You are on page 1of 150

VOLTAGE STABILITY ASSESSMENT OF DUBAI POWER GRID

USING A DETAILED LOAD MODEL

by

Salha Ali Al Disi

A Thesis Presented to the Faculty of the


American University of Sharjah
College of Engineering
in Partial Fulfillment
of the Requirements
for the Degree of

Master of Science in
Electrical Engineering

Sharjah, United Arab Emirates

June 2013
2013 Salha Al Disi. All rights reserved.
Approval Signatures

We, the undersigned, approve the Masters Thesis of Salha Ali Al Disi.
Thesis Title: Voltage Stability Assessment of Dubai Power Grid Using A Detailed
Load Model

Signature Date of Signature


___________________________ _______________
Dr. Ahmed Osman-Ahmed
Associate Professor
Department of Electrical Engineering
Thesis Advisor

___________________________ _______________
Dr. Awad Ibrahim Al-Baraasi
Libya's Deputy Prime Minister
Thesis Co-Advisor

___________________________ _______________
Dr. Ayman El-Hag
Associate Professor
Department of Electrical Engineering
Thesis Committee Member

___________________________ _______________
Dr. Amr El-Nady
Associate Professor
Department of Electrical and Computer Engineering
University of Sharjah
Thesis Committee Member

___________________________ _______________
Dr. Mohamed El-Tarhuni
Head
Department of Electrical Engineering

___________________________ _______________
Dr. Hany El Kadi
Associate Dean
College of Engineering

___________________________ _______________
Dr. Leland Blank
Interim Dean
College of Engineering

___________________________ _______________
Dr. Khaled Assaleh
Director of Graduate Studies
Acknowledgments

All the praises and thanks are to Allah the Exalted, the Merciful for bestowing on me
his blessings and providing me with the strength to attain my Master Degree, which
has been a real challenge, yet an interesting experience. I am indebted to many
people, who in one way or another, contributed and extended their valuable support in
the preparation and completion of this thesis.

First and foremost, I would like to express my sincerest gratitude and appreciation to
my thesis advisors, Dr. Ahmed Osman and Dr. Awad Ibrahim, for their persistent
guidance, encouragement, understanding and patience throughout the course of my
thesis, whilst allowing me the room to work in my own way. Without their support
and motivation this thesis would not have been possible.

I also would like to express my warmest thanks and appreciation to my thesis


committee members, Dr. Ayman El-Hag and Dr. Amr Elnady, for their constructive
comments and valuable suggestions. My special thanks and appreciation also go to
Dr. Mohamed El-Tarhuni, Head of Department of Electrical Engineering, for his
unlimited support and assistance during my study at AUS.

My appreciation must be extended to Dubai Electricity and Water Authority (DEWA)


Management for sponsoring me to attain my Master Degree at the American
University of Sharjah (AUS).

I am also grateful to my family, and most especially my husband, for their


unconditional support, care and patience during the toughest times. I would like also
to express my gratitude to all my friends and colleagues at DEWA and AUS for
helping me get through the difficult times, and for all the emotional support they
provided.

I cannot end without emphasizing my sincere and profound gratitude to Dr. Awad
Ibrahim again for believing in me, asserting me to continue my graduate studies and
sparing me lots of his valuable time despite his tremendous responsibilities and
commitments.
To the memory of my Parents
May Allah rest their souls in Heaven
Abstract

Voltage stability problem has become one of the major concerns for power utilities in
recent years. This is due to the exponentially growing demands and the associated
stress on the power transmission resources. Moreover, voltage instability has been
responsible for severe network collapses world-wide and subsequently, the possible
threat of voltage instability is becoming more pronounced in power utilities. Dubai
Power Grid is undergoing similar circumstances. The increased stress on the power
resources in addition to the high proportion of motor driven loads, embedded in Air
Conditioning (AC) appliances, have raised the necessity to assess the voltage stability
of Dubai Power Grid. During large system disturbances, the transmission system
voltage can fall below a critical threshold, resulting in induction motors stalling or
tripping depending on several factors such as motor type, size and control. The
severity increases during peak load conditions, when the system load is dominated by
AC appliances. Recently, Dubai Power Grid had experienced several system
disturbances that were accompanied by small/large voltage variations. These
variations were followed by inadvertent disconnection of load. The existing Dubai
Power Grid load model is not capable of reflecting the actual system behavior
following the experienced disturbances. Having an accurate load model capable of
capturing load behavior during system disturbances is crucial in voltage stability
assessment. This thesis presents a detailed load model for Dubai Power Grid and
validates it against recorded disturbances. The updated load model will be used to
assess voltage stability margin against the increasing use of power transmission
resources, growing demand and associated stress on available and planned active and
reactive power resources.

Search Terms: Voltage Stability Assessment, Load Modeling, Measurement-Based


Approach, Component-Based Approach, Parameter Estimation.

6
Table of Contents

Abstract ........................................................................................................ 6
List of Figures ..................................................................................................... 9
List of Tables .................................................................................................... 15
List of Abbreviations ........................................................................................ 17
List of Symbols.................................................................................................. 18
Chapter 1: Introduction ................................................................................. 20
1.1 Overview ............................................................................................................20
1.2 Motivation ..........................................................................................................21
1.3 Thesis Objectives...............................................................................................22
1.4 Contributions .....................................................................................................23
1.5 Thesis Outline ....................................................................................................23
Chapter 2: Theoretical Background .............................................................. 24
2.1 Power System Stability Definition...................................................................24
2.2 Power System Stability Classification .............................................................24
2.2.1 Rotor Angle Stability..............................................................................26
2.2.2 Frequency Stability .................................................................................27
2.2.3 Voltage Stability .....................................................................................27
2.3 Voltage Stability Assessment ...........................................................................29
2.3.1 Static Voltage Stability Analysis ...........................................................30
2.3.2 Dynamic Voltage Stability Analysis .....................................................34
2.4 Power System Components Modeling for Voltage Stability Assessment.....35
2.4.1 Loads ........................................................................................................35
2.4.2 Generators................................................................................................35
2.4.3 Reactive Power Compensation ..............................................................36
2.4.4 Protection and Controls ..........................................................................36
2.5 Load Modeling For Voltage Stability Assessment .........................................36
2.6 Basic Load Modeling Concepts........................................................................38
2.6.1 Static Load Model...................................................................................39
2.6.2 Dynamic Load Model .............................................................................41
2.6.3 Acquisition of Load Model Parameters.................................................42
2.7 Literature Review on Load Modeling ..............................................................43

