulation and enhance
system transient stability and operating flexibility [7]. The increasing size and
complexity of power systems with high penetration levels of MGs/DGs distributed
along a large geographical area or located in remote zones, however, make voltage
control challenging.
As the number of modern renewable energy sources (RESs) with improved static
and dynamic reactive power capability grows, a significant challenge is to integrate
them into existing voltage control schemes as well as to develop effective coordi-
nation schemes [8–10]. Better detection and forecasting techniques through the
concept of wide-area voltage monitoring are also needed to allow full use of control
capabilities and realize wide-area monitoring structures.
In the recent past, there has been renewed interest in the use of automatic volt-
age regulation schemes for wind and solar photovoltaic (PV) farms connected to
the bulk power systems [9–11]. Studies show that static VAR compensators (SVCs)
located adjacent to large wind and PV farms can be used to maximize reactive
power reserves and improve voltage profiles. Voltage–VAR control can contribute
to the overall power system angle and voltage stability and result in improved sys-
tem operation and security. This is a subject that is receiving increasing attention.
The integration of advanced voltage–VAR controls in the wind and solar genera-
tors, on the other hand, raises several complex issues [8, 11]. First, farm-level volt-
age control introduces a hierarchical control system that needs to be optimized to
improve the overall system response to system perturbations. Further, coordina-
tion with other nearby generators or network reactive power compensation
devices may be needed as in many cases, wind farms require additional reactive
power support, especially during transient conditions.
One of the critical issues in modeling the inverter-based generators is reactive
power capability. Some interesting phenomena of voltage and reactive power
responses from solar PV generators have been observed and investigated, such as
high voltages under normal conditions, high transient voltages, and sustained oscil-
lations following a fault or change in the control characteristics of wind and solar PV
farms [1]. These problems may cause further reliability concerns such as overload of
subtransmission and distribution facilities, unexpected generation tripping for
overvoltage or under-excitation, and even system-wide transient instability.
Experience with the application of primary and secondary voltage control in
power systems in European countries shows that coordination of reactive sources
5.2 Voltage Control Areas: A Background 147
may result in enhanced system wide-area control and reliability [2, 12–14]. To
avoid undesirable interactions, voltage control at the various levels should tempo-
rally and spatially independent. This requires splitting the system into non-
interacting zones in which voltage is controlled individually.
Large-scale coordination of reactive power sources is challenging due to a large
number of control characteristics, the location and type of controllers, and the
characteristics of each device. Issues such as reserve capacity control and the effi-
cient utilization and coordination of reactive power sources must be addressed to
achieve fast automatic voltage control and keep the capacitive output margin
against system contingencies.
This chapter discusses the experience in the development of data-driven analysis
techniques to identify and update voltage control zones and the associated reactive
power resources. A systematic methodology for the identification of voltage con-
trol zones is first introduced. The proposed procedure consists of three main steps:
(i) the identification of strongly connected buses showing coherent behavior,
(ii) the identification of generators and SVCs participating in the critical zones,
and (iii) the determination of distance measures indicating relationships between
bus voltage magnitudes and reactive power sources. These methods are suitable for
large-scale applications and can be used to coordinate multiple available reactive
compensation devices, including SVCs, synchronous condensers, generator exci-
tation systems, and modern RESs equipped with closed-loop voltage control
schemes.
The design methodology is demonstrated on a complex test system with signif-
icant wind penetration in which several SVCs are used to control system voltage.
Results show that properly coordinated reactive power sources may have an
important impact on system dynamic behavior.
Recent years have witnessed the development and application of wide-area voltage
control systems with the ability to monitor voltage deviations at key transmission
buses, update set points of major closed-loop controllers, and coordinate reactive
power sources to regulate network voltages. Due to the local nature of voltage
behavior, a major issue in these hierarchical control schemes pertains to the iden-
tification of nearly independent voltage control zones.
A general review of the voltage control structures is presented in Chapter 1. In
this chapter, a brief overview of these methods, in the context of voltage monitor-
ing techniques is given. The discussion begins with a review of fundamental
concepts in the development of practical wide-area voltage control of power
148 5 Wide-Area Voltage Monitoring in High-Renewable Integrated Power Systems
systems. Then, the need for data-driven approaches to identify voltage control
zones is established.
−1 ∂Q
ΔQ = J QV − J Qθ J Pθ J PV ΔV = ΔV = J QV R ΔV 52
∂V Vo
and, therefore,
−1 ∂V
ΔV = J QV R ΔQ = ΔQ = J VQR ΔQ
∂Q V o ,Qo
where matrix J QV R is called the reduced steady-state Jacobian matrix of the system
and J VQR is a sensitivity matrix with coefficients:
∂V 1 ∂V 1 ∂V 1
∂Q1 ∂Q2 ∂Qn
∂V 2 ∂V 2 ∂V 2
∂Q1 ∂Q2 ∂Qn
J VQR = J VQRij =
∂V n ∂V n ∂V n
∂Q1 ∂Q2 ∂Qn
where n is the total number of nodes, and in the interest of simplicity, all buses are
assumed to be PQ buses.
Simplified approaches for calculating sensitivity relations are described in
Ref. [15]. Further, Ref. [16] describes alternative approaches to calculate
5.2 Voltage Control Areas: A Background 149
ΔQgi ∂Qgi ∂V HV i
= ,i k 53
ΔV pk ∂V HV i V o ,Qo ∂V pk Qo
where V HV i is the high side bus of the machine at bus i and Qgi represents reactive
∂Qgi ∂V HV i
power injections at bus i; the terms ∂V HV i , ∂V pk represent sensitivity coefficients.
∂Qgi
Numerically, the coefficients ∂V HV i can be obtained directly from the load flow
∂V HV i
Jacobian matrix; the second term, ∂V pk , represents an electrical distance and is
a subproduct of the calculation of electrical distances in Section 5.2.2. Sensitivity
matrices are real and nonsymmetrical and reflect the propagation of voltage var-
iation following reactive power injection at a bus [17]. Variations of this model
using other system formulations are described in [4, 15–17], and references
therein.
or ΔVi = αijΔVj, where, in general, αij αji. Physically, αij ≈ 1, when buses i, j are
electrically close and has a small value when they are electrically distant. Other
approaches to determining approximate sensitivity relations based on the load
flow equations are described in [4, 17].
An attenuation matrix can then be obtained from the notion of entropy or infor-
mation theory. The electrical pairwise distance, dij, between buses i and j can be
150 5 Wide-Area Voltage Monitoring in High-Renewable Integrated Power Systems
defined as dij = dji = − log(αij. αji) that has two important properties, positivity and
symmetry.
Extending this approach to the multidimensional case, one can define the matrix
of attenuation between all the buses, A, as:
α11 α12 α1n
α21 α22
A = αij = α2n 55
1) Subdivide the attenuation data A into subsets or clusters which are pairwise
disjoint and connected;
2) Select a robust node or bus that represents the average or dominant system
behavior;
3) Determine sensitivity relations between dominant bus voltage behavior and
reactive power sources.
These approaches are usually hierarchical; first, a voltage control zone associ-
ated with the largest sensitivities is determined. Then, the remaining zones are
determined using an iterative approach [1, 2, 4].
1) Pilot buses should be chosen among the strongest nodes in the system using
criteria such as the maximum short-circuit current.
2) Coupling between pilot nodes associated with different control areas should be
low to avoid interactions among control systems.
Cluster j
Inter-cluster
distances
Cluster i
α12 α23
α24
1 α13 mi 3
Cluster k
4
α34
α25
α15 α45
α35
Figure 5.1 Illustration of clusters associated with weighted, connected graphs. Each node
within cluster i has several connections (dense connection) with other members of the
cluster and fewer links (sparse connections) with nodes outside the cluster.
that they belong to the control area under analysis, the size of their reactive power
capability, and the highest electrical coupling with the pilot nodes (see Eq. (5.3)).
These methods, however, are commonly characterized by several limitations:
•• The methods do not fully recognize the dynamic nature of system behavior;
In many applications, not only frequently pilot nodes need to be reselected fol-
lowing topology changes but also the control areas and the associated reactive
power sources need to be updated;
Voltage control
Area 1
∼ ∼ Q charts
Qg Direct ...
1 clustering
Wide-area Voltage control
Dimensionality areas,
V1 ∠θ1 voltage reduction Pilot buses,
∼ Vp1 ∠θp1 monitoring Trajectory
Q-V correlation,
classification
Qgn Map visualization
...
∼ ∼
Figure 5.2 Overview of the steps in computing voltage control areas. Here, V p j represents
the set point of pilot buses, Vj and θj are the bus voltage magnitude and phases, and Qg j
represent reactive power output from major generators and network reactive power
compensation devices.
5.4 Theoretical Framework 155
knowledge about the strength of interactions between multimodal data, i.e. volt-
age and reactive power. Further, measurement techniques give some insight into
the nature of control structure and communications required for efficient imple-
mentation of practical WAMSs, especially in power systems with RESs [25].
Three main activities are of interest here:
• Bus voltage magnitudes from different physical regions may show similar
behavior thus obscuring physical interpretation. Phase information may be
needed to improve the extraction of coherent structures;
• Practical voltage control systems should only consider selected voltage system
behavior (model reduction) from which global system behavior can be
reconstructed;
dij x i , x j, t j = x i t j − x j t j 57
Distance
function
where xi(tj) and xj(tj) represent the instantaneous values of the time sequences at
time instance tj (a time slice in the spatiotemporal representation in Figure 5.1). In
practice, distances can be obtained for a time window of interest and a full distance
t1 tk tN
...
...
...
Similarity
(distance) Normalization
matrix
Vertex i v Coarse-graining
i
weight Transition • Dimensionality
wij reduction
vj probability
• Anomaly detection
matrix
• Visualization
Undirected graph
Figure 5.3 Illustration of distance (similarity) matrices and the associated undirected
graphs.
5.4 Theoretical Framework 157
matrix can then be defined as D = [dij]. In this case, however, the distance defi-
nition does not account for the time-ordering of the signals.
This allows the application of graph-clustering algorithms or coarse-grained
methods to distance (similarity) matrices.
wij = x i t − x j t
with
m
wii = wij
j=1
xi t − x j t
K = k ij = exp − , 1 ≤ i, j ≤ m 58
εi ε j
where the kernel bandwidth εi, εj controls the hop size of the random walk and can
be automatically tuned based on the distribution of matrix A. The use of other ker-
nels is discussed elsewhere.
Physically, matrix K can be interpreted as the adjacency matrix of a graph, each
of whose nodes represents one sensor or trajectory. The kernel bandwidth scale, εi,
εj modulates the notion of distance in (5.7) and can be tuned to cluster system
behavior. If εiεj maxij xi(t) − xj(t) , then for all edges {i, j}, kij ≈ 1; for low kernel
bandwidth values, εiεj 0, the pairwise distances kij become increasingly similar
and large fluctuations in the density of points in the high-dimensional space are
smoothed out.
From the matrix kernel in (5.8), one now can define a diagonal matrix D = [dij]
whose entries are the row sums of K, namely:
m
j = 1 K 1j 0 … 0
m
D = dij = 0 j = 1 K 2j … 0
m
0 0 … j = 1 K mj
m
where dii = j = 1 K ij is the degree of node xi.
The distance matrix has an interesting interpretation in terms of an undirected
probabilistic graph. In this concept, the transition probability pij from i to j can now
m
be obtained as pij = k ij k = 1 k ik. The right-stochastic Markov matrix or probabil-
ity matrix, M, can then be defined as:
k 11 k 12 k 1m
m m … m
j = 1 K 1j j = 1 K 1j j = 1 K 1j
k 21 k 22 k 2m
m m … m
M = mij = D −1
K= j = 1 K 2j j = 1 K 2j j = 1 K 2j
k m1 k m2 k mm
m m … m
j = 1 K mj j = 1 K mj j = 1 K mj
59
in which the matrix element mij = pij can be interpreted as the probability pij of
hopping from point i to point j in t steps of a discrete random walk [32].
Physically, the transition probability matrix, M, defines the random walk of a
particle on the graph. Formally, suppose that the initial probability of the particle
being at a vertex vj is poj j = 1, …, n . It follows that the probability of the trajectory
5.4 Theoretical Framework 159
vj taking the edge wij is mij poj. The extension to the multivariate case follows along
the same lines.
Paramount to the automated extraction of low-dimensional representations is
the adaptive computation of the Gaussian kernel widths, ε. In general, the scale
parameter is related to the statistics and the geometry of the data points. Following
Ref. [29], let Xj denotes a cloud of points around xi. The variance of the distance
between the point xi to all the points xj Xj is given by Ref. [30]:
2
xi − x j − X i
εi =
xi X j
Xi
where
Xi = xi X
xi − x j Xi
mij ≥ 0
j m ij = 1 for j = 1, …, n row stochastic
M − λ j I ψ j = 0, j = 1, …, n 5 10
160 5 Wide-Area Voltage Monitoring in High-Renewable Integrated Power Systems
M = D − 1K = D − 1 2
D − 1 2 KD1 2
D−1 2
= D − 1 2MsD − 1 2
5 11
and thus
M s = D − 1 2 KD1 2
= D − 1 2 DMD − 1 2
= D1 2 MD − 1 2
is a normalized affinity matrix (a normalized kernel), where use has been made of
the identity K = DM.
Since matrices M and Ms are related by a similarity transformation, they share
the same eigenvalues. To prove this, let ψ denote the eigenvalues of M and Φ
denote those of Ms. Observing that Ms = D−1/2KD−1/2, it can be inferred that
Ms is a symmetric matrix, which allows the use of special techniques for calculat-
ing the associated singular values. Griffith [31] gives an interesting interpretation
of these matrices in the context of spatial analysis and filtering.
Collecting all eigenvalues yields
MΨ = λ j Ψ
MsΦ = λ jΦ
with Ψ = ψ o ψ 1 ψ m − 1 and Φ = ϕo ϕ1 ϕm − 1 .
From these relationships, the eigenvalues λj of Ms satisfy
M s − λ j I ϕ j = 0 = D1 2 MD − 1 2 − λ j I ϕ j = M − λ j I D − 1 2 ϕ j = 0
ψj
Observational
High-dimensional data
(input) space t1 tk tN
Reconstruction
Clusters
Linear of data Inverse
(nonlinear)
mapping
mapping
embedding . . . .
Ψ εd (xm) λ1ψ1 (m) λ2ψ2 (m) λdψd (m)
..
.
Ensemble
• The eigenvectors associated with the largest singular values correspond to slow
modes governing the long-time system evolution, i.e. ψ j, j = 1, …, d;
λ1 ≥ λ2 ≥ … ≥ λ d λd + 1 ≥ … λm
Slow motion
M i=M×M×…
i times
1) Like the case of the spectral representation in DMs, the first eigenvector has a
unit value and can be associated with average behavior;
2) The second and third eigenvector capture dominant system behavior.
Hydro generator
To Areas 2 and 3 Thermal/nuclear
Bus
Figure 5.5 Single-line diagram of the study region showing the 230-/400-kV transmission
system and its interconnection with other regional systems.
166 5 Wide-Area Voltage Monitoring in High-Renewable Integrated Power Systems
153 19 495
33 19 063
25 18 859
32 18 859
108 18 843
73 16 809
94 14 372
65 13 982
126 13 549
167 13 409
136 13 074
61 10 405
44 9157
5.5 Case Study 167
Table 5.3 The largest attenuation values αij computed using (5.4). Base operating case.
Table 5.3, in turn, shows the electrical distances αij computed using the sensitiv-
ity relation:
ΔV i ∂V i ∂V j
αij = =
ΔV j ∂Q j ∂Q j
108
2
32, 25, 33
153
0
Figure 5.6 Sensitivity coefficients associated with bus 153 (unnormalized values).
168 5 Wide-Area Voltage Monitoring in High-Renewable Integrated Power Systems
Table 5.4 Candidate pilot buses selected using the information from electrical distances.
sorted, electrical distances can be used to divide the system into several voltage
control zones. Candidate pilot buses selected from this analysis include buses
153, 73, 136, 167, 61, 108, and 156 (refer to Table 5.4).
Based on the results of this exploratory analysis, seven major voltage control areas
or electrical zones were initially identified for the regional voltage control scheme:
Figure 5.7 shows the spatial distributions and principal transmission resources
associated with these voltage control zones. In practice, clusters can be combined if
they share certain properties or lack of generation sources.