7
Chapter 3: Updating Dubai Power Grid Load Model .................................. 57
3.1 Overview of Dubai Power Grid........................................................................57
3.2 Problem Statement.............................................................................................59
3.3 Methodology ......................................................................................................64
3.3.1 Component Based Approach Processes ................................................66
3.3.2 Measurement Based Approach Processes .............................................68
3.3.3 Validation of the Developed Load Model.............................................88
Chapter 4: Voltage Stability Assessment of Dubai Power Grid ................... 94
4.1 Overview of Dubai Power Grid Planning Standards ......................................94
4.2 Voltage Stability Assessment Methodology....................................................96
4.2.1 Software Tool ..........................................................................................96
4.2.2 Load Model Representation ...................................................................96
4.2.3 Study Considerations and Scenarios......................................................97
4.2.4 Steady State Voltage Analysis Results................................................100
4.2.5 Dynamic Voltage Analysis Results .....................................................114
Chapter 5: Conclusions, Recommendations and Future Work .................. 120
5.1 Conclusions......................................................................................................120
5.2 Recommendations ...........................................................................................122
5.3 Future Work .....................................................................................................123
References .................................................................................................... 125
Appendix A .................................................................................................... 129
Appendix B .................................................................................................... 130
Appendix C .................................................................................................... 138
Vita .................................................................................................... 150

8
List of Figures

Figure 2.1: Classification of Power System Stability .................................................. 26

Figure 2.2: Voltage Stability Phenomena and Time Responses .................................. 28

Figure 2.3: A Simple Two Bus System ....................................................................... 31

Figure 2.4: PV Curves for Some Power Factor Values ............................................... 32

Figure 2.5: Different VQ Curves and Critical Operating Points ................................. 33

Figure 2.6: Induction Motor Equivalent Circuit .......................................................... 41

Figure 2.7: Component-Based Modeling Approach .................................................... 43

Figure 3.1: Dubai Power Grid Geographical Map ....................................................... 57

Figure 3.2: Electricity Installed Capacity & Peak Demand (2002-2012) .................... 58

Figure 3.3: Percentages of Energy Consumption for different Consumer Categories


by Year 2012 ............................................................................................ 59

Figure 3.4: Dubai Power Grid Load Composition (Existing Load Model) ................. 60

Figure 3.5: Recorded Voltage at MUSH 132 kV bus during MUSH Disturbance ...... 61

Figure 3.6: Frequency, System Requirement and Interchange Flow Trend during
MUSH Disturbance .................................................................................... 62

Figure 3.7: Zoomed in part of the Frequency Trend Curve during MUSH
Disturbance ................................................................................................ 62

Figure 3.8: Recorded vs. Simulated Voltage Trend at MUSH 132 kV Bus MUSH
Disturbance ............................................................................................... 63

Figure 3.9: Simulated Total System Load during MUSH Disturbance


(Existing Load Model) ............................................................................. 63

Figure 3.10 : Illustration of the Developed Load Modeling Methodology.................. 65

Figure 3.11: Result of Implementing Component Based Approach on Dubai Power


Grid ........................................................................................................... 68

Figure 3.12: Equivalent Circuit of the Selected Aggregate Load Model Structure ..... 69

Figure 3.13: Parameter Estimation Procedures ............................................................ 74

Figure 3.14: Measured Instantaneous Voltage and Current Signals


Incident No.1 ............................................................................................ 75

9
Figure 3.15: Calculated Line Positive Sequence Voltage and Current
Incident No.1 ............................................................................................ 75

Figure 3.16: Measured Instantaneous Voltage and Current Signals


Incident No.2 ............................................................................................ 76

Figure 3.17: Calculated Line Positive Sequence Voltage and Current


Incident No.2 ............................................................................................ 76

Figure 3.18: Measured Instantaneous Voltage and Current Signals


Incident No.3 ............................................................................................ 77

Figure 3.19: Calculated Line Positive Sequence Voltage and Current


Incident No.3 ............................................................................................ 77

Figure 3.20: Measured Instantaneous Voltage and Current Signals


Incident No.4 ............................................................................................ 78

Figure 3.21: Calculated Line Positive Sequence Voltage and Current


Incident No.3 ............................................................................................ 78

Figure 3.22: Measured and Estimated Active and Reactive Power Incident No.1 ... 79

Figure 3.23: Measured and Estimated Active and Reactive Power Incident No.2 ... 80

Figure 3.24: Measured and Estimated Active and Reactive Power Incident No.3 ... 80

Figure 3.25: Measured and Estimated Active and Reactive Power Incident No.4 ... 80

Figure 3.26: Aggregated Motor Currents..................................................................... 81

Figure 3.27: Aggregated Motor Torque ....................................................................... 82

Figure 3.28: Aggregated Motor Speed......................................................................... 82

Figure 3.29: Measured Instantaneous Voltage and Current Signals ............................ 83

Figure 3.30: Calculated Line Positive Sequence Voltage and Current ........................ 83

Figure 3.31: Measured and Simulated Active and Reactive Power............................. 84

Figure 3.32: Fast Tripping Pattern-Ilustration-1 .......................................................... 85

Figure 3.33: Fast Tripping Pattern-Ilustration-2 .......................................................... 85

Figure 3.34: Different Tripping Schemes (Fast Tripping Pattern) .............................. 86

Figure 3.35: Slow Tripping Pattern-Ilustration-1 ........................................................ 87

Figure 3.36: Slow Tripping Pattern-Ilustration-2 ........................................................ 87

10
Figure 3.37: Extracted Slow Tripping Pattern related to the Triggering Voltage
Dip ............................................................................................................ 88

Figure 3.38: Recorded vs. Simulated Voltage Trend at MUSH 132 kV bus
MUSH Disturbance .................................................................................. 90

Figure 3.39: Recorded vs. Simulated Voltage Trend at MUSH 132 kV bus
MUSH Disturbance (New and Old Load Model) .................................... 90

Figure 3.40: Simulated Total System Load during MUSH Disturbance


(New Load Model) ................................................................................... 91

Figure 3.41: Recorded vs. Simulated Voltage Trend at MUSH 132 kV bus
WRSN Disturbance .................................................................................. 92

Figure 3.42: Recorded vs. Simulated Voltage Trend at BKRA 400 kV bus
WRSN Disturbance .................................................................................. 92

Figure 3.43: Simulated Total System Load during WRSN Disturbance ..................... 93

Figure 4.1: Typical Substation Layout a 400/132 kV Substation ................................ 98

Figure 4.2: Example of N-1 Contingencies ................................................................. 99

Figure 4.3: Example of N-2 Contingencies ................................................................. 99

Figure 4.4: Example of N-3 Contingencies ................................................................. 99

Figure 4.5: Calculating Active Power Transfer Margin from PV Curves ................. 100

Figure 4.6: Calculating Reactive Power Reserve Margin from VQ Curves .............. 101

Figure 4.7: PV Curves for NHDA 400/132 kV Substation Base Case and All
Contingencies ......................................................................................... 102

Figure 4.8: PV Curves for NHDA 400/132 kV Substation


All N-1 Contingencies ............................................................................ 102

Figure 4.9: PV Curves for NHDA 400/132 kV Substation


All N-2 Contingencies ............................................................................ 103

Figure 4.10: PV Curves for NHDA 400/132 kV Substation


All N-3 Contingencies ............................................................................ 103

Figure 4.11: PV Curves for NHDA 400/132 kV Substation


Worst Contingencies .............................................................................. 104

Figure 4.12: PV Curves for MUSH 400/132 kV Substation


Worst Contingencies .............................................................................. 106

11
Figure 4.13: PV Curves for CARX 400/132 kV Substation
Worst Contingencies .............................................................................. 106

Figure 4.14: PV Curves for MBCH 400/132 kV Substation


Worst Contingencies .............................................................................. 107

Figure 4.15: QV Curves for NHDA 400/132 kV Substation


Base Case and All Contingencies ........................................................... 108

Figure 4.16: QV Curves for NHDA 400/132 kV Substation


All N-1 Contingencies ............................................................................ 109

Figure 4.17: QV Curves for NHDA 400/132 kV Substation


All N-2 Contingencies ............................................................................ 109

Figure 4.18: QV Curves for NHDA 400/132 kV Substation


All N-3 Contingencies ............................................................................ 109

Figure 4.19: QV Curves for NHDA 400/132 kV Substation


Worst Contingencies .............................................................................. 110

Figure 4.20: VQ Curves for MUSH 400/132 kV Substation


Worst Contingencies .............................................................................. 112

Figure 4.21: VQ Curves for CARX 400/132 kV Substation


Worst Contingencies .............................................................................. 112

Figure 4.22: VQ Curves for MBCH 400/132 kV Substation


Worst Contingencies .............................................................................. 113

Figure 4.23: Illustration of Different Fault Locations ............................................... 115

Figure 4.24: Voltage Trends at NHDA 400 kV Bus for the Worst Contingencies
(Normal Fault Clearance Time) ............................................................. 116

Figure 4.25: Voltage Trends at NHDA 132 kV Bus for the Worst Contingencies
(Normal Fault Clearance Time) ............................................................. 117

Figure 4.26: Voltage Trends at NHDA 400 kV Bus for the Worst Contingencies
(Breaker Failure- Backup Protection Time) ........................................... 117

Figure 4.27: Voltage Trends at NHDA 132 kV Bus for the Worst Contingencies
(Breaker Failure- Backup Protection Time) ........................................... 118

Figure A.1: Simulink Model For Load Model Parameter Estimation ....................... 129

Figure B.1: Dubai Power Grid Planned 400 kV Network Topology Year 2014 .... 130

Figure B.2: Dubai Power Grid Geogrphical Map ...................................................... 131

12
Figure C.1: PV Curves for MUSH 400/132 kV Substation
Base Case and All Contingencies ................................................................ 138

Figure C.2: PV Curves for MUSH 400/132 kV Substation


All N-1 Contingencies ................................................................................... 138

Figure C.3: PV Curves for MUSH 400/132 kV Substation


All N-2 Contingencies ................................................................................... 138

Figure C.4: PV Curves for MUSH 400/132 kV Substation


All N-3 Contingencies ................................................................................... 139

Figure C.5: QV Curves for MUSH 400/132 kV Substation


Base Case and All Contingencies ................................................................ 140

Figure C.6: QV Curves for MUSH 400/132 kV Substation


All N-1 Contingencies ................................................................................... 140

Figure C.7: QV Curves for MUSH 400/132 kV Substation


All N-2 Contingencies ................................................................................... 140

Figure C.8: QV Curves for MUSH 400/132 kV Substation


All N-3 Contingencies ................................................................................... 141

Figure C.9: PV Curves for CARX 400/132 kV Substation


Base Case and All Contingencies ................................................................ 142

Figure C.10: PV Curves for CARX 400/132 kV Substation


All N-1 Contingencies ................................................................................... 142

Figure C.11: PV Curves for CARX 400/132 kV Substation


All N-2 Contingencies ................................................................................... 142

Figure C.12: PV Curves for CARX 400/132 kV Substation


All N-3 Contingencies ................................................................................... 143

Figure C.13: QV Curves for CARX 400/132 kV Substation


Base Case and All Contingencies ................................................................ 144

Figure C.14: QV Curves for CARX 400/132 kV Substation


All N-1 Contingencies ................................................................................... 144

Figure C.15: QV Curves for CARX 400/132 kV Substation


All N-2 Contingencies ................................................................................... 144

Figure C.16: QV Curves for CARX 400/132 kV Substation


All N-3 Contingencies ................................................................................... 145

Figure C.17: PV Curves for MBCH 400/132 kV Substation


Base Case and All Contingencies ................................................................ 146

13
Figure C.18: PV Curves for MBCH 400/132 kV Substation
All N-1 Contingencies ................................................................................... 146

Figure C.19: PV Curves for MBCH 400/132 kV Substation


All N-2 Contingencies ................................................................................... 146

Figure C.20: PV Curves for MBCH 400/132 kV Substation


All N-3 Contingencies ................................................................................... 147

Figure C.21: QV Curves for MBCH 400/132 kV Substation


Base Case and All Contingencies ................................................................ 148

Figure C.22: QV Curves for MBCH 400/132 kV Substation


All N-1 Contingencies ................................................................................... 148

Figure C.23: QV Curves for MBCH 400/132 kV Substation


All N-2 Contingencies ................................................................................... 148

Figure C.24: QV Curves for MBCH 400/132 kV Substation


All N-3 Contingencies ................................................................................... 149

14
List of Tables

Table 3.1: Substations and Lines Statistics for Dubai Power Grid (2002-2012) ......... 58

Table 3.2: Load Tripping Rules for Motor Load (Existing Load Model) ................... 61

Table 3.3: Sample of Load Classes/Mixes in Dubai Power Grid System Level ...... 66

Table 3.4: Typical Load Composition for Different Load Classes .............................. 67

Table 3.5: Modified Load Composition for Dubai Load Composition ....................... 67

Table 3.6: Parameters to be identified for the Aggregate Load Model ....................... 73

Table 3.7: Estimated Load Composition and ZIP Load Model Parameters for each
Incident ...................................................................................................... 81

Table 3.8: Estimated Unified Aggregate Motor Load Parameters .............................. 81

Table 3.9: Extracted Load Tripping Scheme (Fast Tripping Pattern) ......................... 86

Table 4.1: Steady State Voltage Levels for 400 kV and 132 kV levels....................... 95

Table 4.2: Maximum Active Power Transfer Margin for NHDA 400/132 kV
Substation Base Case and All Contingencies ....................................... 105