Further, Table 5.5 lists the bus pilot nodes selected for analysis taking into
account generation resources, suggesting that bigger areas could be obtained by
relaxing the initial design criteria.
5.5 Case Study 169
105
Zone 6
156 25
Zone 1
153 32
108
87 113 33
Zone 5
7 29 61 151
36
83 173 21
34
162
Zone 4 65 22 24
94 39 Zone 3
20 88 4 160
1 8
18
167 125 149 136 53
26 70
170 165 SVC-2 100
17 13
81 69
84 148 44 73 66
SVC-1 150 2 Zone 2
10 40 122
101 90
117 15 132
35 67
57
124
Figure 5.7 Estimated spatial distribution of voltage control areas based on sensitivity
analysis.
1 153
2 73
3 136
4 167
5 61
Generator reactive power output 105 generator reactive power output signals
signals (XQg)
SVC reactive power output signals Three reactive power output signals (SVC-1,
(XSVC) SVC-2, and SVC-3)
bus voltage magnitudes and phases, and generator and SVC reactive output power.
Table 5.6 summarizes the main characteristics of signals selected for analysis.
For each of these scenarios, a set of six contingencies of interest was examined
for system characterization (see Table 5.7).
Based on these results, the snapshot (measurement) data is defined as:
T
X V = V 1 V 2 … V 173 T
= X busvolt X buvoltwfs
X θ = θ1 θ2 … θ173 T 5 17
T
X Q = Q1 Q2 … Q105 T
= X Qgen X Qsvc
Contingency
scenario Description Remarks
4
Q [MVAr]
2 G. 25 U1
G. 25 U2
0 G. 25 U3
–2
6 7 8 9 10 11 12 13 14 15
1
Q [MVAr]
G. 32 U1
0 G. 32 U2
–1 G. 32 U3
–2
6 7 8 9 10 11 12 13 14 15
34
Q [MVAr]
32 G. 35 U1
G. 35 U2
30 G. 35 U3
28 G. 35 U4
G. 35 U5
6 7 8 9 10 11 12 13 G. 35 U6 15
Q [MVAr]
39
38
G. 156
37
6 7 8 9 10 11 12 13 14 15
Time [sec]
Figure 5.8 Reactive power deviations following a double-line outage for contingency
scenario CE06.
172 5 Wide-Area Voltage Monitoring in High-Renewable Integrated Power Systems
Bus 2
Q [MVAr]
–13.5
–14
–14.5
–15
6 7 8 9 10 Bus 73 11 12 13 14 15
Q [MVAr]
–8
–9
6 7 8 9 10 Bus 81 11 12 13 14 15
Q [MVAr]
15
14
6 7 8 9 10 Bus 100 11 12 13 14 15
Q [MVAr]
0.5
0
–0.5
6 7 8 9 10 11 12 13 14 15
Time [sec]
Figure 5.9 Reactive power output for generators in zone 2 for the contingency
scenario CE06.
10
–10
–20
SVC 1
–30 SVC 2
SVC 3
–40
–50
–60
–70
6 7 8 9 10 11 12 13 14 15
Time [sec]
Figure 5.10 SVC reactive output power for contingency scenario CE06.
1.05
1.04
1.03
1.02
Voltage [pu]
1.01
0.99
0.98
0.97
6 7 8 9 10 11 12 13 14 15
Time [sec]
Cluster Buses
2 40, 67
3) Construct the mapping from the original space to the DM space (refer to
Figure 5.4):
Ψdε = ψ 1 ψ 2… ψd ,
105
156
25
153 32
108
87 113 33
7 29 61 151
36
83 173 21
34
162
22 24
65 94 39
20 88 4 160
1 8
18
167 125 149 136 53
26 70
170 165 SVC-2 100
17 SVC-1 13
81 69
84 148 44 73 66
150 2
10 40
101 90 122
15 132
117
35 67
57 SVC-3 124
Wind farms
Figure 5.12 Approximate boundaries and geographical locations for clusters extracted
using DMs. Modified case with an SVC at bus 40 and a large cluster of WFs connected at
bus 67.
Contingency
scenario Bus (magnitude of mode shape)
CE01 153 (0.259), 32 (0.220), 31 (0.175), 24 (0.156), 152 (0.156), 108 (0.146),
156 (0.141), 113 (0.130), 33 (0.128), 25 (0.111), 36 (0.102), 106 (0.126),
95 (0.083), 87 (0.086)
CE02 32 (0.229), 153 (0.229), 156 (0.195), 111 (0.172), 31 (0.166), 108 (0.158),
24 (0.137), 110 (0.153), 157 (0.152), 152 (0.139)
CE03 32 (0.227), 153 (0.226), 111 (0.168), 31 (0.162), 107 (0.154), 24 (0.139),
110 (0.152), 156 (0.149), 25 (0.132), 109 (0.135)
CE04 32 (0.212), 153 (0.212), 111 (0.158), 155 (0.173), 31 (0.161), 107 (0.147),
157 (0.141), 110 (0.144), 152 (0.135), 24 (0.135)
CE05 159 (0.132), 65 (0.123), 54 (0.108), 7 (0.106), 68 (0.103), 25 (0.093),
33 (0.094), 8 (0.088), 99 (0.094), 160 (0.096)
176 5 Wide-Area Voltage Monitoring in High-Renewable Integrated Power Systems
105
Sensor placement (DMD)
156 25
108 153 32 Pilot node
87 113 33
7 29 61 151
36
83 173 21
34
162 24
65 20 94 22
88 4 39
1 8 160
18
167 126 149 136 70 53
26 100
170 165 SVC-2 13 SLC
17
81 69
84 148 44 73 66
SVC-1 150 2
10 40
101 122
LCP 90 132
117 15
35 67
57
124
Figure 5.13 The candidate pilot node locations. Empty dashed circles show pilot node
locations from Table 5.5. Filled gray squares show the candidate node locations obtained
from DM information in Table 5.9.
Similar results are obtained using the MCL algorithm in Figure 5.15. In this case,
the leading eigenvector is obtained solving the standard eigenvector equation:
M MCL r − λMCL2 I ϕMCL2 = 0
where the power coefficient r is set to r = 4, and ϕMCL2 denotes the second eigen-
vector in the spectral decomposition.
(a)
×10–3
4
0
Real part of mode
–2
–4 Bus 156
Bus 33
–6
Bus 108
–8 Bus 32 Bus 153
Bus 25
–10
Bus
(b)
×10–3
2 Bus 69 Bus 100 Bus 160
Bus 2 Bus 53
1 Buses
20,73
–1
Magnitude
Bus 167
Bus 113
–2
–3
Bus 33
–4
Bus 156
Bus 108 Bus 153
–5 Bus 25 Bus 32
–6
Bus
Figure 5.14 Leading eigenvector of the Markov transition matrix: (a) contingency scenario
CE04 and (b) contingency scenario CE05 (unnormalized values).
0.1
–0.1
Real part of mode
Bus 108
–0.2 Bus 156
–0.3
–0.4
–0.5
Bus 25 Bus 33 Bus 153
–0.6
Bus
Figure 5.15 Dominant eigenvector extracted using the MCL for contingency
scenario CE04.
X = XV X Qgen X Qsvc 5 18
Table 5.10 shows the five largest diffusion coefficients obtained using this
approach. Comparisons with reactive power sensitivities show that DMs can pro-
vide an accurate characterization of the strength of interactions between bus volt-
age magnitudes and reactive power reserves and also have a strong physical
interpretation. It is noted that the adopted approach identifies both SVCs and gen-
erators showing the largest strength of interaction with nearby buses.
The practical application of these concepts to complex power systems with
increased wind penetration is deferred to Chapter 8.
O1 XTY
X= 5 19
YTX O2
R = UΣV T 5 21
Here, U and V are orthonormal matrices called saliences containing the left
and right singular vectors of R, respectively, and Σ is a diagonal matrix of the
nonzero singular values. The latent variables, Lx and Ly, which are linear combi-
nations of the original values, are obtained by projecting the data matrices onto
their respective saliences as:
Lx = XV
and
Ly = YU
When coupled with real-time information, these techniques have also the poten-
tial to correlate reactive power reserves with bus voltage deviation and provide
operational warnings.
–1
Real part of mode
–2
Bus 148
–3 units 1–2
–4
Bus 73
–5 units 1,2,4,5
Bus 2 Bus 100
units 1–5 Bus 81 units 1–6
–6 units 1–6
–7
0 10 20 30 40 50 60 70 80 90 100
Generator
Figure 5.16 shows partial least squares regression results for data matrices let
X = XV and Y = X Qwfs X Qgen X Qsvc . Results correlate well with previous results
in Figures 5.14 and 5.15.
This example demonstrates that correlation analyzes may identify key relation-
ships between observed bus voltage deviations and output reactive power from
wind farms and synchronous generators, such as sensitivity relationships, or levels
of strength associated with reactive power control areas. It is found that both sim-
ple correlation analysis and partial least squares regression perform well for the
simulated records.
5.6 Summary
Over the last few decades, various forms of voltage monitoring and reactive power
management systems have been developed. These architectures offer the possibil-
ity to calculate in near-real-time several aspects of voltage control and VAR man-
agement such as distances to voltage collapse and network reactive control,
distances to instability, and sensitivity calculations. The (geographical) dispersion
of dynamic recorders and the changing network structure and operating condi-
tions, however, create wide-area voltage monitoring issues that must be addressed
using advanced data-based analysis approaches and correlation techniques.
In this chapter, a new use of spectral analysis tools for dimensionality reduction
and clustering of high-dimensional datasets, based on the normalized graph Lapla-
cian has been introduced. The proposed technique combines the inherent abilities
of graph-theoretical techniques with spectral clustering and visualization methods
to identify loosely interconnected voltage control areas and reconstruct system
behavior using selected measurements. Attention has been focused on three main
aspects, namely the identification of critical system zones showing a coherent
behavior and the associated reactive power sources, the use of spectral analysis
to identify the critical buses, and the computation of reduced-order models.
Issues such as reserve capacity control and the efficient utilization and coordi-
nation of reactive power sources need to be addressed to achieve fast voltage mon-
itoring and control and keep the reactive power output margin against system
contingencies.
References
1. Coordinated Voltage Control in Transmission Networks. (2007). CIGRE Task
Force C4.602 (February 2007).
182 5 Wide-Area Voltage Monitoring in High-Renewable Integrated Power Systems
2. Corsi, S., Possi, M., Sabelli, C., and Serrani, A. (2004). The coordinated automatic
voltage control of the Italian transmission grid - Part I: reasons of the choice and
overview of the consolidated hierarchical system. IEEE Transactions on Power
Systems 19 (4): 1723–1732.
3. Cañizares, C.A., Cavallo, C., Pozzi, M., and Corsi, S. (2005). Comparing secondary
voltage regulation and shunt compensation for improving voltage stability and transfer
capability in the Italian power system. Electric Power Systems Research 73: 67–76.
4. Lagonotte, P., Sabonnadiere, J.C., Léost, J.Y., and Paul, J.P. (1989). Structural
analysis of the electrical system: application to secondary voltage control in France.
IEEE Transactions on Power Systems 4 (2): 479–486.
5. Taylor, C.W. (1994). Power System Voltage Stability. McGraw-Hill.
6. Taylor, C.W. (1993). Survey of effective and practical solutions for longer-term
voltage stability. Electrical Power and Energy Systems 15 (4): 217–220.
7. Taylor, C.W. (2000). The Future in on-line security assessment and wide-area
stability control. Proceedings of the 2000 IEEE Power Engineering Society Winter
Meeting.
8. AEMO. (2013). Australian energy market operator, wind turbines plant capabilities
report. Wind Integration Studies. https://www.aemo.com.au/-/media/Files/PDF/
Wind_Turbine_Plant_Capabilities_Report.pdf/.
9. Miller, N.W., Guru, D., and Clark, K. (2008). Wind generation applications for the
cement industry. Proceedings of the IEEE Cement Industry Technical Conference
Record.
10. Miller, N., MacDowell, J., Chmiel, G. et al. (2012). Coordinated voltage control for
multiple wind plants in eastern wyoming: analysis and field experience.
11. Ullah, N.R. and Bhattacharya, K. (2008). Wind farms as reactive power ancillary
service providers—technical and economic issues. IEEE Transactions on Energy
Conversion 24 (3): 661–672.
12. Taranto, G.N., Martins, N., Falcao, D.M. et al. (2000). Benefits of applying
secondary voltage control schemes to the Brazilian system. Proceedings of the 2000
Power Engineering Society Summer Meeting (July 2000).
13. Sancha, J.L., Fernandez, J.L., Cortez, A., and Abarca, J.T. (1996). Secondary voltage
control: analysis, solutions and simulation results for the Spanish transmission
system. IEEE Transactions on Power Systems 11 (2): 630–638.
14. Paul, J.P., Leost, J.Y., and Tesseron, J.M. (1987). Survey of the secondary voltage
control in France: present realization and investigations. IEEE Transactions on
Power Systems PWRS-2 (2): 505–511.
15. Kundur, P. (1994). Power System Stability and Control, The EPRI Power System
Engineering Series. New York, NY: McGraw-Hill.
16. Carpentier, J.L. (1987). “CRIC”, a new active-reactive decoupling process in load
flows, optimal power flows and system control. IFAC Proceedings 20 (6): 59–64.
References 183
17. Zhong, J., Nobile, E., and Bose, A. (2004). Localized reactive power markets using
the concept of voltage control areas. IEEE Transactions on Power Systems 19 (3):
1555–1561.
18. Cotilla-Sanchez, E., Hines, P., and Barrows, C. (2013). Multi-attribute partitioning
of power networks based on electrical distance. IEEE Transactions on Power
Systems 28 (4): 4979–4987.
19. Tyuryukanov, I., Popov, M., van der Meijden, M.A.M.M., and Terzija, V. (2018).
Discovering clusters in power networks from orthogonal structure of spectral
embedding. IEEE Transactions on Power Systems 33 (6): 6441–6451.
20. Jiang, T., Bai, L., Ji, L., and Li, F. (2007). Spectral clustering-based partitioning of
volt/VAR control areas in bulk power systems. IET Generation, Transmission &
Distribution 11 (5): 1126–1133.
21. Saugata, S., Biswas, T., and Srivastava, A.K. (2014). Performance analysis of a new
synchrophasor based real time voltage stability monitoring (RT-VSM) tool.
Proceedings of the 2014 North American Power Symposium (NAPS).
22. Sun, H., Guo, Q., Zhang, B. et al. (May 2013). An adaptive zone division-based
automatic voltage control system with applications in China. IEEE Transactions on
Power Systems 28 (2): 1816–1828.
23. Morison, K., Wang, X., Moshref, A., and Edris, A. (2008). Identification of voltage
control areas and reactive power reserve; an advancement in on-line voltage
security assessment. Proceedings of the 2008 IEEE Power and Energy Society
General Meeting – Conversion and Delivery of Electrical Energy in the
21st Century.
24. Roman-Messina, A. (2015). Wide-area Monitoring of Interconnected Power Systems,
IET, Power and Energy Series 77, Stevenage. UK: IET, The Institute of Engineering
and Technology.
25. Li, H., Li, F., Xu, Y., Y. et al. (2010). Adaptive voltage control with
distributed energy resources: algorithm, theoretical analysis, simulation,
and field test verification. IEEE Transactions on Power Systems 25 (3):
1638–1647.
26. Messina, A.R. and Vittal, V. (2007). Extraction of dynamic patterns from wide-area
measurements using empirical orthogonal functions. IEEE Transactions on Power
Systems 2 (2): 682–692.
27. Newman, M. (2010). Networks, 2e. Oxford, UK: Oxford University Press.
28. Dray, S., Legendre, P., and Peres-Neto, P.R. (2006). Spatial modelling: a
comprehensive framework for principal coordinate analysis of neighbour matrices
(PCNM). Ecological Modelling 196: 483–493.
29. Arvizu, C.M. and Messina, A.R. (2016). Dimensionality reduction in transient
simulations: a diffusion maps approach. IEEE Transactions on Power Delivery
31 (5): 2379–2389.
184 5 Wide-Area Voltage Monitoring in High-Renewable Integrated Power Systems
30. David, G. and Averbuch, A. (2012). Hierarchical data organization, clustering and
denoising via localized diffusion folders. Applied and Computational Harmonic
Analysis 33 (1): 1–23.
31. Griffith, D. (2000). Eigenfunction properties and approximations of selected
incidence matrices employed in spatial analyzes. Linear Algebra and Its
Applications 321 (1-3): 95–112.