Table 4.3: Maximum Active Power Transfer Margin for NHDA, MUSH, CARX and
MBCH 400/132 kV Substation Base Case and Worst Contingencies .. 107

Table 4.4: Available Reactive Power Reserve Margin for NHDA 400/132 kV
Substation Base Case and All Contingencies ....................................... 111

Table 4.5: Available Reactive Power Reserve Margin for NHDA, MUSH, CARX and
MBCH 400/132 kV Substation Base Case and All Contingencies....... 113

Table 4.6: Details of Simulated Disturbances............................................................ 114


Table B.1: Comprehensive Contingency List for NHDA 400/132 kV Substation .... 132

Table B.2: Comprehensive Contingency List for MUSH 400/132 kV Substation .... 134

Table B.3: Comprehensive Contingency List for CARX 400/132 kV Substation .... 135

Table B.4: Comprehensive Contingency List for MBCH 400/132 kV Substation.... 136

Table B.5: WECC Voltage Stability Criteria ............................................................. 137

Table C.1: Maximum Power Transfer Margin for MUSH 400/132 kV Substation
Base Case and All Contingencies ............................................................ 139

Table C.2: Available Reactive Power Margin for MUSH 400/132 kV Substation
Base Case and All Contingencies ............................................................ 141

15
Table C.3: Maximum Power Transfer Margin for CARX 400/132 kV Substation
Base Case and All Contingencies ............................................................ 143

Table C.4: Available Reactive Power Margin for CARX 400/132 kV Substation
Base Case and All Contingencies ............................................................ 145

Table C.5: Maximum Power Transfer Margin for MBCH 400/132 kV Substation
Base Case and All Contingencies ............................................................ 147

Table C.6: Available Reactive Power Margin for MBCH 400/132 kV Substation
Base Case and All Contingencies ............................................................ 149

16
List of Abbreviations

AC Air Conditioner
AGC Automatic Generation Control
AVR Automatic Voltage Regulator
BKRA Bukidra
CARX Car Complex
CDSM Composite Dynamic Static Model
DCP District Cooling/Chiller Plant
DEWA Dubai Electricity and Water Authority
DFR Digital Fault Recorder
EPRI Electric Power Research Institute
EUVLS Embedded Under Voltage Load Shedding
GA Genetic Algorithm
GABPE Genetic Algorithm Based Parameter Estimation
HILP High Impact Low Probability
HVDC High Voltage Direct Current
LTC Load Tap Changer
LV Low Voltage
LV AC Low Voltage Alternating Current
MBCH Mamzer Beach
MUSH Mushrif
NERC North American Electric Reliability
NHDA Nahda
PSO Particle Swarm Optimization
PSS/E Power System Simulator for Engineers
PTI Power Technologies Incorporation
PV Active Power versus Voltage
RMS Root Mean Square
SCADA supervisory control and data acquisition
SVC Static Var Compensator
SVS Static Var System
ULTC Under Load Tap Changer
VQ Voltage versus Reactive Power
WECC Western Electricity Coordination Council
WRSN Warsan
ZIP Constant Impedance- Constant Current- Constant Power

17
List of Symbols

A Torque Coefficient Proportional to the Square of the Speed


B Torque Coefficient Proportional to the Speed
C Constant Torque Coefficient
1 Percentages of Active Power Consumed by Static Load
2 Percentages of Reactive Power Consumed by Static Load
1 Percentage of Active Power Consumed by Dynamic Load
2 Percentage of Reactive Power Consumed by Dynamic Load
E Electromotive Force
I Current
I+ Positive Sequence Current
ids Direct Access Stator Current
iqs Quadratic Access Stator Current
P Active Power
p.u. Per Unit
PIM Active Power Consumed by Induction Motor
PL Active Power Consumed by Aggregate Load
PZIP Active Power Consumed by Static Load
Q Reactive Power
Power Angle
QIM Reactive Power Consumed by Induction Motor
QL Reactive Power Consumed by Aggregate Load
QZIP Reactive Power Consumed by Static Load
R Resistance
Rr Rotor Resistance
Rs Stator Resistance
Te Electrical Output Torque
TL Load (Mechanical) Torque
To Initial Load (Mechanical) Torque
V Voltage
V+ Positive Sequence Voltage
vdr Direct Access Rotor Voltage
vds Direct Access Stator Voltage
vqr Quadratic Access Rotor Voltage
vqs Quadratic Access Stator Voltage
b Motor Angular Electrical Base Frequency
r Rotor Angular Electrical Speed
e Stator Angular Electrical Frequency

18
X Reactance
Xlr Rotor Leakage Reactance
Xls Stator Leakage Reactance
Xm Magnetizing Leakage Reactance
dr Direct Access Rotor Flux Linkage
ds Direct Access Stator Flux Linkage
md Direct Access Magnetizing Flux Linkage
qr Quadratic Access Rotor Flux Linkage
qs Quadratic Access Stator Flux Linkage
mq Quadratic Access Magnetizing Flux Linkage

19
Chapter 1: Introduction

1.1 Overview

Currently, most of the power systems around the world are being operated
under much more stressed conditions than were usual in the past. Environmental
pressure on transmission expansions, exponentially growing demands and penetration
of new types of loads (such as inverter-based appliances) at demand side are some of
the responsible factors for these stressed conditions. Under such stressed operational
conditions, a power system may exhibit instability behaviors that are characterized by
either slow or sudden voltage drops, i.e. voltage instability. Under certain conditions,
voltage instability may escalate to a form of voltage collapse which intimidates
system security. This was evidenced in several network collapses and blackouts
world-wide. Subsequently, voltage stability has become a major concern for power
system utilities [1].

Dubai Power Grid, being planned and operated by Dubai Electricity and Water
Authority (DEWA), is undergoing similar circumstances that made voltage stability a
critical issue. The growing demands and associated stress on the available and
planned system resources as well as the relatively limited geographical area of the city
had resulted in allocating most generation stations at one side of the city, hence,
feeding the load centers through long transmission circuits. Additionally, the
electrical load of Dubai has a particular nature; a significant amount of the supplied
load, especially in summer, is dominated by induction motor driven loads, specifically
Air Conditioning (AC) appliances. During system disturbances, such as faults,
transmission voltages may drop below certain thresholds resulting in either motor
stalling or tripping. Excessive motor tripping or stalling may result in either voltage
collapse or cascaded generator tripping, particularly if the reactive power
compensation facilities of the power system are not adequately sized. Therefore,
proper modeling of power system loads, with a focus on induction motor driven loads,
is essential for voltage stability assessment of a power system.

20
This thesis shall comprehensively assess Dubai Power Grid voltage stability margin
against the increasing use of power transmission resources, growing demand and
associated stress on available and planned active and reactive power resources; with a
special attention to the special load nature of the city.