32. Moghadas, S.M. and Jaberi-Douraki, M. (2019). Mathematical Modelling. Hoboken,
NJ: Wiley.
33. Román-Messina, A. (2020). Data Fusion and Data Mining for Power System
Monitoring. Boca Raton, FL: CRC Press.
34. S. Van Dongen, Graph Clustering by Flow Simulation. PhD thesis, University of
Utrecht, May 2000.
35. Enright, A.J., Van Dongen, S., and Ouzounis, C.A. (2002). An efficient algorithm
for large-scale detection of protein families. Nucleic Acids Research 30 (7):
1575–1584.
36. Malmstrom, R.D., Lee, C.T., Van Wart, A.T., and Amro, R.E. (2014). Application of
molecular-dynamics based Markov state models to functional proteins. Journal of
Chemical Theory and Computation 10: 2648–2657.
37. Cazade, P.A., Zheng, W., Prada-Garcia, D. et al. (2015). A comparative analysis of
clustering algorithms: O2 migration in truncated hemoglobin I from transition
networks. The Journal of Chemical Physics 142: 025103-1-15.
38. Bezdek, J.C. (1981). Pattern Recognition with Fuzzy Objective Function Algorithms.
New York, NY: Plenum Press.
39. Bezdek, J.C., Ehrlich, R., and Full, W. (1984). FCM: The Fuzzy c-means clustering
algorithm. Computers and Geosciences 10 (2–3): 191–203.
40. Zadeh, L.A. (1971). Similarity relations and fuzzy orderings. Information Sciences
3: 177–200.
41. Barocio, E., Pal, B.C., Thornhill, N.F., and Roman-Messina, A. (2015). A dynamic
mode decomposition framework for global power system oscillation analysis. IEEE
Transactions on Power Systems 30 (6): 2902–2912.
42. Lindenbaum, O., Yeredor, A., Salhov, M., and Averbuch, A. (2020). Multi-view
diffusion maps. Information Fusion 57: 127–149.
43. Krishnan, A., Williams, L.J., b., McIntosh, A.R., and Abdi, H. (2011). Partial least
squares (PLS) methods for neuroimaging: a tutorial and review. Neuroimage
5: 455–475.
44. Abdi, H. and Williams, L.J. (2013). Partial least squares methods: partial least
squares correlation and partial least square regression. In: Computational
Toxicology, Methods in Molecular Biology, 930, vol. II (eds. B. Reisfeld and A.N.
Mayeno), 543–579. New York, NY: Human Press-Springer.
185
6.1 Introduction
60 Hz Area 2
Intertial
response
Nadir
Local COI
dynamics
δCOI1, MCOI1
COG [∆PL – ∑ PCOI
tie
– ∆PiMMG]
Area n
tie
PCOI δCOG , MCOG MCOG j
i,COG
j,COG Mi (∆Pmech –∆Pelec)
tie ΔfCOIi = ΔfCOG
PCOI M 2πD
n,COG 1 + COG Δf
Mi (∆Pmech –∆Pelec) COG
Area i
δCOIk, MCOIk
[∆PL – ∑ PCOI
tie
,COG – ∆Pi
MMG]
i
Mi j
Mj (∆Pj –Pijtie – ηj∆PiMMG)
Δfj = ΔfCOI
1 + Mi 2πD i
Δf conv. ESS+
Mj (∆Pj – Pijtie– ηj∆PiMMG) COIi δCOIi, MCOIi + M MMGs
Figure 6.1 Conceptual overview of the adopted control scheme showing the grid-parallel
configuration of interconnected MGs.
6.2 Frequency Dynamics Enhancement 187
L1
L2
F1cosδ δ
h1 F1
h2
COG
F2
Furthermore, the fictitious reactance Xtie can be calculated using (2.28), i.e.
min S
s = 1 ps ξs− − ξs+
X tie
COI , COG
i
st
n n
1 V COI i V COG
sin δCOI i − δCOG + ξs− − ξs+ = ΔPMMG
ω
i = 1 COG X tie
COI i ,COG i=1
i,s
−
ξs , ξs+ ≥ 0
0 < X tie
COI i ,COG ≤1
65
188 6 Advanced Control Synthesis
According to (6.3) and (6.5), the Eq. (6.1) for the equivalent system of Figure 6.1
can be reorganized as
ΔP1 ΔP2 ΔP3 ΔPn
X1 + X2 + X3 + … + Xn = 0 66
M1 M2 M3 Mn
The equilibrium condition (6.6), as the basic principle of the inertia-based
control scheme for frequency support, relies on the following simple yet effective
premise [11]: “Any variation in the equivalent reactance, X in Figure 6.1, following
a disturbance, will be compensated by a corresponding change in ΔPi/Mi so that
the equilibrium condition is satisfied.” Area i in Figure 6.1 schematically describes
the variation of F, through manipulation of M, in the equivalent system. Referring
to Figure 6.1 and taking the contribution of multi-MGs (MMGs) into account, the
swing Eq. (6.3), relative to the COG, can be expressed in the form of (6.7) [11].
1
f COI i = ΔPi − 2πDi Δ f COI i − Ptie
COI i ,COG − ΔP i
MMG
67
2πM COI i
M i = M Conv
COI i + M
ESS
+ M MMGs 68
ESS
where M Conv
COI i , M , and MMMGs represent, respectively, the conventional synchro-
nous inertia and the inertia provided by the ESS and MMGs.
Key parameters of interest to be controlled via manipulation of (6.8) include the
frequency nadir, the rate of change of frequency (RoCoF), and the frequency devi-
ation during a given time interval of interest. Using this framework, the ESS output
is manipulated to reduce the RoCoF and the frequency nadir through the emulated
virtual inertia. Disconnecting the MMGs from the network (islanding) in a
time horizon greater than the response time of ESS, on the other hand, increases
the frequency evolution to satisfy the rolling window criterion. The effects of the
MMGs islanding strategy on the frequency response are examined based on (6.8)
for two different operating conditions of the MGs, i.e. power import from the grid
and power export to the grid. These cases are discussed separately as follows [11]:
6.2 Frequency Dynamics Enhancement 189
1) Power import operating mode: In the inertial response period where Tm < Te,
the MG islanding causes the torque Te to decrease, which in turn decreases the
RoCoF and nadir. On the other hand, for the time interval beyond the nadir,
where Tm > Te, islanding leads to a greater acceleration torque and thus a faster
frequency recovery.
2) Power export operating mode: In this case, the MG islanding, reinterpreted as
loss of inertia according to (6.8), renders frequency dynamics faster. The iner-
tia-based control strategy imposes the islanding of the MMGs beyond the fre-
quency arrest period.
Δ f COI i = Δ f COG 6 18
M COI i ΔPmech − ΔPelec
or equivalently,
M ΔPmech − ΔPelec
Δ f COI i = COI i Δ f COG 6 19
M COG 2πD
1+ Δ f COG
M COI i ΔPmech − ΔPelec
6.2 Frequency Dynamics Enhancement 191
As is apparent in (6.19), the MMGs capacity ΔPMMG affects the COI dynamics
through its effect on the power imbalance and the associated sensitivity factor.
Let now the swing Eq. (6.3) be rewritten as
1
fj= ΔP j − Ptie
ij − 2πD j Δ f j − η j ΔP i
MMG
6 20
2πM j
η j = 1; j i; i 6 21
j
and ηj is defined as the ratio of the total MMGs capacity at bus j to the total MMGs
capacity in the associated area i. Following the same procedure as that in (6.19), it
can be shown that
Mj ΔP j − Ptie
ij − η j ΔP i
MMG
Δf j = Δ f COI i 6 22
M COI i 2πD
1+ Δ f COI i
M j ΔP j − Ptie − η ΔPMMG
ij j i
Start
nadir –
fj >f?
Yes
or
nadir –
dfj df
> ?
dt dt ESS dispatching
NO
– Yes
Δfj 15–sec > Δf ?
NO MMG islanding
End
ΔPmech – 1 1 Δfj
+
– (MConv. + ηj MMMG)s 2π
and safe reconnection of the two grids [15, 16]. The early static switch closing
causes low-frequency beat voltage across the switch. This beat voltage can cause
a significant beat current, which leads to adverse effects such as possible resonance
with mechanical structure and additional power loss. Figure 6.5 shows the beat
voltage and the current flow from the switch for the early closing of the switch.
Figure 6.5 Beat voltage for early closing of the switch; upper plot: beat voltage across the
switch, second plot: current of phase A, third plot: current of phase B, fourth plot: current of
phase C.
194 6 Advanced Control Synthesis
2
V [pu]
–2
10 16 22 28 34 40 46 52
Time [s]
Monitoring the voltage across the static switch reveals that the MMGs and the
utility should be synchronized when the phase difference between voltages of
the two grids is approximately zero. Figure 6.6 shows voltages on either side of the
static switch. From these results, the interval time between 33 and 42 seconds is an
appropriate synchronization time to close the switch [17].
1) Two-area system:
The single-line diagram of the system is shown in Figure 6.7. The disturbance
considered is the trip of 1400 MW generation, i.e. ΔPL = 14 pu, in Area 2. The
Area 1 Area 2
1 10 20 3 101 13 120 11
Gen3
Gen1
110
Gen2 Gen4
2 4 12
14
reduced equivalent model of the system is derived based on (6.5) to gain insight
into the nature of the frequency behavior. For this purpose, 10 random scenar-
ios of equal probability, i.e. Ps = 0.1 in (6.5), are generated. Minimization of
(6.5) reveals that the reduced equivalent model can be characterized by
X tie
COI 1 ,COG = 0 2134; X tie
COI 2 ,COG = 0 3172 6 23
1 67
Δf1 = Δ f COG 6 24
1 + 0 97Δ f COI 1
1 16
Δ f 12 = Δ f COG 6 25
1 + 0 08Δ f COI 2
Equations (6.24) and (6.25) suggest that the frequency nadir, for generator buses
in Area 2, limits the ability of the host grid to accommodate high levels of MGs
generation. A related problem of interest is that of determining the number of ESSs
as well as the generation capacity of the MMGs that can be integrated into the grid
to efficiently support system frequency. From (6.2) and for the base system, one
can write
df 1 14pu
= 6 26
dt M1
and, for the system with high penetration of MGs generation and installed ESSs,
we have
df 2 14pu
= 6 27
dt M 2 + M ESS
where M2 = M1 + MMMG. Dividing (6.27) by (6.26) gives
df 2 M1
= 6 28
df 1 M 2 + M ESS
Multiplying both sides of (6.28) in the term Δf1, on the other hand, gives
df 2 M1
× Δf1 = × Δf1 6 29
df 1 M 2 + M ESS
Equation (6.29) can be rewritten in the compact form as
M1
Δf2 = Δf1 6 30
M 2 + M ESS
196 6 Advanced Control Synthesis
PGenset 5
M MMG = 1 89 MMG
= 1 89 = 0 43 s 6 32
P 22
where PGenset and PMMG define the generated power of the synchronous-based DG
and the overall MG generation, respectively (details on how to calculate (6.32) are
given in Chapter 2), and the inertia constant in (6.32) is on a 15-kVA base. Accord-
ingly, MMMG=1.4 would be realized in the base of the system, i.e. 100 MW, by
penetration of
1 4 × 100
PMMG = × 0 022 = 477 52 MW 6 33
0 43 × 0 015
which allows a 17.5% penetration level of MMGs. On the other hand, providing
0.9% of the base system inertia by ESSs results in [11]
0 5J VI ω2 − VAhESS VAhESS
M ESS = 0 009 × M 2 = 0 009 × 6 5 =
VArated 900
VAhESS = 52 65
6 34
It then follows that the number of ESS units, which could provide such an
energy, is
0
Δf2 [Hz]
–0.1
0 2 4 6 8 10 12 14 16 18 20
Time [s]
1 015
Δ f 12 = Δ f COG 6 36
1 + 0 01Δ f COI 2
Comparing (6.25) with (6.36) suggests that the inertia-based control scheme can
effectively reduce the frequency nadir. Figure 6.8 exhibits the appropriateness of
the adopted control strategy, assuming that, in the online control system center,
the frequency event is detected after receiving five samples with increasing or
decreasing rate [17]; in our simulations, the ESSs are triggered 0.1 second follow-
ing the inception of the fault.
2) NYNE system:
NYNE test system of Figure 6.9 is used to further illustrate the efficiency of the
inertia-based control scheme. Five different contingency scenarios, including
the trip of major generating units and load shedding at selected buses, are con-
sidered for the studies. Results in Table 6.1, comparing the capability of the host
grid to accommodate MGs with and without the advanced control scheme,
show the effectiveness of the inertia-based control scheme.
Table 6.1 reveals that the penetration level of MMGs into the system is limited by
the frequency nadir at bus 60, in Area 1, for contingency 3. This contingency
causes the COG and COI to experience the frequency nadirs of 49.78 and
49.89 Hz, respectively. Using (6.19) and (6.22), the frequency dynamics of bus
60 can be expressed in terms of the COG frequency behavior by the sensitivity
relation
1 346
f nadir = f 6 37
60
1 + 0 363 f COI 1 COG
then
1 346
f nadir
60 = 49 78 = 49 19 6 38
1 + 0 363 49 89 50
198 6 Advanced Control Synthesis
14 Area 3 Area 1
66 8
41 40 48 47 60
1
25 26 29
53
28
10 1 2 61
27
62 9
30
31 3
11
63 18 17
16 24
46 38 32
33 15 22
Area 5 34
9
4
21
19 58
49 14
35 6
42 36 5 20
8 6 12 56 23
67 64
12
57 4
15 45
51 5
59
50 44 37 11 13 7
7 54
2
65 10
52 43 13
39
55
68
3
Area 4 16 Area 2
Figure 6.9 Single-line diagram of the 68-bus system showing coherent areas and their
interconnections.
Table 6.1 Comparison of maximum penetration level for NYNE test system with and
without the controller.
Inspection of (6.38) shows that when the COG frequency nadir reaches 49.78 Hz,
the frequency at bus 60 drops to 49.19 Hz, thus limiting the penetration of MG gen-
erations into the system. Installing 11 MW of ESSs at bus 60 increases the first
swing frequency amplitude to 49.42 Hz, and the penetration level of the system
increases by 17%. Table 6.2 reports the required ESSs to increase the penetration
level of MMGs. It should be noted that as generators G14–G16 represent the aggre-
gated behavior of areas 3–5 connected to the New York power system, the
6.2 Frequency Dynamics Enhancement 199
specification of the ESSs capacity for such equivalent areas is far from reality.
While the reported results in columns 2 and 3 of Table 6.2 are expressed in per-
centage based on each area generation capacity, the last column is expressed in
percentage based on the total system load.
It should be emphasized that using this approach, each area in the system is
characterized by two equations of the form (6.30) and (6.31). Given a system with
n areas, solving the set of 2 × n equations allows to determine the number of ESSs
and MMGs ratings in the overall system.