1.2 Motivation

Dubai Power Grid had experienced several system disturbances that were
accompanied by small/large voltage variations. These voltage variations were
followed by unplanned disconnection of load. The largest amount of load loss was
37% of the total system load encountered following a single phase to ground fault.
This indicates the large proportion of the motor driven load represented mainly by AC
appliances. In general, motor driven loads such as AC appliances may respond to a
voltage dip in different ways depending on several factors such as motor type, size
and control. For example, a window AC unit typically has simple control; hence,
following a voltage dip the electrical torque may decrease below the mechanical load
torque causing the motor to stall, and the motor will eventually be disconnected by
thermal overload protection. On the other hand, a split system AC unit has contactors
that may respond to a voltage dip by disconnecting the motor before it stalls. The
consequences of excessive loss of load or excessive stalling may intimidate voltage
stability and hence system security. Such short-term voltage instability scenarios need
be studied in order to examine the dynamic interactions between the system loads and
the power grid.

Though, all the previous voltage stability studies that were conducted for
Dubai Power Grid show that voltage stability is maintained for all reasonable
contingencies and even for most contingencies beyond the normal design criteria. The
results of these studies are not in line with the actual system behavior following the
experienced disturbances. The reason behind these discrepancies is that all the
previous voltage stability studies were based on a tentative dynamic load model which
has three basic (load versus voltage) components: constant power, constant current
and constant impedance. These components of load model, while useful in the
absence of better information, do not always give an adequate characterization of a
systems load versus voltage characteristic. Indeed, understanding the load nature is

21
very important for voltage stability assessment. The potential benefits of improved
load representation fall into the following categories [2]:

If the existing load representation produces overly-pessimistic results:

- At planning level, the benefits of improved modeling will be in avoiding


premature investments by deferring or avoiding the expense of unnecessary
system modifications or equipment additions.
- At operational level, the benefits will be in increasing power transfer limits,
and more flexibility in operating the system with resulting economic
savings.

If the existing load representation produces overly-optimistic results:

- At planning level, the benefit of improved modeling will be in avoiding


system inadequacies that may result in costly operating limitations.
- At operational level, the benefit may be in preventing system emergencies
resulting from overly-optimistic operational limits

Based on the above, it is becoming necessary to update Dubai Power Grid


Load Model to make it capable of representing, with reasonable accuracy, the load
behavior when subjected to actual voltage variations and hence assess the voltage
stability of Dubai Power Grid based on this updated load model.

1.3 Thesis Objectives

The objectives of this thesis are:

To provide a methodology for developing aggregate load models for utility


power systems using real system data and measurements and test it on a real
power system.
To update Dubai Power Grid load model, using the developed methodology,
to make it capable of capturing load behavior during system disturbances and
enhance the accuracy of voltage stability studies.

22
To evaluate the voltage performance of Dubai Power Grid against the
increasing use of transmission system resources, growing demand and
associated stress on available active and reactive power resources.

1.4 Contributions

The main contributions of this thesis are:

Developing a new hybrid methodology for developing aggregate load models


consisting of a combination of component based approach and measurement
based approach using real system data and measurements.
Developing the load model using small voltage variation disturbances and
validating it against large voltage variation disturbances.
Analysis of load characteristic during small and large voltage variations based
on field measurements, identification of load model parameters and embedded
load self-disconnection of induction motor loads (represented mainly by Air
Conditioning AC Appliances).

1.5 Thesis Outline

Chapter 1 of this thesis gives a general overview of the research topic, the
motivation and the objectives for carrying this research. Chapter 2 presents the
theoretical background of voltage stability problem, analysis techniques, load
modeling and its importance in voltage stability assessment, followed by a literature
review on the previous work conducted in the area. Chapter 3 starts with overview of
Dubai Power Grid and the problem statement, followed by the developed load
modeling methodology and its implementation for updating Dubai Power Grid Load
Model. Chapter 4 presents the results of voltage stability assessment of Dubai Power
Grid based on the developed load model. The conclusion of this research and
recommendations for future work are outlined in Chapter 5.

23
Chapter 2: Theoretical Background

2.1 Power System Stability Definition

A typical modern power system is a high-order non-linear multivariable


dynamic system. The dynamic behavior of a modern power system is influenced by
its various components and their characteristics and response rates. Power system
stability has fundamental mathematical substructures that are comparable to the
stability of any other dynamic system. Accurate definitions of stability are available in
literatures that deal with the mathematical foundations of dynamic system stability.
According to IEEE/CIGRE Joint Task Force on Stability Terms and Definitions,
Power System Stability is defined as [3]:

Power system stability is the ability of an electric power system, for a given
initial operating condition, to regain a state of operating equilibrium after
being subjected to a physical disturbance, with most system variables bounded
so that practically the entire system remains intact.

Power system components operate in a constantly changing environment


(dynamic behavior). As inferred from the above definition, the stability of the power
system, when exposed to a disturbance, depends on the system initial operating
condition as well as the nature of the disturbance [3].

2.2 Power System Stability Classification

Although power system stability is a single problem in principle, the various


forms of instabilities that a power system may experience cannot be appropriately
understood and effectively dealt with by treating it as such. Overall, stability is a
condition of balance between opposing forces. Different sets of opposing forces may
experience sustained imbalance depending on the system operating condition,
network topology, and the form of disturbance. Accordingly, power systems may
experience different types of instability. Therefore, it is essential to classify power
system stability phenomena to help power system engineers to analyze these
instabilities [1].

Previously, transient stability (large disturbance rotor angle stability) has been
the main stability concern for most power systems, and hence has been the focus of

24
industrys attention. Different forms of system instability have emerged as a
consequence of power systems evolvement aroused from continuing growth in
interconnections, use of new technologies and controls, and the increased operation in
highly stressed conditions. For example, voltage stability, frequency stability and
inter-area oscillations have become greater concerns than in the past. A clear
understanding of different types of instability and how they are interrelated is
necessary for the satisfactory design and operation of power systems [3].

The classification of power system stability as proposed by per IEEE/CIGRE


Joint Task Force on Stability Terms and Definitions is based on the following
considerations [4]:

The physical nature of the resulting form of instability as indicated by the


main system variable in which instability can be observed.
The size of the disturbance which affects the method of calculation and
prediction of stability.
The power system components and the time duration that must be taken into
consideration in order to assess stability.

In other literatures, power system stability is classified according to the


driving force of instability into generator-driven (rotor angle and frequency stability)
and load-driven (voltage stability). However, the terms (Generator-Driven) and
(Load- Driven) does not exclude the contribution of other system components to the
instability mechanism [1]. Power system stability can be classified according to
physical nature into three categories: rotor angle stability, frequency stability and
voltage stability. Figure 2.1 shows a general classification of the power system
stability problem [3].