Figure 6.10 compares the penetration levels as well as the frequency nadirs for
the system with and without the inertia-based controller. For completeness, the
30
25
Penetration [%]
20
15
10
5
0
1 2 3 4 5 6 7 8
49.4
49.3
49.2
fmin[Hz]
49.1
49
48.9
48.8
1 2 3 4 5 6 7 8
Scenario
1
With ESS
Without ESS
f [pu]
0.96
0.92
0 1 2 3 4 5 6 7 8 9 10
Time [s]
Δδi t = δi t − δ0i 6 39
and
ΔV i t = V i t − V 0i 6 40
where δ0i and V0i are the initial values of rotor angle and terminal voltage related to
ith generator, respectively. To establish the effective criteria for triggering of the
inertia-based control scheme, parameters normalization is suggested. For this pur-
pose, consider the maximum variable variations as:
and
ΔV max t = max ΔV i t 6 42
Accordingly, one could rewrite (6.39) and (6.40) in the normalized form of
Δδi t
Δδi t = 6 43
Δδmax t
and
ΔV i t
ΔV i t = 6 44
ΔV max t
Economic reasons in addition to environmental constraints cause transmission
lines to operate close to their limits. Therefore, the desired operation of a power
system after being subjected to a disturbance could be theoretically achieved when
the system returns to the planned operating point, characterized by voltage profile,
nominal frequency, and transmitted power. Considering the power flow between
buses i and j, following a system perturbation, in the form of one could define the
following conditions to guarantee the desired performance:
V 0i + ΔV i t V 0j + ΔV j t
Pij = sin δ0i + Δδi t − δ0j + Δδ j t
X ij
6 45
202 6 Advanced Control Synthesis
Ideally, all the connected generators satisfy the aforementioned conditions, and
hence, they are characterized by (1, 1) in the normalized (Δδ − ΔV) plane. How-
ever, in a real power grid, the system returns to a new operating point different
from the initial one, and hence, the voltage and the difference between the rotor
angle deviations of the generators differ from each other. Therefore, control
actions should be taken to force the terminal voltage deviations and the difference
between the rotor angle deviations into the desired value, i.e. zero. In other words,
for the stable, secure, and reliable system operation, all the generators’ operating
points must be located in the minimum distance from (1, 0) in the normalized
(Δδ − ΔV ) plane. This could be mathematically represented by the minimization
problem of
min Δδi + ΔV i − 1, 0 6 46
which could be rewritten as
2 2
min Δδi + ΔV −1 6 47
=0 6 48
2 2
2 Δδi + ΔV i
Hence,
d ΔV i
2Δδi + 2ΔV i =0 6 49
d Δδi
from which it follows that
Δδi = ΔV i = 0 707 6 50
As stated, the inertia-based control scheme tries to conduct all the committed
generators to the prefault condition, i.e. (1,1) in the normalized plane. Therefore,
one could define the secure operation region in the plane as:
Δδi ≥ 0 707 6 51
and
ΔV i ≥ 0 707 6 52
6.3 Small Signal Stability Enhancement 203
While for the generating units located in the desired region CDEFC, dis-
patching of the ESSs and consequentially increasing the effective grid
1.41
G 0.707 A ΔV
45
E 0.707
F
D C B
204 6 Advanced Control Synthesis
inertia can decrease the parameter variations for the machines out of the
desired region, disconnecting the MMGs decreases effective grid inertia
which in turn increases the parameter variations.
Accordingly, the inertia-based control scheme changes Δδi, ΔVi and the maxi-
mum variable variations in (6.43) and (6.44) and hence, causes the generators to
move toward the desired region. Of note that the control scheme would be realized
in several steps in such a way that in each step, and for generators in the region of
interest, only 10% of the installed ESSs would be dispatched. On the other hand, only
10% of the synchronized MMGs are disconnected in each step of the algorithm.
0.35
0.3
0.25
Virtual Inertia
0.2
0.15
0.1
0.05
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Generator Number
(a)
–0.2
–0.4
ΔδN
–0.6
–0.8
–1
–1 –0.9 –0.8 –0.7 –0.6 –0.5
ΔVN [deg]
(b)
–0.975
–0.98
–0.985
ΔδN
–0.99
–0.995
–1
–1 –0.9 –0.8 –0.7
ΔVN [deg]
Figure 6.14 Trace of generating units in the Δδ – ΔV plane in response to the visualization
of flexible inertia after the outage of generator 2.
committed generators in the normalized Δδ – ΔV plane just 1 second after the out-
age of generator 2. It can be seen that except for six generators, others are out of the
desired region. As discussed earlier, dispatching the ESSs in the buses associated
with the generators in the desired region and disconnecting the MMGs in other
buses improve the system performance. Figure 6.14b suggests that the flexible-
inertia-based control scheme can return all the generating units to the desired
region.
Further, the effectiveness of the flexible inertia concept to improve the voltage
regulation as well as the small-signal stability performance is assessed considering
206 6 Advanced Control Synthesis
(a)
–0.2
–0.4
ΔδN
–0.6
–0.8
–1
–1 –0.9 –0.8 –0.7 –0.6 –0.5
ΔVN [deg]
(b)
–0.99
–0.992
–0.994
ΔδN
–0.996
–0.998
–1
the outage of generator 16. Figure 6.15 reveals that all the generating units are
located in the desired region in about 20 seconds after the inception of the fault.
The time-domain simulation result for the applied fault further demonstrates the
high capability of the proposed control scheme (Figure 6.16).
Regarding the time horizon of voltage instability (>2 seconds) and transient
instability (about 2 seconds), the sampling time interval to dispatch the ESSs
and disconnect the MMGs can be set to 2 seconds. However, the proposed control
strategy employs the first sample at a smaller time (less than 2 seconds) to improve
its reliability.
References 207
60
40
δ [deg]
20
0
0 5 10 15 20 25 30 35 40
1.1
V [p.u.]
0.9
0 5 10 15 20 25 30 35 40
Time [s]
Figure 6.16 Effects of flexible inertia on voltage regulation and small-signal stability.
6.4 Summary
This chapter deals with the designing of advanced control schemes to improve
power system stability. The discussed advanced control schemes rely on the inertia
manipulation in the system to mitigate undesired frequency, voltage, and rotor
angle dynamics. The developed control schemes depend on the stochastic equiv-
alent model of the power system to enable high penetration levels of MGs. The
approaches combine the adaptive ESS dispatch strategy with the MG controlled
islanding scheme to provide a stability support for the host grid.
References
1. Hatziargyriou, N. et al. (2020). Stability definitions and characterization of dynamic
behavior in systems with high penetration of power electronic interfaced
technologies. IEEE, Tech. Rep.
2. Hatziargyriou, N., Asano, H., Iravani, R., and Marnay, C. (2007). Microgrids. IEEE
Power and Energy Magazine 5 (4): 78–94.
3. Ulbig, A., Borsche, T.S., and Andersson, G. (2013). Impact of low rotational inertia
on power system stability and operation. arXiv preprint arXiv:1312.6435.
4. Hatziargyriou, N. et al. (2017). Contribution to bulk system control and stability by
distributed energy resources connected at distribution network, IEEE, PES-TR22.
5. Lasseter, R. et al. (2002). Integration of Distributed Energy Resources. The CERTS
Microgrid Concept. Berkeley, CA: Lawrence Berkeley National Labaratory (LBNL).
208 6 Advanced Control Synthesis
In the past decades, a variety of approaches have been introduced to assess small
and large systems performance, from linear, model-based techniques to data-
driven approaches. This includes efforts to extract spatial shapes, determine modal
properties, and assess the energy exchange between interconnected systems,
among other issues. A common approach is to simultaneously analyze the data
sequences using techniques such as the multichannel Prony method or Koopman
mode decomposition. Despite the importance of these methods, more general,
global multiscale methods that rely on the joint analysis of large data sets are
needed to cope with the ever-increasing utilization of distributed generators
(DGs) and the quantity and complexity of the recorded system behavior.
As the number and distribution of inverter-based generation resources grow, it
becomes increasingly challenging to extract modal properties from heterogeneous
and complex, measured responses following system perturbations. Knowledge of
the fundamental characteristics of global oscillation modes provides valuable
information about the stability of oscillatory phenomena and may help to identify
machines and wind and solar photovoltaic (PV) farms and their control systems
involved in the exchange of oscillating energy and the design or modification of
controllers.
This chapter identifies basic issues important in understanding the oscillatory
performance of wind and solar PV penetrated power systems. A fundamental
study of the characterization of the dynamic behavior of power systems with
increased renewable generation is presented. The study is motivated by the need
to further clarify the participation of wind and PV farms in interarea oscillations.
A second goal is to assess the performance of modern data-driven modal tools and
analysis techniques.
The use of modal characterization techniques is illustrated on two test systems: a
6-machine, 10-bus test system, and a 5449-bus, 635-generator test system. By using
a simple example, the effect of renewable generation on system dynamic
Measured data sets are often relatively large and may exhibit nonlinearities and
other artifacts. In this section, several data-driven approaches are examined to var-
ious aspects associated with the integration of RESs. First, a conventional mass–
spring system is used to introduce the nature of oscillatory phenomena. The
notation is also introduced.
where n is the number of modes, the v j s are the natural modes, and the terms
aj(t) = Aj sin(ωjt + θj) represent the time modulation of the natural modes, in
which Aj, ωj, and θj are the modal amplitude, frequency, and phase, associated
with the jth mode respectively.
Let now a t = a1 t a2 t … an t T ᑬn × 1 be the vector containing the
time evolution of the modes, aj(t) = Aj sin(ωjt + θj), j = 1, …, n, and
V = v 1 v 2 … vn ᑬnxn be the modal matrix of modal coordinates. It follows
that:
x t = Va t 72
By evaluating (7.2) for a given time interval, t1, t2, …, tN, the system response can
be written in terms of the modal matrix and the matrix of modal coordinates as
x1 t 1 x 1 t2 x1 t N
x2 t 1 x 2 t2 x2 t N
X t = x t1 x t2 … x tN = = VA t 73
xn t 1 xn t 2 xn t N
where, x k t j = x k t j xk t j … xk t j , j = 1, …, N, and
a1 t 1 a1 t 2 a1 t N
a2 t 1 a2 t 2 a2 t N
A=
an t 1 an t 2 an t N
is the matrix of time-dependent coefficients.
7.2 Modal Characterization Using Data-Driven Approaches 215
y t = y t + ε t , t = 0, …, N − 1 75
where N is the number of samples.
Assume further that the ringdown waveform y(t) can be written as a superposi-
tion of Q exponentials as [18, 22]
Q
yt = Ai eλi t cos 2π f i t + ϕi 76
i=1
Defining
Bi = Ai e jϕi
zi = eλi Δt = e σi + jωi Δt
Equation (7.6) can be rewritten in the discrete domain (t = tk = kΔt), in the form
n
yk = Bi zki 77
i=1
0 0 0 ao
1 0 0 a1
Cn = 0 1 0 a2 Cn × n
0 0 1 an − 1
where the ajs are unknown coefficients to be determined and det(zIn − Cn(a)) =
p(z), in which In is the n × n identity matrix. It is noted that the eigenvalues of
Cn coincide with the zeros of the Prony polynomial in (7.9).
It can be verified after some algebra that (7.9) satisfies a forward linear predictor
model (LPM) of the form [25, 26]
n
am y k − m = 0
m=0
Using the known data values, y(k), this equation can be rewritten in matrix
form as
yn y n−1 y1 y n+1
y n+1 yn y2 a1 y n+2
a2 y n+3
y n+2 y n+1 y3 = − , n+1≤k≤N
an
y N −1 y N −2 y N −n a
yN
Y y
7 10a
or
y n−1 y n−2 y0 yn
y n−1 a1
yn y1 y n+1
a2 y n+2
y n+1 yn y2 = − , n ≤ k ≤ N −1
an
y N −1 y N −2 y N −n−1 a
y N −1
Y y
7 10b
in which it is assumed that ao = 1.
Equations (7.10a) and (7.10b) can be rewritten in the form Ya = y, where Y is a
matrix of dimension (N − n) × n. This system of equations can be solved directly
for the ajs if N = 2n, as a = Y−1y [23].
218 7 Small-Signal and Transient Stability Assessment Using Data-Driven Approaches
an
y N −1 y N −3 y N −n−1 a
y N −1
Y b
7 11
which has a noniterative, least-squares solution of the form
−1
a = Y †b = − Y T Y Y Tb
y1 N − 1 y1 N − 2 y1 N − n − 1 a1
a2 y1 N − 1
= −
yM n − 1 yM n − 2 yM 0 yM n
an
yM n + 1
yM n yM n − 1 yM 1 am
Block M
yM N − 1 yM N − 2 … yM N − n − 1 yM N − 1
ym
Ym
7 13
where the a j s are the modified signals’ amplitudes, and each signal is defined by a
vector, y j = y j 0 y j 1 y j N − 1 T , j = 1, …, M.
Given this result, the following three-step procedure is used to compute the
Prony parameters
YM
220 7 Small-Signal and Transient Stability Assessment Using Data-Driven Approaches
with
yj n−1 yj n−2 yj 0
yj n yj n−1 yj 1
Yj= , j = 1, …, M
yj N −1 yj N −2 yj N −n−1
y1 0 y2 0 yM 0
y1 1 y2 1 yM 1
y1 2 y2 2 yM 2 = V vand B
y1 N − 1 y2 N − 1 yM N − 1
with
1 1 1
z1 z2 zn
1 2 M
Bn Bn Bn
y2 y3 y4 yL
y3 y4 y5 y L+1
Y H = Y ij = y4 y5 y6 y L+2
yL y L+1 y L+2 yN
ᑬL × L , a Hankel matrix
where L = N − M + 1, M = N/2, and the elements Y ij depend only on the sum of the
indices, i.e. Y i + 1,j + 1 = y i + j , i + j = 0, 1, 2, …, L.
An algorithm is now suggested for extracting Prony modes from multichannel
data based on Hankel SVD analysis. As outlined in previous sections, the
minimum norm solution to the linear prediction equation Y a = b is given by
†
a = Y b. A more efficient solution for noisy measurements can be obtained from
SVD analysis of matrix Y .
Suppose the SVD of the Hankel matrix Y H is given by
Y H = UΛV T 7 14
where Λ is a diagonal matrix with positive, real values in the upper left and zeroes
elsewhere, and U,V are the left and right singular vector matrices, respectively.
†
In Refs. [31, 32], it is shown that the pseudo inverse Y H of Y H is related to the
SVD of Y H by the relation
†
YH = V Σ† U T 7 15
in which Σ†is obtained from Σ by replacing each positive diagonal entry by its
reciprocal. Related approaches are introduced in Refs. [22, 32], where connections
are made with the Matrix Pencil method and the eigensystem realization algorithm.
222 7 Small-Signal and Transient Stability Assessment Using Data-Driven Approaches
From the SVD of the Hankel matrix in Eq. (7.15), it is then possible to construct a
least-squares estimate of the Hankel matrix for the ideal, noiseless signal. Once the
model (7.15) has been computed and the LPM is solved, the roots, z, of the char-
acteristic polynomial, are obtained; the associated Ritz eigenvalues [12], and their
modal amplitudes are computed from
λk = conj − log zk Δt
znk = eλk Δt 7 16
j−1
Z jk = znk
where use has been made of (7.18). Physically, (7.20) indicates that the observable
g(xk) is decomposed into vector coefficients, vj, called Koopman modes whose tem-
poral behavior is given by the associated eigenvalues λj.
As discussed further, the phase of the eigenvalues determines its frequency,
while its modulus determines the growth rate. The magnitude φj(xo)vj is used
as a measure of the relative participation of a mode to the modal decomposition.
Refer to [12, 34] for the detailed implementation of the method. Applications to
other power system dynamic phenomena are discussed in [37].
x 1 t1 x 1 t2 x 1 tN
x2 t1 x 2 t2 x 2 tN
X = X N1 = x 1 x2 xN = ᑬm × N
x m t1 x 2 t2 x m tN
x1 xN
7 21
T
where x j = x 1 t j x2 t j … xm t j , j = 1, …, N, and the superscript, N,
denotes end time; m represents the number of sensors or measurement locations.
The method assumes that the data sequences or snapshots, xj, in (7.21) are gen-
erated by a discrete-time linear dynamical system whose evolution is governed by
the linear mapping. Formally, these methods postprocess the sequence of snap-
shots (7.21) and relate two consecutive data fields through a linear mapping of
the form
x k + 1 = Ax k + ηk , k = 0, 1, …, N − 1 7 22
where A is an unknown operator matrix of dimension m×m for a time step Δ t, and
matrix A is real, asymmetrical, and high-dimensional; ηk is some noise process.
This equation defines a Krylov sequence [38]
x 1 = Ax o
x 2 = Ax 1 = A2 x o
x 3 = Ax 2 = A3 x o
7 23
x N = Ax N − 1 = AN x o
or, in a compact form
X N2 ᑬm × N = x 2 x3 xN = A x1 x2 x N − 1 = AX N1 − 1
−1
XN
2 XN
1
7 24
with
7.2 Modal Characterization Using Data-Driven Approaches 225
X N2 ᑬm × N = x 2 x3 xN
X N1 − 1 ᑬm × N = x 1 x2 xN − 1
†
A = X N2 X N1 − 1 7 25
The process of computing A directly from (7.25) may not be feasible, especially
when the size of the data set increases or when the model is ill-conditioned. Several
formulations to estimate the eigenvalues of matrix A have been developed and
tested. For completeness, a brief review of these methods is presented further.