25
Figure 2.1: Classification of Power System Stability [3]

2.2.1 Rotor Angle Stability

Rotor angle stability can be defined as the ability of synchronous machines of


an interconnected power system to stay in synchronism after being exposed to a
disturbance. It depends on the capability of the synchronous machines in the system to
maintain and/or restore balance between electromagnetic torque and mechanical
torque. The resulting instability occurs in the form of growing angular swings of some
generator units and subsequently to their loss of synchronism with other generator
units [3]. Rotor angle instability may be encountered in one of two forms [4]:

1) Undamped mechanical oscillations (lack of damping torque) initiated by


small system disturbances, and thus called steady-state or small signal
stability.
2) Monotonic rotor acceleration leading to loss of synchronism (lack of
synchronizing torque) initiated by large system disturbances and thus called
transient stability.

The time frame of both forms of rotor angle instability is in the range of few
seconds, where automatic voltage regulators (AVRs), excitation systems, turbine and
governor dynamics act, hence, the categorization of both rotor angle and the transient
stability as short term phenomena [1].

26
2.2.2 Frequency Stability

Frequency stability is the ability of a power system to maintain stable


frequency following a severe system incident. It depends on the ability to maintain
and/or restore balance between system generation and load, with minimum
inadvertent loss of load. The resulting instability occurs in the form of sustained
frequency swings and subsequently to cascaded tripping of generating units and/or
loads. The time frame for frequency instability ranges from few seconds (short term
phenomena corresponding to under-frequency load shedding schemes) to several
minutes (long term phenomena corresponding to prime mover dynamics) [3].

2.2.3 Voltage Stability

Voltage stability is the ability of a power system to maintain steady voltages at


all its buses after being exposed to a disturbance from a given initial operating
condition. It depends on the ability to maintain and/or restore balance between load
supply from the power system and load demand. Instability that may result occurs in
the form of a progressive drop or rise of voltages of some or all system buses [3].
Voltage stability covers a wide range of phenomena. Therefore, voltage stability
means different phenomenon to different engineers. It is a fast phenomenon for
engineers involved with induction motors, air conditioning loads or HVDC links. It is
however a slow phenomenon for engineers involved with mechanical tap changers
and other manual operator-initiated actions. Figure 2.2 shows that many system
components and controls play a role in voltage stability [5]. In various literatures,
voltage stability has been classified into short and long-term. Short-term voltage
stability corresponds to a time frame of about a few seconds, and is characterized by
the dynamics of components such as induction motors, static var compensators
(SVCs) and excitation of synchronous generators. On the other hand, long-term
voltage stability corresponds to slower time frames, around several minutes, and is
characterized by dynamic recovery of the load due to the action of on-load tap
changers, current limiter control in generators, corrective control actions such as
reactive compensation and load shedding, operator control actions, etc. [5].

27
Figure 2.2: Voltage Stability Phenomena and Time Responses [5]

Conditions causing the voltage instability may include one or more of the following
situations [4]:

The power flow in the transmission lines is too high.


The voltage/reactive power control resources are too far from the load centers.
The source voltages (at generation side) are too low.
The reactive power compensations facilities are insufficient.

Voltage instability may result in the loss of load in an area, or tripping of


transmission lines and other elements by their protective systems leading to cascading
outages. Loss of synchronism of some generators may also occur as a consequence of
these outages or due to other specific operating conditions that violate field current
limit [1].

The driving force for voltage instability is usually the loads. Following a
disturbance, power consumed by the loads tends to be restored by the action of motor
slip adjustment, distribution voltage regulators, tap-changing transformers, and
thermostats. Restored loads may increase the stress on the transmission network by

28
increasing the consumption of reactive power and causing additional voltage
reduction. Voltage instability may occur when load dynamics attempt to restore power
consumption beyond the capability of the transmission network and the connected
generation [1] and [6].

While the most common form of voltage instability is the progressive drop of
bus voltages, the jeopardy of overvoltage instability has been experienced by some
systems. It is caused by a capacitive behavior of the network (EHV transmission lines
operating below surge impedance loading) as well as by underexcitation limiters
preventing generators and/or synchronous compensators from absorbing the excess
reactive power. In this case, the instability is associated with the inability of the
generation and transmission system to operate below some load level. Transformer
tap changers may cause long-term voltage instability while attempting to restore this
load power [3].

2.3 Voltage Stability Assessment

Voltage stability assessment for a given power system involves the


examination of two aspects [4]:

Proximity to voltage instability: how close is the system to voltage instability?


Distance to voltage instability can be measured in terms of physical quantities
such as system load level, active power flow through a critical interface and/or
reactive power reserve. However, the most appropriate measure for any given
situation is system specific, as it depends on many aspects including planning
versus operating decisions. Consideration must be given to potential
contingencies such as line outages, loss of a generation unit or a reactive
power source, etc.
Mechanism of voltage instability: how and why does voltage instability
occur? What are the key factors contributing to voltage instability? What are
the voltage-weak areas? What are the most effective measures for improving
voltage stability?

29
Voltage stability assessment tools fall into two approaches: static and
dynamic. Static voltage stability analysis, which is based on power flow solutions, is
performed to identify the weak regions in terms of reactive power deficiency of the
system and determine the critical contingencies and voltage stability margins for
various power transfers within the power system. On the other hand, dynamic voltage
stability analysis, which is based on conducting time-domain simulations, is
performed to assess the ability of the power system, with a significant share of
rotating load, to operate satisfactorily following disturbances [7]. A comprehensive
voltage stability assessment would include both steady state and dynamic analyses
techniques as suggested in [8].

2.3.1 Static Voltage Stability Analysis

Static Voltage Stability Analysis (also called steady state voltage stability
analysis) is often used to analyze slower form of voltage instability making use of
power flow simulation as a primary study method. It is commonly tackled by plotting
PV and VQ curves. These two methods determine steady state loadability limits,
which are associated to voltages stability [5]. PV and VQ curves are useful for the
following purposes [9]:

Defining voltage collapse point(s) in the power system network.


Determining the maximum power transfer between different network buses
before voltage collapse point.
Sizing the reactive power compensation devices required at relevant buses to
prevent voltage collapse.
Assessing the impact of generator, loads and reactive power compensation
devices on the network.

In order to clearly understand PV and VQ curves, consider the two bus system as
shown Figure 2.3.

30
Figure 2.3: A Simple Two Bus System

The active and reactive power consumed by load at the receiving end bus is given by
(2.1) and (2.2), respectively [1]:


= (2.1)

2
= (2.2)

By the help of the trigonometric identity (sin2 + cos2 = 1) we get:

2
2 2
2
+ + = 0 (2.3)

The voltage at the receiving end bus can be written as:

2 2
= 2 2 2 (2.4)
2 4

Sample PV curves corresponding to different PF values are shown in Figure 2.4.