Connections to other approaches are reviewed.
x N = c1 x 1 + c2 x 2 + + cN − 1 x N − 1 7 26
x N ≈ c1 x 1 + c2 x 2 + + cN − 1 x N − 1 = X N1 − 1 c 7 27
X N2 = x 2 x3 xN = x2 x3 X N1 − 1 c 7 28
and hence
X N2 = AX N1 − 1 = X N1 − 1 S 7 29
0 0 0 co
1 0 0 c1
S= 0 1 0 c2 ᑬN − 1 × N − 1 7 30
0 0 1 cN − 1
pS λ = λN − cN − 1 λN − 1 − cN − 2 λN − 2 − …c1 = 0
In practical application, linear dependence in (7.27) occurs gradually. As a
result, a residual is obtained from (7.29) of the form [38]
r = X N2 − X N1 − 1 S
AX N1 − 1 − X N1 − 1 S = reTp
where
0 0 ρ1
reTp =
0 0 ρn
The problem is now how to find the vector c, which determines the residues, r,
and therefore matrix S. From the aforementioned results, the eigenvalues of A
approximate those of S when r 2 0.
Following the earlier discussion, it can be seen that matrix S can be
determined by
argmin X N2 − X N1 − 1 S 2
7 31
S
X N1 − 1 = UΣW T 7 32
7.2 Modal Characterization Using Data-Driven Approaches 227
X N2 = AUΣm W T 7 33
−1
XN
1
X N2 = X N1 − 1 S = UΣW T S = AUΣm W T 7 34
ΣW T S = U T A UΣm W T
S = U T AU 7 35
S = U T AU = U T X N2 W T Σm− 1 7 36
S = Y ΛY − 1 7 37
S = Y ΛY − 1 = U T X N2 W T Σm− 1 7 38
and hence,
X N2 = USΣm W m = U Y ΛY − 1 Σm WTm 7 39
or
X N2 = USΣm W m = UY ΛY − 1 Σm W m = ΦΛΓm t 7 40
spatial Temporal
sructure structure
228 7 Small-Signal and Transient Stability Assessment Using Data-Driven Approaches
Ψ = X T V Σ − 1W
x N = Ax N − 1 = AN − 1 x 1
or x i = Ai − 1 x 1 , i = 2, 3, …, N
From linear system theory, each data vector can be expanded in the form (refer
to Eq. (7.1)). Following [40] let
m
x1 t = a jϕ j 7 41
j=1
and
m
xi t = Ai − 1 a j ϕ j , i = 1, 2, …
j=1
7.2 Modal Characterization Using Data-Driven Approaches 229
It is left to the reader to prove that the time evolution of the ith state can be
expressed as
m
x i t, x = λij− 1 a j ϕ j , i = 1, 2, … 7 42
j=1
where the aj s are the time-dependent amplitudes, the φjs are the eigenvectors of
S, and λj = σ j + iωj is the complex frequency of the associated mode.
Equation (7.42) can be conveniently rewritten in the vector-matrix form
X t = ΦV vand c 7 43
where k = t/Δt.
In what follows, analytical criteria to describe the energy relationships in the
observed oscillations are derived and a physical interpretation of the system modes
is suggested. For simplicity and clarity of exposition, assume that the eigenvectors
φj are normalized to unity, so that
ϕj
ϕj =
ϕj 2
It follows from basic considerations that the total averaged energy, E(T), over a
time interval [0, T] can be expressed as
T T 2
r−1
1 2 1 t Δt
E T = x t dt = λ j c jϕ j dt
T T j=0
0 0
or
T 2
r−1
1 t Δt
T
t Δt
E T = λj c jϕ j λj c jϕ j dt 7 46
T j=0
0
Noting, finally, that elogx = x, and integrating (7.47) with respect to t gives
1 2
T
2σ t 2 e2σ j T − 1
Ej = ϕj λ j j dt = ϕ j 7 48
T 0 2σ j T
7.3 Studies of a Small-Scale Power System Model 231
11 10 4 2 1
∼
∼
3
∼ ∼
12
SC
7 5 6
∼
∼
Tap location
Load Generation
Vref ,...
Plant level
fref ,... control Vter Pgen Qgen
Pgen Drive
ωref
train Pord
Pm
In the studies described further, wind turbines of the wind farms (WFs) are mod-
eled to represent typical (Type 3) doubly fed induction generator (DFIG) config-
urations, similar to the GE WFs’ models [46, 47]. Three main control loops are
considered in the analysis (see Figure 7.2 adapted from [46, 47]):
Case D. A modified case in which the infinite bus 1 and the thermal generator at
bus 6 are replaced by equivalent wind turbine generators.
Frequency Damping
Mode Eigenvalue (Hz) (%) Oscillation pattern
0.501 Hz 0.894 Hz
1 1
Magnitude
Magnitude
0.5
0.5
0
0
6 12 11 4 3 1 12 11 4 3 1 6
Bus Bus
1.433 Hz 1.752 Hz
1 1
Magnitude
0.5 0.5
0
0
–0.5
3 4 11 1 6 12 4 11 1 6 12 3
Bus Bus
2.326 Hz
1
Magnitude
0.5
0
11 1 6 3 12 4
Bus
Table 7.7 Electromechanical modes of the system (Case B with a wind farm at bus 1).
Table 7.8 Electromechanical modes of the system (Case C with a WF farm at bus 6).
In all analyzed cases, damping improves with high wind generation, but as dis-
cussed earlier the effect of wind generation on system dynamic performance
depends on several interacting factors.
In the following subsections, data-driven approaches that supplement informa-
tion on conventional small-signal performance are investigated.
236 7 Small-Signal and Transient Stability Assessment Using Data-Driven Approaches
Table 7.9 Electromechanical modes of the system (Case D with wind farms at buses
1 and 6).
(a)
64.3
Wind farm at bus 1
Frequency (Hz)
64.2
64.1
64
63.9
0 2 4 6 8 10 12 14 16 18 20
Time [sec]
(b)
60.1 Bus 1
Frequency [Hz]
Bus 2
60.05 Bus 3
Bus 4
60 Bus 5
59.95
59.9
0 2 4 6 8 10 12 14 16 18 20
Time [sec]
Figure 7.4 Speed deviations response following a three-phase fault at bus 7 (Case B);
(a) bus 1, and (b) buses 1–5.
7.3 Studies of a Small-Scale Power System Model 237
50
40
30
Angle [deg]
20
Bus 1
10 Bus 2
Bus 3
Bus 4
0 Bus 5
–10
–20
0 2 4 6 8 10 12 14 16 18 20
Time [sec]
Figure 7.5 System rotor angle deviations following a three-phase fault at bus 7 (Case C).
60.04
60.03
60.02
60.01
Frequency [Hz]
60
59.99
Bus 1
59.98 Bus 2
Bus 3
Bus 4
59.97 Bus 5
Bus 6
59.96 Bus 7
Bus 8
59.95 Bus 9
Bus 10
59.94
0 2 4 6 8 10 12 14 16 18 20
Time [sec]
Figure 7.6 System bus frequency deviations following a three-phase fault at bus 7 (Case B).
Due to the small number of measurements, multisignal Prony analysis was used
to identify the modes observed in the transient stability simulations [19].
A discussion of other modal extraction techniques is postponed until Section 7.4.
Table 7.10 shows Prony results for the generator speed deviations and genera-
tors’ active output power (Case B) following a three-phase fault at bus 7. The good
agreement between modal and data-driven approaches in Tables 7.5 and 7.8
238 7 Small-Signal and Transient Stability Assessment Using Data-Driven Approaches
1.08
Bus 1
Bus 2
Bus 3
1.07 Bus 4
Bus voltage magnitude [pu]
Bus 5
Bus 6
1.06 Bus 7
Bus 8
Bus 9
Bus 10
1.05
1.04
1.03
1.02
0 2 4 6 8 10 12 14 16 18 20
Time [sec]
Figure 7.7 System bus voltage deviations following a three-phase fault at bus 7 (Case B).
Table 7.10 Multisignal Prony analysis for speed deviations and generator active power at
buses 1, 2, 4, 6, 11, and 12 (Case B).
ensures the correctness of the model. Results are found to be consistent, although
generator active power is seen to provide a better characterization of the damping
characteristics of mode 1 in Table 7.6, especially for the damping of mode 1.
A comparison of analytical results in Tables 7.6 with data-based results allows us
to validate the small-signal analysis results.
7.3.3.2 Case D
In the discussion on the effect of renewable generation on the system dynamic per-
formance, and to highlight the impact of inertia reduction, a two-WF case is now
considered. In this analysis, generators at buses 1 and 6 were replaced by
7.3 Studies of a Small-Scale Power System Model 239
(a)
64.5
Bus 1
Frequency [Hz]
Bus 6
64
63.5
0 2 4 6 8 10 12 14 16 18 20
Time [sec]
(b)
60.1
Frequency [Hz]
60
Bus 11
Bus 12
59.9 Bus 3
Bus 4
0 2 4 6 8 10 12 14 16 18 20
Time [sec]
Figure 7.8 Speed deviations of system machines and wind farms following a three-phase
fault at bus 7 (Case D); (a) buses 1 and 6, and (b) buses 3, 4, 11, and 12.
equivalent type 3 WFs. For simplicity and clarity of exposition, the same set of
parameters for the control systems was utilized.
Figure 7.8 shows the system response to a three-phase fault at bus 7. Of interest
in this plot are correlated motions of WFs at buses 1 and 6, which suggest the
importance of identifying dynamic patterns and utilizing dynamic equivalencing
techniques.
Further insight into the participation of WFs into system oscillatory behavior
can be obtained from the analysis of PFs for the 1.318 Hz mode in Table 7.8. Sim-
ilar results are obtained for other modes and are not discussed here.
As shown in Figure 7.9, the participation of the WFs to the modes of interest can
be easily identified. The phase of PFs, however, does not have a direct interpreta-
tion, and other measures may be needed to supplement information on energy
exchange, as well as the interaction paths.
Based on this analysis, data-driven approaches are used to supplement informa-
tion on the effect of renewable generation on system dynamic behavior. To dem-
onstrate the application of the aforementioned formulation, consider the time
evolution of speed deviations following a three-phase stub fault at bus 7 in
Figure 7.8. The time window selected for analysis is 20 seconds.
Figure 7.10 shows the time evolution of the extracted modes using Koopman
analysis. Similar plots and results are obtained using DMD and other approaches.
From Figure 7.10, system dynamic behavior is essentially dominated by two well-
damped modes: a 1.21 Hz mode, and a 1.35 Hz mode.
240 7 Small-Signal and Transient Stability Assessment Using Data-Driven Approaches
Gen 3
Angle, speed
1
0.9 Gen 12
Angle, speed
0.8
0.7
Magnitude
0.6
0.5
0.4
WF 1
Gen 4
0.3 WF 1 Converter
Angle, speed
Converter Gen 3
Gen 12 WF 6
0.2 Ψd
Ψq Converter
0.1
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
State
Table 7.11 shows the captured dominant modes. Multichannel Prony analysis
indicates two modes with frequencies near 1.17 and 1.35 Hz that capture over
99% of the observed system response. Mode selection (column 3 in Table 7.11)
was based on the transient energy ratio described in Refs. [43, 49].
0.15
Mode 2
0.1 Mode 1
0.05
Real part of mode
–0.05
–0.1
–0.15
4 6 8 10 12 14 16 18 20
Time [sec]
d
Figure 7.10 Dominant DMD modes from the decomposition j = 0a j t ψ Tj following a
three-phase fault at bus 7.
7.3 Studies of a Small-Scale Power System Model 241
The importance of the modal approximation in Eq. (7.20) or Eq. (7.28) is now
obvious. Simulation results suggest that the time evolution of the system states
can be approximated by the reduced-order model X t = ψ 1 t aT1 t + a2 t ψ T2 ,
where, as discussed earlier, the temporal coefficients ajs provide the temporal
(amplitude) evolution of the modes, and the vector ψ give the approximate mode
shape.
Table 7.12 Voltage-based mode shape of the interarea mode at 1.35 Hz extracted using
Prony analysis for bus voltage magnitudes in Table 7.7 (Case D).
Q
Δω j t = Ai eλi t cos 2π f i t + ϕi , j = 1, …, 5 7 49
i=1
Table 7.12 shows Prony results (PRS) for the speed-deviation signals.
Several conclusions can be drawn for this analysis.
• The WFs at buses 1 and 12 and the WF at bus 6 swing nearly in phase and in
opposition to generators at buses 11, 3, and 4.
• Bus 3 has the largest participation in the mode, followed by the generators at
buses 12, 4, and the WF at bus 1.
0.2
0.1
Real part of mode
–0.1
–0.2
–0.3
–0.4
1 11 12 3 4 6
Bus
Figure 7.11 Speed-based mode shape for mode 1 in Table 7.3 computed using DMD
analysis (Case D).
7.3 Studies of a Small-Scale Power System Model 243
Figure 7.12 shows the time evolution of the centroids for clusters 1 and 2 along
with the buses included in the clusters.
Cluster Buses
1 12
2 4, 6, 11
3 3
4 1
(a)
60.1
Frequency [Hz]
60
59.9 Bus 11
Centroid 1
Bus 4
59.8
0 2 4 6 8 10 12 14 16 18 20
Time [sec]
(b)
60.1
Frequency [Hz]
60.05
60
59.95 Bus 12
Centroid 2
59.9 Bus 3
0 2 4 6 8 10 12 14 16 18 20
Time [sec]
Figure 7.12 Speed signals following a three-phase fault at bus 7; (a) Cluster 1, and
(b) Cluster 2.
Note that, by applying this technique, the center of angle (frequency) can be
determined without simplifying assumptions.
In the system studied thus far, attention has been confined to basic active and reac-
tive control strategies. The purpose now is to introduce some specific control alter-
natives in the study of a practical test system.
61 Area 2
85 Area 3
107
69 1
89
2
62 11
Area 1 194
143
124 137 153 Wind farms
21–26
192 184
3 31 115
105
47
26 Area 7
Area 4 Area 6
8 117 83 15
42 18 109
146 161
Area 5 92
47 95
78 7
144
Wind farms
11–20
Wind farms
1–10
Figure 7.13 Schematic illustration of the seven-area test system showing major buses and
transmission lines.
•• All WFs are modeled by type 3 (DFIG) generic wind turbine models.
Wind plant reactive power control is enabled.
Figures 7.16 and 7.17 show plots of frequency deviations of WFs 1–26 for three
critical contingency scenarios in Table 7.16. In all cases, a nonlinear trend is
observed, especially in the first few seconds of the simulations.
Studies conducted to extract dominant system behavior are discussed further.
Table 7.17 shows the dominant system modes extracted using DMD applied to
the generator speed deviation signals in Figure 7.16.
(a)
Vrfq
KIV Qmax
Vc s +
1 1 1
1 + sTR Σ Σ
fN 1 + sTFV
KPV +
1 + sTV Qmin
Pfarep
tan
(b)
MWbase Pelec wg
MWrat
Vterm
Pmax
Speed RPmax
+ IPmax
1 KIP 1 ÷
π
1 + sTPWR – Σ Σ
KPP +
s sTPP
RPmin
Pmin
MWcap
MWrat
Ipcmd
Figure 7.14 Active and reactive control characteristics: (a) reactive and (b) active control
characteristics. Adapted from [46, 47].
Tmech – + ωo
Pmech Δωt
1 + ωt
÷ ∑ 2Mts ∑
+
Turbine
–
+
Δωtg Δωtg 1
Dshaft ∑ Kshaft
s
– + ωo
Pe + Δωg
1 ωg
÷ ∑ ∑
2Mgs
Te – + Generator
Table 7.18 compares the amplitude and frequency of the extracted modes using
Prony analysis and DMD analysis for the 66-speed deviation signals (data set 1)
with the eigensolution of the linear model x = Ax. Results between Prony analysis
and DMD analysis are found to be in good agreement although some discrepancies
are found.
Frequency Damping
Mode Eigenvalue (Hz) (%) Swing pattern
Contingency
scenario Description
CS01 Loss of the largest unit in the system (generators 1 and 2, Table 8.1)
(a)
44.6
Frequency [Hz]
44.4
44.2
44
43.8
0 5 10 15 20 25 30
Time [sec]
(b)
72.3
Frequency [Hz]
72.2
72.1
72
71.9
0 5 10 15 20 25 30
Time [sec]
Figure 7.16 Frequency deviations for contingency scenario CS04; (a) WFs in Area 6, and (b)
WFs in Areas 2 and 3.
Eigenvalues obtained using a small signal stability program exhibit good agree-
ment with the time-domain studies. Study experience shows that conventional
Prony analysis becomes prohibitive for high-dimensional data sets, and some
expertise is needed to determine both the sampling interval and the study time
window.