31
Figure 2.4: PV Curves for Some Power Factor Values

The upper part of the curves represents the stable region with / < 0,
while the lower part of the curves represents the unstable region with / > 0.
Voltage at the point of maximum loading margin, often referred to as nose point, is
known as Critical Voltage. This point is popularly referred to as collapse point [1].

Similarly, equation (2.2) can be rearranged as:

2 cos + = 0 (2.5)

Therefore,

2
= (2.6)

The voltage stability limit is reached at / = 0, this is the critical


operating point of the system [1]. The different VQ curves and critical operating
points are shown in Figure 2.5.

32
Figure 2.5: Different VQ Curves and Critical Operating Points

The right hand side of the curves represents the stable region with
dQ/dV > 0, while the left hand site of the curves represents the unstable region
with dQ/dV < 0. The bottom of the curves, often referred to as knee point,
represents the voltage stability limit [1].

2.3.1.1 PV Curves

PV curves are useful for conceptual voltage stability analysis and for studying
small or radial systems. This method is also used for large and meshed systems where
P is the total load in an area and V is the voltage at a critical or representative bus [5].
In principle, PV curve is a representation of voltage change as a result of increased
active power transfer between two systems (or subsystems). Tracing PV curves
requires a parametric study involving a series of power flows while monitoring the
changes in one set of power flow variables with respect to another. As power transfer
is increased in steps, voltage decreases at some buses on or near the transfer path. The
transfer capacity where voltage reaches a low value criterion (typically 90 to 95% of
the rated voltage) is the low voltage transfer limit. Transfer can continue to increase
until the solution identifies the proximity to the voltage instability, which is a nose
point on the PV curve where the voltage drops steeply in response of an increase in

33
the transfer power flow. Load flow solution will not converge beyond this limit,
indicating a voltage collapse transfer limit [10].

2.3.1.2 VQ Curves

VQ curves are used to determine the reactive power injection required at a bus
in order to control the bus voltage or set it to a required value. Like the PV curves, the
VQ curves are obtained through a series of load flow calculations. Starting with the
existing reactive loading at a bus, the voltage at the bus can be computed for a series
of power flows as the reactive load is increased in steps, until the power flow
experiences convergence difficulties indicating the proximity of a voltage collapse.
The bottom of the VQ curve, where the change of reactive power dQ, with respect to
voltage dV (dQ/dV) is equal to zero, represents the voltage stability limit. Since all
reactive power compensation devices are designed to operate satisfactorily when an
increase in Q results in an increase in V, the operation on the right side of the VQ
curve is stable, whereas the operation on the left side is unstable. The bottom of the
VQ curves defines also the minimum reactive power requirement for the stable
operation. Hence, the VQ curve can be used to examine the type and size of
compensation needed to provide voltage stability. Practically, this can be performed
by superimposing the VQ characteristic curves of the compensator devices on that of
the system [10].

2.3.2 Dynamic Voltage Stability Analysis

Dynamic analysis (also referred to as time-domain analysis) can capture the


evolvement of the instability process by simulating the transient response and timing
of control actions. This is done by solving a set of differential and algebraic equations
representing the system under study. Power systems networks typically include a
large number of dynamic and static components, where each individual component
may need several differential and algebraic equations to be represented. Accordingly,
the total number of differential and algebraic equations of a real power system can be
relatively large [11].

Voltage stability is a dynamic problem by nature; hence, dynamic voltage


stability analysis provides the most accurate response of the actual dynamics of

34
voltage instability when the appropriate modeling is included. Accurate dynamic
simulation reveals system trajectory after a disturbance; hence it is an essential tool
for post-disturbance analysis and the coordination of protection and control devices.
The dynamic voltage stability analysis is normally conducted by simulating the events
and chronology leading to voltage instability [1].

In contrast to static analysis in which equilibrium points of a PV/VQ curves


are time-independent, dynamic voltage stability analysis defines the time-dependent
voltage performance of the system. This method reveals the transient and/or the
longer-term voltage stability of a power system under study [9].

2.4 Power System Components Modeling for Voltage Stability Assessment

The accuracy of voltage stability assessment of a power system is associated


with the accuracy of modeling its components. Power system components need to be
carefully represented in the power system model to simulate their actual behavior
following system disturbances. Therefore, transformer Under Load Tap Changer
(ULTC) action, reactive power compensation at load side and voltage regulators in
the sub-transmission system must be carefully considered. The following are
descriptions of major power system elements that have a significant impact on voltage
stability, hence, need to be carefully represented in power system model [4]:

2.4.1 Loads

Voltage instability is a load driven disturbance; therefore, load characteristics


could be critical to voltage stability analysis. It is important to account for voltage
and frequency dependence of loads in the power system model, thus it is important to
carefully represent induction motors specifically. The appropriate representation of
load characteristics at low voltages is also essential [4].

2.4.2 Generators

Power system disturbances leading to voltage instability often involve


generation-load imbalances. This causes redistribution of power flow and reactive
losses. It is essential to represent how generation units respond to such disturbances
in the power system model [5]. For voltage stability, it is necessary to include the

35
droop characteristics of the AVR rather than to assume zero droop. If load
compensation exists, its effect should be represented. Field current and armature
current limits should also be represented specifically rather than using fixed values for
maximum and minimum reactive power limits. Additionally, for longer term
simulations in the range of few minutes, Automatic Generation Control (AGC)
functions need to be represented appropriately [4].

2.4.3 Reactive Power Compensation

Reactive power compensation is often the most effective way to improve both
power transfer capability and voltage stability. Reactive power compensation can be
divided into series and shunt compensation. It can also be divided into active and
passive compensation. Common forms of reactive compensation are series capacitor
banks, shunt reactors and capacitor banks and static var compensators (SVCs) [5].

For example, when SVCs are operating within the normal voltage control
range, they maintain bus voltage with slight droop characteristics. However, when
they operate at the reactive power limits, they become more or less similar to simple
capacitors or reactors. This could have a significant impact on voltage stability.
Therefore, the characteristics of reactive power compensation should be represented
appropriately in voltage stability studies [4].

2.4.4 Protection and Controls

Generating units and transmission networks protection and control must be


appropriately included in voltage stability analysis. Examples of these protection and
controls are generator excitation protection, armature over current protection,
transmission line overcurrent protection, capacitor bank controls, undervoltage load
shedding, etc. [4].