250 7 Small-Signal and Transient Stability Assessment Using Data-Driven Approaches
(a)
44.3
Frequency [Hz]
44.2
44.1
44
43.9
0 5 10 15 20 25 30
Time [sec]
(b)
72.2
Frequency [Hz]
72.1
72
71.9
5 10 15 20 25 30
Time [sec]
Figure 7.17 Frequency deviations for contingency scenario CS02: (a) WFs in Area 6 and (b)
WFs in Areas 2 and 3.
1 0.399 0.389
2 0.547 7.000
3 0.666 8.29
3 0.696 7.473
5 0.798 9.527
3
Generators in Areas Generators in Areas
2 2 and 3 5 and 6
1
Frequency [Hz]
–1
–3
–4
–5
0 5 10 15 20 25 30
Time [sec]
Figure 7.18 System response to a three-phase fault at the POIS (contingency scenario
CS03) Signals detrended for clarity of presentation.
Table 7.18 Modal extraction analysis for simulation results in Figure 7.18.
Similar results are obtained using DMD analysis. The overall reconstruction
error, ε = 1.277× 10−3, is computed as
N
2
xk
k=1
ε xk , xk = N
7 52
2
xk
k=1
Table 7.19 Comparison of Koopman and DMD results for data set 2.
(a)
0.08
0.06
Bus frequency deviations [p.u.]
0.04
0.02
–0.02
–0.04
–0.06
5 10 15 20 25 30
Time [s]
(b)
0.08
0.06
Bus frequency deviation [p.u.]
0.04
0.02
–0.02
–0.04
–0.06
5 10 15 20 25 30
Time [s]
Figure 7.19 Bus frequency signals following a three-phase fault at the POIS (contingency
scenario CS03): (a) original, simulated signals, and (b) reconstructed signals using DMD.
7.5 Analysis Results and Discussion 253
obtained from the same disturbance. They could represent measurements from
multisignal records or represent historic information.
Let each data set be of the form X set 1 ᑬN 1 × m1 , X set 2 ᑬN 2 × m2 , …, X set L
ᑬN L × mL . If each data matrix has the same length (N1 = N2 = … = NL), one can
express the concatenated matrix as
X set 1
X set 2
X=
X set L
This motivates the need for a joint analysis of multiple data sets much as in the
case of feature-based fusion of data. Three main issues are of interest here: (i) to
study the complementary nature of data types, (ii) to assess the effect of a given
data type of modal behavior and which combination of data modalities can be used
for specific analysis or control purpose, and (iii) to find common structures and
relationships between data types.
0.15
0.1
Real part of mode
0.05
Wind generators
Synchronous generators
(Areas 1–3)
0
Synchronous generators Area 6
(Areas 6, 7)
–0.05
–0.1
0 10 20 30 40 50 60
Generator/wind farm
References
1. Vittal, E., O’Malley, M., and Keane, A. (2012). Rotor angle stability with high
penetrations of wind generation. IEEE Transactions on Power Systems 27 (1):
353–362.
2. Vittal, V. and Ayyanar, R. (2013). Grid Integration and Dynamic Impact of Wind
Energy. New York, NY: Springer.
3. Gautam, D., Vittal, V., and Harbour, T. (2009). Impact of increased penetration of
DFIG based wind turbine generators on transient and Small-signal stability of
power systems. IEEE Transactions on Power Systems 24 (3): 1426–1434.
4. Power System Dynamic Performance Committee, Task Force on Identification of
Electromechanical Modes, Sanchez-Gasca, J.J. Chair. (2012). Identification of
electromechanical modes in power systems. IEEE/PES, Special Publication TP462
(June 2012).
5. Wilson, D., Bialek, J., and Gustavsson, N. (2008). Identifying sources of damping
issues in the icelandic power system. Proceedings of the 16th Power Systems
Computation Conference (PSCC’08).
6. Castellanos, R.B., Messina, A.R., Gonzalez, M.R., and Guizar, G.C. (2018).
Assessment of frequency performance by grid integration in a large-scale power
system. Wind Energy 21: 1359–1371.
7. CIGRE Working Group C.6.08. (2011). Grid integration of wind generation
(February 2011).
8. Browne, T.J., Vittal, V., Heydt, G.T., and Messina, A.R. (2008). A comparative
assessment of two techniques for modal identification from power system
measurements. IEEE Transactions on Power Systems 23 (3): 1408–1415.
9. Grebe, E., Kabouris, J., López Barba, S. et al. (2010). Low frequency oscillations in
the interconnected system of Continental Europe. Proceedings of IEEE PES
General Meeting.
10. Sanchez-Gasca, J.J., Miller, N.W., Price, W.W. (2004). A modal analysis of a two-
area system with significant wind power penetration. Proceedings of the IEEE PES
Power Systems Conference and Exposition.
11. Messina, A.R. and Vittal, V. (2007). Extraction of dynamic patterns from wide-area
measurements using empirical orthogonal functions. IEEE Transactions on Power
Systems 22 (2): 682–692.
12. Barocio, E., Pal, B.C., Thornhill, N.F., and Messina, A.R. (2015). A dynamic mode
decomposition framework for global power system oscillation analysis. IEEE
Transactions on Power Systems 30 (6): 2902–2912.
13. Feeny, B.F. and Kappagantu, R. (1998). On the physical interpretation of proper
orthogonal modes in vibrations. Journal of Sound and Vibration 211 (4): 607–616.
14. Goebel, C.J. and Epstein, S.T. (1980). Motion of damped oscillators: expansion in
normal modes. American Journal of Physics 48 (4): 289–291.
256 7 Small-Signal and Transient Stability Assessment Using Data-Driven Approaches
15. Hamdan, A.M.A. and Nayfeh, A.H. (1986). Coupling measures between modes and
state variables. International Journal of Control 43 (3): 1029–1041.
16. Kundur, P. (1994). Power System Stability and Control. New York, NY:
McGraw-Hill.
17. Hauer, J. (1991). Application of Prony analysis to the determination of modal
content and equivalent models for measured power system response. IEEE
Transactions on Power Systems 6 (3): 1062–1068.
18. Pierre, D.A., Trudnowski, D.J., and Hauer, J.F. (1990). Identifying reduced-order
models for large nonlinear systems with arbitrary initial conditions and multiple
outputs using prony signal analysis. Proceedings of the 1990 American Control
Conference.
19. Trudnowski, D.J., Johnson, J.M., and Hauer, J.F. (1999). Making Prony analysis
more accurate using multiple signals. IEEE Transactions on Power Systems 14 (1):
226–231.
20. Zhou, N., Pierre, J., and Trudnowski, D. (2013). Some considerations in using
prony analysis to estimate electromechanical modes. Proceedings of the 2013 IEEE
Power & Energy Society General Meeting.
21. Berti, E., Cardoso, V., González, J.A., and Sperhake, U. (2007). Mining information
from binary black hole mergers: a comparison of estimation methods for complex
exponentials in noise. Physical Review D 75 (124017): 1–17.
22. Borden, A. and Lesieutre, B. (2014). Variable projection and Prony analysis for
power system modal identification. IEEE Transactions on Power Systems 29 (6):
2613–2620.
23. Hildebrand, F.B. (1956). Introduction to Numerical Analysis. New York, NY: Dover
Publications, Inc.
24. Potts, D. and Tasche, M. (2013). Parameter estimation for nonincreasing
exponential sums by Prony-like methods. Linear Algebra and Applications 439:
1024–1039.
25. Koehl, P. (1999). Linear prediction spectral analysis of NMR data. Progress in
Nuclear Magnetic Resonance Spectroscopy 34: 257–299.
26. Kul’ment’ev, A.I. (2004). Analysis of positron lifetime spectra via a fast Prony
algorithm. The European Physical Journal of Applied Physics 25: 191–201.
27. Khazaei, J., Fang, L., Jiang, W., and Manjure, D. (2016). Distributed Prony analysis
for real-world PMU data. Electric Power Systems Research 133: 113–120.
28. Golpira, H., Atarodi, A., Amini, S. et al. (2020). Optimal energy storage system-
based virtual inertia placement: a frequency stability point of view. IEEE
Transactions on Power Systems 35 (6): 4824–4835.
29. Barkhuijsen, H., De Beer, R., and Van Ormondt, D. (1987). Improved algorithm for
noniterative time-domain model fitting to exponentially damped magnetic
resonance signals. Journal of Magnetic Resonance 73: 553–557.
References 257
modern power systems are expected to introduce new interactions and operational
challenges [1–6].
Typically, factors that lead to integration issues include vulnerability to major
perturbation, reduced system inertia, rotor angle, and voltage stability, transient
stability and frequency performance, and postdisturbance oscillations [7–13]. In
addition, the successful integration of distributed RESs requires an increased level
of visibility (and controllability) of these devices [14, 15]. A minimum level of
expected performance during power system disturbances is needed, which in turn
requires knowledge of power system dynamic characteristics and the aggregated
behavior of distributed RESs [16]. Sensitivity studies and data-based analysis tech-
niques are of interest since they can supplement information on the system behav-
ior, for instance, to understand the commitment and dynamic patterns, arising
from the integration of new wind and PV generation.
Some recent investigations have also revealed a significant decline in the fre-
quency performance of some major grids [10, 11, 17], prompting concerns on
the effect of increased wind and solar penetration on the future inertial frequency
performance. Understanding the nature of this trend is therefore critical to coor-
dinating operational and control actions.
Recent advances in inertia emulation techniques have allowed modern wind
turbines to provide inertial response functionalities under low-frequency condi-
tions [18, 19]. Other developments include the use of advanced schemes for inertia
emulation [20]. However, as discussed by several authors, the role of inverter-
based inertial response and wind generation frequency response characteristics
is not well understood [17], or knowledge of the real capabilities and limitations
of inertial response from wind turbines is limited [21].
Other factors affecting system dynamic response include [22, 23]:
• The lack of adequate dynamic voltage support near areas or zones with a high
concentration of renewable generation,
• The intermittent nature and variability in power flow patterns and a more com-
plex real and reactive power dispatch,
• Inertia reduction resulting from large increases in wind and solar generation
and the associated displacement of thermal and hydro generation, along with
high reactive power losses; and
Figure 8.1 is a simplified single-line diagram of the study system showing the main
areas of concern for this study. Key buses and major transmission paths are indi-
cated on the schematic map. The base system used in the studies described further
262 8 Solar and Wind Integration Case Studies
Areas 1–3
Wind-farms
20–26
PV farms
1–3
Area 5
Area 4
Area 7
± 300
MVAr
Hydro
SVC 2 ∼ Generation
Area 6 (4970 MW)
SVC 1
± 300
Wind-farms
MVAr 11–19
Wind-farms 2500
MW
1–10
is an expanded version of the given test system in Chapter 7 and includes the
detailed representation of 635 generators, 26 explicitly modeled inverter-based
wind farms (WFs), 3 solar PV farms, 5449 buses, and 5292 transmission lines. Tur-
bine governors are represented on major machines. See Section 7.1 for detailed dis-
cussions of other modeling considerations. This model has been validated using
measurement data from synchronized phasor measurement units (PMUs) [33].
Two wind and solar penetration scenarios are of interest here: (i) grid integration
of wind energy in Area 6 and (ii) wind and solar integration in Area 4 and Area 5.
The emphasis in the following analyses is placed on the scenario (i).
To evaluate the performance impact of wind and solar PV farms, appropriate
models of wind and solar farms have been developed together with special reme-
dial action schemes. Simulation studies include the detailed representation of the
machines that provide governor response and discrete control devices such as
underfrequency load shedding (UFLS) schemes.
8.3 Wind Power Integration in the South Systems 263
The integration studies described further have two specific objectives: (i) define
limits of integration of wind energy in the system, and (ii) evaluate actions to
improve frequency regulation and voltage control on the 400 kV grid.
Because of its geographical location and the sparse nature of the bulk transmission
network, wind integration in Area 6 is of special concern. In this system, WFs 1–19
are located at the south extremity of Area 6; the current installed wind capacity in
the study region is about 50% of the local hydropower generation. As wind pene-
tration in the study area is expected to increase, hydro plants providing governor
response will likely be displaced or be required to operate at reduced output (spin-
ning reserve) to accommodate wind variations and provide primary frequency
response as discussed in the following subsections.
Further, faults in this system resulting in loss-of-generation events will affect fre-
quency performance and result in postdisturbance oscillations.
Area
To Area Gens 1,2 7
7
∼
∼ Gen 14
∼
∼ Gen 13
Vpois 4970 MW
∼
SVC 2 Gens 11,12
∼ Gen 10
SVC 1
Areas Hydro machines
4 and 5 Area 6
Wind
farms 1–19
(2500 MW)
Figure 8.2 Simplified schematic of the study region within Area 6. Dashed lines represent
interconnections to neighboring systems.
• Wind generation was uniformly increased in proportion to its base case condi-
tion and generation at local hydro generators was decreased. The rationale
behind the adopted approach is discussed in Section 8.5.2.
• Among the various generating sources, hydro generators 10–14 are used for bal-
ancing wind fluctuations and utilized as a spinning reserve for major
contingencies.
1 Nuclear 750 No No
2 Nuclear 750 No No
3 Hydro 7.5 No No
4 Hydro 66 No No
5 Thermal 46 No No
6 Thermal 46 No No
7 Thermal 7 No No
8 Hydro 122 No No
1) Wind power concentration: The concentration of wind generation and low load
density within the study region. As discussed earlier, a large amount of wind
power is concentrated on a small geographical region.
2) Wind variability: Wind power fluctuations ranging from 0 to 100 MW within an
hour are not uncommon, especially early in the morning (during off-peak
hours). Further, wind generation is out of phase with load demand thus adding
to system variability.
3) Relative size: As noted earlier, the size of the WFs relative to hydro generation in
the south systems is high (approximately 50%). As a result, the automatic gen-
eration control (AGC) system must compensate for fluctuations in wind power
and load variations,
4) Point of interconnection: WFs 1–19 are connected radially to the 400 kV trans-
mission system. A large ±300 MVar SVC is used to support local voltage. As a
result, loss of SVC voltage support or loss of major transmission lines may result
in system instability or poor postdisturbance system response.
5) Impact on reserve margin: Present levels of wind power penetration in the south
systems may affect operating reserve margins.
In addition to these limitations, power transfers to the bulk 400 kV system are
constrained by grid stability and voltage considerations.
Detailed transient stability studies were conducted to evaluate the impact of high
wind penetration on system stability and performance. The following discussion
describes system response in the context of grid integration studies.
Contingency
scenario Remarks
In these analyses, three scenarios of wind penetration were selected for study:
Wind scenario 1 (WS1): This is the base operating case with about 400 MW wind
generation in the south systems, representing about 4% wind penetration rela-
tive to the installed hydro generating capacity in generators 10–14 in Figure 8.2.
Wind scenario 2 (WS2): A low wind generation scenario.
Wind scenario 3 (WS3): A 20% wind scenario.
In all cases, flow patterns were adjusted to reflect current operating policies in
the system. Note that, because some of the hydropower plants are equipped with
PSSs, adding wind generation may affect the damping of major interarea modes
associated with the study region.
80
WF 10
70
60
50
WFs 3,9,17,19
Power [MW]
40
30 WFs 14,15
20
10
WFs 4–8, 11–13,16,18
0 WFs 1,2
0 5 10 15 20 25 30
Time [sec]
Figure 8.3 WFs’ active power output following a stub three-phase fault at the POIS
(contingency scenario CS03).
1.04
1.03
1.02
Voltage [pu]
1.01
SVC 1
1
0.99
0.98
0 5 10 15 20 25 30
Time [sec]
Figure 8.4 Wind farms terminal voltages following a three-phase stub fault at the POIS
(contingency scenario CS03).
For reference, Table 8.4 presents a comparison of the frequency and damping of
the dominant mode for the speed deviations of WFs for the contingency scenarios
in Table 8.3. In all cases, WFs are found to have dominant participation in the
8.4 Impact of Increased Wind Penetration on the System Performance 269
200
100
Reactive power [MVAr]
–100
–200
SVC 1
–300 SVC 2
SVC 3
0 5 10 15 20 25 30
Time [sec]
Figure 8.5 SVC reactive power output following a three-phase stub fault at the POIS
(contingency scenario CS03).
0.394 Hz mode. From these results, contingency scenarios CS03, CS06, and CS07
are selected for analysis in the discussion that follows.
60.1
60.08
60.06
Frequency [Hz]
60.04
WFs 20–26 WFs 1–19 0.396 Hz
60.02
60
59.98
0 5 10 15 20 25 30
Time [sec]
Figure 8.6 WFs’ bus frequencies following a three-phase stub fault at the POIS
(contingency scenario CS03).
270 8 Solar and Wind Integration Case Studies
600 WF 12
500
400 WF 1
300 WF 14
Qwind [MVAr]
200
100 WFs 2,4–5,9–11,13–19
0
–100
WF 8
–200
WF 7 WFs 3,6
–300
–400
0 5 10 15 20 25 30
Time [sec]
Figure 8.7 Wind farms reactive power output (contingency scenario CS03).
Table 8.4 Koopman analysis of rotor speed deviations for different contingency scenarios
(wind scenario WS1). The bold entries indicate critical contingency scenarios with the lowest
post-disturbance damping characteristics.
WFs in areas 1–3 have an important participation in the slowest interarea mode in
the system.
Insight into the nature of the WFs’ contribution to these oscillations can be
gleaned from the analysis of the mode phase for selected machines. Table 8.5
shows the amplitude and phase associated with the 0.394 Hz mode for WFs 1,
5, 24, and 25 and Gens 10–13 in Table 8.1, extracted using Koopman mode anal-
ysis. It is observed that WFs 1 and 5 swing in opposition to WFs 24 and 25 and Gens
8.4 Impact of Increased Wind Penetration on the System Performance 271
Table 8.5 Phase relationships for dominant system generators extracted from generator
speeds (contingency scenario CS03).
10–13 suggesting a local exchange of energy through the 400 kV transmission sys-
tem in Figure 8.2. WFs 1 and 5 are found to have the largest participation in the
0.394 Hz mode.
20
WFs
3,17,19
15
10 WFs
4,5,7,8,16
5
Q [MVAr]
–5
WFs 9,10
–10
–15
0 2 4 6 8 10 12 14 16 18 20
Time [sec]
700
SVC 12
600
500
400
Q [MVAr]
300
200
SVC 1
100
0 SVC 8
–100 SVC 7
–200
0 2 4 6 8 10 12 14 16 18 20
Time [sec]
simulations for contingency scenario CS03. In this analysis, it is assumed that wind
generation displaces hydroelectric generators equipped with PSSs.
From the results of Chapter 7, simulated data are approximated by a Koopman
representation of the form
m
yk = λkj ψ j x o v j 81
j=1
Close analysis of simulations for high wind generation cases in Figures 8.8 and
8.9 shows undamped reactive power oscillations involving two major electrome-
chanical modes: a local mode at about 1.36 Hz and the 0.394 Hz interarea mode
1 (m = 2 in Eq. (8.1)). The 1.36 Hz mode is of special interest since it has no coun-
terpart in linear (small signal) studies; similar findings are described in Ref. [29].
Simulation results highlight three key points: (i) careful analysis of Figures 8.8
and 8.9 shows that reactive power outputs at WFs swing in phase with nearby
SVCs (SVCs 1,7, 8, and 12); (ii) while not explicitly addressed, the WFs swing
out of phase with nearby hydro generators (Gens 10–14); and (iii) the oscillations
manifest themselves as nonlinear oscillations with a dominant component at
1.36 Hz.
As a first step toward assessing the nature of energy exchange associated with
these modes, phase relationships for several generators and wind generators were
determined using dynamic mode decomposition (DMD) analysis in Chapter 7. In
the interest of simplicity, Gens 1–14 in Table 8.1 and 19 WFs (Gens 1–19) were
selected for the study.
8.4 Impact of Increased Wind Penetration on the System Performance 273
Toward this end, the measurement matrix, Xω, associated with wind and SG in
Area 6 is represented as a row-concatenated matrix
X ω = X ωwind X ωgen 82
X ω = x1 x2 … xd = ΦΛΓm t 83
or
X ω = x1 x2 … xd = ΦV vand c 84
where, for clarity of illustration, the DMD method is used in the analysis, and the
symbols in Eqs. (8.3) and (8.4) have the usual interpretation in Section 7.2.3.
In the studies presented further, generation was rescheduled to account for
increases in wind generation. Starting from the base case, wind generation was
increased, and active and reactive power was rescheduled.
For reference and validation, Table 8.6 shows the extracted modes using
Koopman/DMD analysis for the high wind penetration scenario. Similar results
are obtained for other operating conditions.
Further, Figure 8.10 shows the speed-based mode shape for the 1.36 Hz mode
extracted using this approach.
Analysis of the mode shape in Figure 8.10 reveals several significant results.
First, WFs 1–19 swing in opposition to machines 1–14 revealing the local nature
Table 8.6 The slowest modes of oscillation of the test system (contingency scenario CS03).
1 0.399 0.357
2 0.545 7.001
3 0.696 5.26
4 0.701 7.42
5 0.748 3.512
274 8 Solar and Wind Integration Case Studies
0.25
Gen 10
Gens 11,12
0.2
0.15
Real part of mode
Gens 13,14
0.1
Gen 8
0.05
WFs 1–19
0
Generator
WFs 9–11
–0.05
Generator/wind farm
of this mode; Gens 10–12 (hydro machines) and Gens 13–14 are seen to have the
largest contribution to the mode. The WFs, on the other hand, show a similar con-
tribution in terms of amplitudes and exhibit the largest participation in the mode.
Physically, the form of the mode indicates the oscillatory mechanism involved in
the observed time-domain responses.
a loss of generation event in the test system [33]. As discussed in Chapters 4 and 7,
the location of contingency reserves plays a critical role in the ability of the system
to limit frequency variations [34]. A simple criterion to estimate the effect of gen-
erator dropping on the frequency behavior at bus j, Δfj, can be obtained from the
sensitivity relationship [27]
∂ f j ∂Pgout
Δf j = ΔPwgen 85
∂Pgout ∂Pwgen
where Pgout and Pwgen are the amount of generator outage and WF active power,
respectively, and the sensitivities can be estimated from measurements or simula-
tion studies involving frequency behavior to selected generator outages and wind
generation scheduling strategies.
Another way of looking at frequency dynamics is by using two-area inertia-
based dynamic equivalents derived from the center-of-gravity (COG) formulation
in Chapter 6. Figure 8.11a gives a conceptual representation of this model, show-
ing a recorded frequency disturbance event following the loss of a large hydro gen-
erating plant in Area 6. Figure 8.11b, in turn, shows the equivalent system for the
analysis of wind and hydro coordination and frequency response analysis.
Using the same notation as in Section 6.2, the frequency deviation of the partial
centers of inertias (COIs) can be approximated as [35, 36]
Mi ΔPmech − ΔPelec
Δ f COI i = Δ f COG 86
M COG 2πD
1+ Δ f COG
M i ΔPmech − ΔPelec
From (8.6), the relationship between the frequency of the COG and the motion
of the local centers of angle is first determined, and expressions to compute local
frequency deviations following major disturbances can be derived. Formally, the
frequency deviation at bus j can then be approximated as
Mj ΔP j − Ptie
ij − η j ΔP i
MMG
Δf j = Δ f COI i 87
Mi 2πD
1+ Δ f COI i
M j ΔP j − P − η ΔPMMG
tie
ij j i
(a)
60 Hz
Intertial
response Area 2
Nadir
Local COI
dynamics
∂fi ∂fCOIi
Δfj(t) = Δf (t)
∂fCOIi ∂fCOG COG
(b)
P
COG SG
–C
OG
G
CO
F–
W
P
PWF–SG ∼ Gen 14
VPOIS
∼ Gen 13
∼ Gens 11,12
∼ Gen 10
WFs 1–19 SVC 1
Hydro machines
Figure 8.11 The COG dynamic equivalent adopted for assessing energy exchange and wind
and hydro coordination; (a) multiarea representation and (b) two-area dynamic equivalent.
8.5 Frequency Response 277
Once frequency sensitivities are estimated, the partial sensitivities in (8.5) can be
determined, as suggested in Section 8.5.3.
To examine the applicability of this approach, active power responses were cal-
culated for each wind generator and SG in the study system.
From partial least squares (PLS) regression analysis in Chapter 5, the measure-
ment data are defined as
T
X = X sm = P gsmj1 P gsmj2 … P gsmj14
T
88
Y = X wf = P gwfj1 P gwfj2 … P gwfj19
X = t 1 pT1 + E = TPT + E
89
Y = u1 qT1 + F = UQT + F
(a)
1000
900
Power [MW]
Gen 1
800
700
600
2 4 6 8 10 12 14 16 18 20
Time [sec]
(b)
50
Power [MW]
40 Gen 6
30 Gen 9
20
0 5 10 15 20
Time [sec]
(c)
400
Gen 11
300
Power [MW]
200 Gen 13
Gen 14
100
Gen 10
0
2 4 6 8 10 12 14 16 18 20
Time [sec]
Figure 8.12 Generator active power response following the simultaneous outage of two
generators (Contingency scenario CS07) for: (a) nuclear generation, (b) thermal generation,
and (c) hydro generation.
In addition, one can infer from Figure 8.13 that WFs 3, 9–10, and 17, 19 have the
largest contribution to the postdisturbance scenarios. In turn, nuclear generators 1
and 2 and hydro machines 9–13 show the largest contribution to the oscillation;
the opposite is true for machines 3–8. Because nuclear generators are not expected
to provide governor response, hydro generators 9 through 13 are selected for
analysis.
280 8 Solar and Wind Integration Case Studies
(a)
200
150
Magnitude
100
50
0
0 5 10 15 20
Wind farm
(b)
300
Nuclear
generators
Magnitude
200 Hydro
generators
100
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14
Generator
Figure 8.13 Partial least squares regression between WFs and generation sources in the
south systems (contingency scenario CS07); (a) WFs 1–19 and (b) selected generators in
Area 6.
From the system viewpoint, this information helps to easily identify the location
and distribution of generation to be displaced; these findings can also be used to
coordinate operational actions. A related issue is that of inertia support in hydro
machines, as some of them could be operated as synchronous condensers.
60
Base case
59.98 Base case with 7% wind penetration
Base case with 15% wind penetration
59.96
59.94
Frequency [Hz]
59.92
59.9
59.88
59.86
59.84
59.82
59.8
0 5 10 15 20 25 30
Time [sec]
Figure 8.14 Frequency response at the POIS following the outage of a 750 MW nuclear
unit (contingency scenario CS06).
results in Section 8.5.2. The metrics of interest are the RoCoF, the frequency nadir,
and the settling frequency [10, 11].
Figure 8.14 is an example of postdisturbance frequency oscillations following a
750-MW generation loss (contingency scenario CS06). As shown in this plot, large
loss-of-generation events result in negatively damped oscillations with a frequency
of about 1.36 Hz (5–25 seconds). The mode at 1.36 Hz is identified as a local oscil-
lation between the western portion (WFs 1–19) and the eastern portion (hydro
generators) of Area 6.
For the base case, the frequency response reaches a nadir of about 59.9 Hz at
approximately 3.8 seconds after the generation trip; while not shown in
Figure 8.14, the frequency settles back to the prefault (nominal) condition at about
40 seconds. Several differences are obvious when comparing system performance.
It can be seen that the slope of the inertial response increases as the amount of
connected wind generation increases. Additional studies show that this effect is
more pronounced in the case of stressed operating conditions.
Table 8.7 shows the eigenvalues computed using Koopman mode analysis for
the contingency operating conditions in Figure 8.4. In this analysis, the study win-
dow is chosen to coincide with the frequency primary response period (5–15
seconds).
Analytical experience with other operating scenarios shows that frequency
nadirs range from 59.92 Hz for low wind penetration levels to 59.65 Hz for high
wind penetration levels for the cases with double- and triple-generation
282 8 Solar and Wind Integration Case Studies
Table 8.7 Prony results for frequency deviation signals in Figure 8.14 following the loss of
critical system generators for contingency scenarios CS06 and CS07 in Table 8.1 (time
interval 15–30 seconds).
CS07-loss of two
CS06-loss of unit 1 of generating units
Gen 1 (750 MW) (180 MW each)
59.95
59.9
300 MW generator tripping
Frequency nadir [Hz]
59.85
59.8
59.7
59.65
1000 1500 2000 2500 3000
Wind penetration level [MW]
The advent of modern, inverter-based wind and solar farms offers a direct means
for controlling system voltage. Specifically, the WFs in the south systems are
expected to provide local voltage support (Figure 8.16). Two main factors influence
the effect of voltage control on system dynamic performance: (i) the location, size,
and control characteristics of reactive support sources near the wind/solar PV
plants; and (ii) the control characteristics of WFs.
The first issue is studied here. Based on these studies, the need for coordinated
voltage control is established.
Vref
Ibranch +
1
∣Vreg – (Re+ jXc)Ibranch∣ 1
∑
1 + sTstfr –
Vreg 0 qmax
emax Qext
+ K
Qbranch + KPq + qi 1
∑ s 1 + sTp
Kc
emin
qmin
Qbranch 1
∑
1 + sTstfr –
+
Qref
Figure 8.16 depicts the WF reactive power control loop used in the simulations.
The primary voltage control loop includes a proportional–integral controller and
droop control; the adjustable droop is obtained by the feedback of the WF reactive
current and voltage and modulates the reference voltage, Vref. The difference value
is then processed through the PI controller to obtain the reactive command for the
output, Qext.
Further, through the modification of the reactive power reference, Qref, the WFs
can maintain constant reactive power output. A slow reactive control loop, auto-
matic reactive power regulator (AQR) can also be utilized for the WFs. A suitable
coordination control between the SVCs and the WFs [39] may be necessary as dis-
cussed in Chapter 5.
As a first step, extensive power flow analyses were conducted to determine the
impact of transmission system limitations on the energy deliverability of the wind
plant output. In these simulations, static compensation with constant reference
voltage control is considered at SVCs 1 and 2 in Figure 8.2. Table 8.8 shows the
SVC output reactive power as a function of the WFs’ reactive power output
Table 8.8 SVC reactive power as a function of WFs’ loading. The voltage control loop in
Figure 8.16 is enabled.
a MVAr.
8.6 Effect of Voltage Control on System Dynamic Performance 285
obtained from a steady-state power flow solution for the case when WFs 1−19
operate on voltage control mode. Three cases are considered, namely (i) wind sce-
nario 1 (WS1) as the base case, (ii) wind scenario 2 (WS2), and (iii) wind scenario
3 (WS3).
By increasing the WFs’ output, the reactive losses on the transmission system
connecting the WFs with the 400 kV system (see Figure 8.11) increase, and the
margin of dynamic reactive power to support voltage recovery decreases. Increased
reactive losses, in turn, increase the possibility of voltage collapse as discussed
further.
Based on these results, studies were conducted to assess voltage stability con-
straints. Emphasis was placed on the impact of solar and wind generation on
power system performance. From the perspective of system modeling, three main
aspects are of interest [1]: (i) optimizing voltage control, (ii) tuning voltage/VAR
control loops [40], and (iii) providing sufficient static and dynamic reactive power
capability [41]. These features are interrelated as discussed further.
To pursue these concepts, Figure 8.17 shows plots of terminal voltage at SVC 1
for various control strategies. The postdisturbance oscillatory response shows a
marginally stable ( f = 1.366 Hz, ξ = 0.011% oscillation and a negatively damped
oscillation at 0.396 Hz (ξ = − 0.759%). Modal analysis discloses higher harmonics
at 2.732 Hz (ξ = − 0.012%) revealing the nonlinear nature of the observed
oscillations.
(a)
1.03
Voltage [pu]
1.028
1.024
0 5 10 15 20 25 30
Time [sec]
(b)
1.03
Voltage [pu]
1.028
0 5 10 15 20 25 30
Time [sec]
Figure 8.17 Bus voltage magnitudes at the POIS for various control and dynamic support
characteristics (contingency scenario CS06); (a) SVC voltage support at the POIS and nominal
SVC gain and (b) SVC voltage support at the POIS with low SVC gain.
286 8 Solar and Wind Integration Case Studies
Bus 81
1.1 Bus 132
Bus 136
Bus 73
Bus 40
Bus voltage magnitude [pu]
1.05
0.95
0.9
0 5 10 15 20 25 30
Time [sec]
A crucial observation is that the study results in Figure 8.17a show that constant
Q control may result in mid-term voltage instability. In turn, it can be seen in
Figure 8.17b that improper SVC tuning can also result in degraded performance
and severe voltage fluctuation or instability. Voltage instability is observed in
Figure 8.17a at about 28 seconds into the simulation. The effect becomes more pro-
nounced as the system is further stressed.
The need for additional (or more coordinated) dynamic reactive support is evi-
dent and deserves further investigation.
Based on the preceding results, system studies were conducted to identify modal
characteristics in the observed oscillations. Figure 8.18 illustrates the results of
transient stability simulations for contingency scenario CS06.
The nature of these oscillations becomes evident from the Fourier spectra of the
observed oscillations in Figure 8.19 and Prony analysis of selected simulations in
Table 8.9. From this Table, hydro plant generators and WFs in Area 6 (WFs 1 and
8) are seen to swing in opposition to WFs in Area 3 (WFs 23 and 26) and solar PV
farm 1 in Area 4 as expected from physical considerations. It should be noted that
the larger modal voltage swings for the dominant mode are located in zones or
buses without voltage regulation (Bus 73) in agreement with Fourier spectral anal-
ysis in Figure 8.19.
8.6 Effect of Voltage Control on System Dynamic Performance 287
Bus 81
40 Bus 136
Bus 73
Bus 40
20
Magnitude [dB]
–20
–40
–60
–80
0 0.5 1 1.5 2 2.5 3
Frequency [Hz]
Figure 8.19 Fourier spectra of bus voltage magnitude deviations in Figure 8.18.
WF 1 6 0.0105 −85.22
WF 23 3 0.0059 32.12
WF 26 3 0.0061 39.04
PV 1 4 0.0032 43.05
288 8 Solar and Wind Integration Case Studies
8.7 Summary
In this chapter, studies to evaluate the impact of high wind penetration levels on
the dynamic performance of a large, realistic power system model have been con-
ducted. The studies examine the effect of geographically disperse wind and solar
generation on transient stability and frequency regulation. The major emphasis
was directed toward the problem of wind and hydro coordination and the study
of the effect of coordinated voltage control on system dynamic performance.
The applicability of correlation techniques to determine the conventional SG gen-
eration to be displaced by WF integration is also pointed out.
System stability studies have shown that dynamic performance considerations
including voltage collapse, oscillatory and frequency instability, and postdistur-
bance oscillations may limit high wind and solar PV integration. The combination
of these constraints makes the analysis of the high integration of wind and solar PV
penetration difficult.
References
1. Miller, N.W., Larsen, E.V., and MacDowell, J.M. (2004) Advanced control of wind
turbine-generators to improve power system dynamic performance. Proceedings of
the IEEE International Conference on Harmonics and Quality of Power.
2. CIGRE Working Group C.6.08. (2011). Grid integration of wind generation
(February 2011).
3. Alberta Electric System Operator, Wind Integration Impact Studies Phase 2. (2006).
Assessing the impacts of increased wind power on AIES operations and mitigating
measures (18 July 2006), Revision 2.
4. Miller, N.W., Guru, D., and Clark, K. (2008). Wind generation applications for the
cement industry. Proceedings of the 2008 IEEE Cement Industry Technical
Conference Record.
5. Winter, W. (ed.) (2010). European wind integration study (EWIS) – EWIS final
report, Brussels. http://www.wind-integration.eu (accessed 2020).
6. Sharma, S., Huang, S.H., and Sarma, N., System inertial frequency response
estimation and impact of renewable resources in ERCOT interconnection.
Proceedings of the 2011 IEEE Power and Energy Society General Meeting.
7. European Network of Transmission System Operators for Electricity (ENTSO-E).
Future system inertia – Report prepared by Energienet.dk, Fingrid, Stanett and
Svenska krafnät, Brussels. https://eepublicdownloads.blob.core.windows.net/
public-cdn-container/clean-documents/Publications/SOC/Nordic/
Nordic_report_Future_System_Inertia.pdf (accessed 2020).
References 289
8. Vittal, E., O’Malley, M., and Keane, A. (2010). A steady-state voltage stability
analysis of power systems with high penetrations of wind. IEEE Transactions on
Power Systems 25 (1): 433–442.
9. Vittal, V. and Ayyanar, R. (2013). Grid Integration and Dynamic Impact of Wind
Energy. New York, NY: Springer.
10. Tan, J., Zhang, Y., You, S. et al. (2018). Frequency response study of a U.S. Western
interconnection under extra-high photovoltaic generation penetrations.
Proceedings of the IEEE Power and Energy Society General Meeting.
11. Tan, J., Zhang, Y., Veda, S. (2017). Developing high PV penetration cases for
frequency response study of U.S. Western Interconnection. Proceedings of the
IEEE Green Technologies Conference.
12. Quintero, J., Vittal, V., Heydt, G.T., and Zhang, H. (2014). The impact of increased
penetration of converter control-based generators on power system modes of
oscillation. IEEE Transactions on Power Systems 29 (5): 2248–2256.
13. Clark, K., Freeman, L.A., Jordan, G.A. et al. (2010). Impact of high levels of wind
and other variable renewable generation on the grid operation: summary of major
US studies. CIGRE 2010, C5_209_2010.
14. AEMO. (2019). Australian energy market operator, maintaining power system
security with high penetrations of wind and solar generation. International
Insights for Australia (October 2019).
15. MIGRATE – Massive InteGRATion of Power Electronic Devices. (2016).
Deliverable D.1.1, Report on Systemic Issues. Tech. Rep. https://www.h2020-
migrate (accessed 2020).
16. European Network of Transmission System Operators for Electricity (ENTSO-E).
(2014). Dispersed generation impact on CE region security: dynamic study – 2014
report update – report of ENTSO-E SG SPD. http://www.entsoe.eu (accessed 2020).
17. Gevorgian, V., Zhang, Y., and Ela, E. (2015). Investigating the impacts of wind
generation participation in interconnection frequency response. IEEE Transactions
on Sustainable Energy 6 (3): 1004–1012.
18. Caldas, D., Fischer, M., and Engelken, S. (2015). Inertial response provided by full-
converter wind turbines. Proceedings of the Windpower 2015 Conference and
Exhibition.
19. Laudahn, S., Seidel, J., Engel, B. et al. (2016). Substitution of synchronous
generator based instantaneous frequency control utilizing inverter-coupled DER.
Proceedings of the 2016 IEEE Seventh International Symposium on Power
Electronics for Distributed Generation Systems (PEDG).
20. Hydro-Québec TransÉnergie (2009). Transmission provider technical
requirements for the connection of power plants to the hydro Québec transmission
system. http://www.hydroquebec.com/transenergie/fr/commerce/pdf/
exigence_raccordement_fev_09_en.pdf (accessed February 2009).
290 8 Solar and Wind Integration Case Studies
21. Fischer, M., Engelken, S., Mihov, N., and Mendonca, A. (2016). Operational
experiences with inertial response provided by type 4 wind turbines. IET Renewable
Power Generation 10 (1): 17–24.
22. Gautam, D., Vittal, V., and Harbour, T. (2009). Impact of increased penetration of
DFIG-based wind turbine generators on transient and small signal stability of
power systems. IEEE Transactions on Power Systems 24 (3): 1426–1434.
23. MIT Energy Initiative. (2011). Managing large-scale penetration of intermittent
renewables. An MIT Energy Initiative Symposium (20 April 2011).
24. NERC. (2012). Frequency response initiative report - the reliability role of
frequency response. Draft. http://www.nerc.com/docs/pc/FRI%20Report%209-30-
12%20Clean.pdf (accessed September 2012).
25. North American Electric Reliability Corporation. (2017). Integrating inverter based
resources into low short circuit strength system, reliability guideline
(December 2017).
26. Milligan, M., Lew, D., Corbus, D. et al. (2009). Large-scale wind integration studies
in the United States: preliminary results. Proceedings of the Eighth International
Workshop on Large Scale Integration of Wind Power and on Transmission
Networks for Offshore Wind Farms.
27. Bustamante, R.C., Roman Messina, A., Gonzalez, M.R., and Guizar, G.C. (2018).
Assessment of frequency performance by grid integration in a large-scale power
system. Wind Energy 21: 1359–1371.
28. Acker, T. (2011). National Renewable Energy Laboratory, IEA Wind Task 24,
Integration of Wind and Hydropower Systems, Volume 1: Issues, Impacts, and
Economics of Wind and Hydropower Integration. Tech. Rep. NREL/TP-5000-50181
(December 2011).
29. You, S., Kou, G., Liu, Y. et al. (2017). Impact of high PV penetration on the inter-
area oscillations in the U.S. Eastern Interconnection. IEEE Access 5: 4361–4436.
30. Tuttelberg, K., Kilter, J., Wilson, D., and Uhlen, K. (2019). Estimation of power
system inertia from ambient wide area measurements. Proceedings of the 2019
IEEE Power & Energy Society General Meeting.
31. Begovic, M.M. and Messina, A.R. (2010). Wide area monitoring, protection and
control, special issue on wide-area monitoring, protection and control. IET
Generation, Transmission & Distribution: 1083–1085.
32. Miller, N.W., Shao, M., D’aquila, R. et al. Frequency response of the US Eastern
Interconnection under conditions of high wind and solar generation. Proceedings
of the 2015 Seventh Annual IEEE Green Technologies Conference.
33. Martinez, E. and Messina, A.R. (2011). Modal analysis of measured inter-area
oscillations in the Mexican Interconnected System: The July 31, 2008 event.
Proceedings of the 2011 IEEE Power and Energy Society General Meeting.
34. Muljadi, E., Gevorgian, V., Singh, M., and Santoso, S. (2012). Understanding
Inertial and frequency response of wind power plants. Proceedings of the IEEE
Symposium on Power Electronics and Machines in Wind Applications.
References 291
Index
h l
Hadamard product 163 LCC, see life cycle cost 119, 120, 134
Hankel matrix 221, 222 least-squares 218, 222, 223
Hankel-SVD 221 life cycle cost (LCC) 119
hierarchical levels of voltage control 7 linear prediction model (LPM) 217
high voltage direct current (HVDC) 6, 27, load damping 33
103, 104, 261 load frequency control (LFC) 2
high-frequency switches 58 load shedding 9, 13, 44, 73, 104, 185,
197, 246
i local frequency 27, 28, 33, 35, 67, 185, 275
ideal voltage source 61 local loads 61
IEEE 50-machine test system 35, 44, 85, 89 long-term power frequency transients 186
impulse response 116 low-frequency equivalent models 27
inertia emulation 105, 114, 116, 124, low-frequency models 26, 27
185, 260
inertia-based control 104, 188, 189, 191, m
197, 199, 200, 201, 202, 203, 204 mapping 114, 160, 161, 162, 173, 224
inertia-weighted perturbation 187 Markov
inflation operator 163 Markov clustering algorithm
inner voltage loop 106, 107 (MCL) 162, 163
instantaneous frequency 82, 122 Markov matrix 158, 162, 163, 172
inter-area modes 7, 41, 91 matrix of attenuation 150
internal combustion (IC) 54 matrix pencil 221
296 Index
maximum penetration level 71, 80, 85, 86, oscillation damping 2, 3, 4, 5, 7, 8, 12,
88, 89, 93, 94, 95, 99 105, 108
MCL, see Markov clustering 159, 162, oscillation-free behavior 61
163, 176
measurement- based approach 28, 169 p
Mei-Sheila algorithm 159 parameter estimation 1
MG, see microgrid 1, 4, 5, 6, 11, 14, 25, 26, parametric uncertainty 5, 135
27, 28, 47, 53, 54, 61, 63, 64, 65, 67, 71, partial least squares
72, 73, 76, 78, 79, 80, 82, 84, 85, 88, 89, partial least squares correlation 179
90, 91, 92, 185, 188, 189, 195, 196, 197, partial least squares regression 181
198, 207 participation factor 11, 137, 138, 215
MGs-penetrated grid 26 percentage of variation (PCTVAR) 278
microgrids (MG) 1, 25, 71, 103, 145 permanent magnet synchronous generators
mid-term dynamics 39 (PMSGs) 108
MIMO, see multi-input multi-output 4 permissible frequency deviation 83, 84
minimization problem 47, 48, 137, 202 phase-locked loop (PLL) 191
minimum inertia 82, 84, 85, 88, 122, phasor measurement units (PMUs) 2,
123, 125 145, 262
modal analysis 5, 36, 41, 61, 126, 213, 222, phasor models 27
230, 254, 285 pilot bus 149, 151, 168, 175
modal decomposition 213, 215, 223, 228 pilot node 149, 150, 151, 152, 166, 168, 178
mode shape PMU, see phasor measurement units 3, 4,
mode shape identification 241 5, 6, 7, 8, 9, 12, 13, 145, 150, 153,
speed-based mode shape 234, 253, 273 154, 155
model point C criterion 82
deterministic 25, 47, 48, 50 point of common coupling 64
stochastic 25, 50, 207 point of interconnection with the system
modular multilevel matrix converter (POIS) 263
(M3C) 108 post-disturbance frequency fluctuation 73
modulation index 59 post-fault stable equilibrium points 46
multi-input multi-output (MIMO) 4 power angle-based stability index 129
multi-MGs (MMGs) 47, 82, 188 balance of plant 119
MUltiple Signal Classification power conversion system (PCS) 119
(MUSIC) 115, 221 power conversion system 119
multiterminal HVDC systems 108 Power references-based VSG control 108
MUSIC, see MUltiple Signal power system modeling 25, 28
classification 115, 116, 221 power system simulator 63
power system stability 1, 4, 14, 26, 53, 71,
n 72, 74, 77, 99, 104, 105, 207
New York New England (NYNE) test power system stabilizer (PSS) 200, 264
system 35, 78, 131, 194 power-electronic interfaced sources 27
primary protection 44
o primary reserve 72
online estimation 13 principal component (PC) 277
optimal placement 103, 104, 113, 114, 121, probability density function (PDF) 49, 135
125, 128, 136 Prony analysis
Index 297
Multisignal Prony analysis 215, 237 spinning reserve 263, 264, 277
Prony polynomial 216, 217, 220 state matrix 90
standard Prony analysis 215 state of charge (SOC) 115, 116, 121, 191
proportional-integral (PI) 55, 284 Static load bank (SLB) 57
pseudo-spectrum 116, 117 static switch 193, 194
pulse function 64 static VAR compensator (SVC) 6, 146, 263
stochastic approaches 46
q stochastic transition matrix 162
quasi-steady-state 132 study zone 27
supervisory control and data acquisition
r (SCADA) 154
rate of change of frequency (RoCoF) 13, swing equation 11, 13, 33, 62, 63, 64, 73,
43, 73, 104, 121, 188, 277 104, 105, 107, 114, 122, 187
reactive control zones 150 synchronism 90
reactive power dispatch 260, 283 synchronizing torque 90
reconstruction error 251 synchronous machine 3, 27, 35, 36, 54, 55,
reduced model 41 69, 90, 104, 105, 128, 264, 278
reliability 11, 26, 46, 71, 72, 74, 147, 206 synchronous-based DG 54, 196
renewable energy sources (RESs) 1, 105, synchronverter 104
146, 212, 259 system inertia 25, 29, 32, 35, 36, 37, 72, 73,
resiliency 48 86, 90, 92, 112, 113, 124, 136, 194, 196,
resonance 12, 193 260, 261, 274, 277
right half plane (RHP) 91 system instability 14, 38, 42, 112, 266
RMSE, see Root Mean Square Error 82 system topology 38, 151
RoCoF, see rate of change of frequency 13,
43, 50, 73, 74, 76, 77, 80, 85, 86, 104, 105, t
108, 109, 110, 113, 114, 121, 122, 123, TCC, see total capital cost 119, 120
126, 129, 188, 189, 191, 200, 277, 281 temporal structure 212, 227
root mean square error (RMSE) 82 three-phase voltage source 59
rotor oscillations 90 tie-line power exchanges 29
rotor-angle stability 27 total capital cost (TCC) 119
transient stability 51, 72, 86, 114, 129, 132,
s 146, 200, 211, 237, 246, 260, 266,
286, 288
sensitivity analysis 6, 90, 92, 134, 135
transient stability assessment (TSA) 86
short-term behavior 42
transition probability matrix 158
similarity matrix 156, 157
transmission system operators
singular value decomposition
(TSOs) 73, 104
(SVD) 179, 215
trial-and-error 72
slowest modes 36
TSOs, see transmission system
small-signal stability assessment
operators 73
(SSSA) 86, 90
turbine dynamics 28
solar photovoltaic 6, 71, 146, 211, 259
turbine-governor model 35
spatial patterns 161, 180, 254
spatial structure 229, 242
spatiotemporal 156, 162, 164, 212, 254 u
spectral graph theory 157 uncertainty analysis 50, 135
298 Index