2.5 Load Modeling For Voltage Stability Assessment

In performing power system analysis, models must be developed for all


relevant power system components. These would include generating stations,
transmission, distribution equipment, and load devices. Models for generation,
transmission and distribution equipment had received great attention. Moreover, their

36
characteristics and parameters are generally very well-known they must comply with
the governing grid codes and regulations. On the other hand the representation of the
loads has received less attention and continues to be an area of greater uncertainty.
Several studies have shown that load representation can have a significant impact on
analysis results, especially in voltage stability assessment. Therefore, efforts directed
at improving load modeling are of major importance. The modeling of loads is
complicated due to a number of factors, including [2]:

Abundance and diversity of load components.


Location and ownership of load devices by customer sites, which make them
not directly accessible to the electric utility.
Continuous change in load composition with time of day and week, seasons
and weather.
Lack of precise information about load composition.
Uncertainties about the characteristics of many load components, particularly
with respect to large voltage or frequency variations.

Consequently, load modeling in power system studies is based on a


considerable amount of simplification. Excluding detailed voltage stability analysis,
the aggregated load is represented in power system models as seen from bulk power
delivery points. The aggregated load represented at a transmission bus usually
includes, in addition to the connected load devices, the effect of sub-transmission and
distribution system lines, cables, reactive power compensation, LTC transformers,
distribution voltage regulators, and even relatively small synchronous or induction
motor [12].

Several literatures indicated that both the static and dynamic properties of
power system loads have a major impact on system stability. Load model uncertainty
was proved, in many publications, to be the major source for simulation inaccuracy.
Therefore, in any system stability study, especially voltage stability studies, it is
necessary to model loads accurately. Having accurate load models capable of
capturing load behavior during system disturbances enhances power system planners
ability to anticipate potential risks and design power systems more precisely.
Therefore, AC appliances must be properly considered while modeling the electrical

37
load of the power system, in order to accurately predict system behavior following
small and large disturbances [13].

2.6 Basic Load Modeling Concepts

The term LOAD can have several meanings in power system engineering
depending on the context. The different meanings of the term "LOAD" include [2]:

Load Device: A device connected to a power system that consumes power.


System Load: The total power (active and/or reactive) consumed by all
devices connected to a power system.
Generator or Plant Load: The power output of a generator or generating
plant.
Bus Load: A portion of the system that is not explicitly represented in a
system model, but rather is treated as if it were a single power-consuming
device connected to a bus in the system model. Bus Load is the one that is of
main concern of this thesis.

When describing the composition of the load, the following terms are recommended
[2]:

Load Component: A load component is the aggregate equivalent of all


devices of a specific or similar type, e.g., air conditioner, fluorescent lighting,
etc.
Load Class: A load class is a category of load, such as, residential,
commercial, or industrial. For load modeling purposes, it is useful to group
loads into several classes, where each class has similar load composition and
load characteristics.
Load Composition: The fractional composition of the load by load
components. This term may be applied to the bus load or to a specific load
class.
Load Class Mix: The fractional composition of the bus load by load classes.
Load Characteristic: A set of parameters, such as power factor, variation of
active power with voltage, etc., that characterize the behavior of a specified

38
load. This term may be applied to a specific load device, a load component, a
load class, or the total bus load.

The following terminology is commonly used in describing different types of load


models [2]:

A load model is a mathematical representation of the relationship between a


bus voltage (magnitude and frequency) and the power (active and reactive) or
current flowing into the bus load. The term load model may refer to the
equations themselves.

Load models are conventionally categorized into static load models and dynamic load
models. The following sections explain both categories.

2.6.1 Static Load Model

Static load model expresses the active and reactive powers at any instant of
time as functions of the bus voltage magnitude and frequency at the same instant.
Static load models are used both for essentially static load components, e.g., resistive
and lighting load, and as an approximation for dynamic load components, e.g., motor-
driven loads [2]. The frequency dependence of loads is not addressed in this thesis,
since in voltage stability incidents the frequency exclusions are not of primary
concern. Various static load-voltage static models are explained below:

Exponential Load Model : A static load model that represents the power
relationship to voltage as an exponential equation, usually in the following form
[2]:


= (2.7)


= (2.8)

Vo, Po and Qo are the reference values for voltage, active power and reactive power
respectively. The exponents and depend on the type of load. Common values for
the exponents and can be found in many literatures including [1] and [5].
However, three particular cases of load exponents are noteworthy:

39
Constant Impedance Load Model (often noted Z): A static load model where
the power varies directly with the square of the voltage magnitude. It may also
be called a constant admittance load model and it is the first form of the
exponential load model when = =2.
Constant Current Load Model (often noted I): A static load model where the
power varies directly with the voltage magnitude. It is the second form of
exponential load when = =1.
Constant Power Load Model (often noted P): A static load model where the
power does not vary with changes in voltage magnitude. It may also be called
constant MVA load model and it is the third form of exponential load model
when = =0.
Polynomial Load Model (often noted ZIP): A static load model that
represents the power relationship to voltage magnitude as a polynomial
equation, usually in the following form [2]:

2
= 1 + 2 + 3 (2.9)

2
= 1 + 2 + 3 (2.10)

With 1 + 2 + 3 = 1 and 1 + 2 + 3 = 1.

Any combination of constant impedance, constant current and constant load


models is called polynomial load model (ZIP).

The main advantage of exponential load model is its simplicity of parameters


identification procedure, since it has only two parameters ( and ) instead of six in
the polynomial (ZIP) load model (p1, p2, p3, q4, q5 and q6). However, the polynomial
(ZIP) load model could appear more of interest due to its physical meaning of
combining more than one load type. If for instance an active load is composed of 50%
of motors and 50% of impedance, it may be modeled using a ZIP load model with
a1 = a3 = 0.5 and a2 = 0 considering that the motor behaves roughly as a constant
power load.

40
2.6.2 Dynamic Load Model

Dynamic load model expresses the active and reactive powers at any instant of
time as functions of the voltage magnitude and frequency at past instants of time and,
usually, including the present instant. Difference or differential equations can be used
to represent such models. A Dynamic load model presents a time dependency that
generally describes a recovery of the load: following a voltage variation, the load
reacts instantaneously before recovering towards a power closer to the previous load
consumption. This class of model can describe phenomena as different as fast
recovery of a motor or as slow recovery of a thermostatic controlled load [2]. Power
systems with large percentage of induction motors require load dynamic
representation. Other dynamic aspects of load components are: the extinction of
discharge lighting with voltage drop, thermostatic control of loads such as space
heaters/coolers, operation of protective relays of some electric devices and response
Under Load Tap Changers (ULTC) of distribution transformers and voltage regulators
[4].

Induction Motor Model: Most dynamic simulation programs include a


dynamic model based on the equivalent circuit shown in Figure 2.6. Other
features available in some programs are additional rotor circuits, saturation,
low voltage tripping, and variable rotor resistance [2].

Figure 2.6: Induction Motor Equivalent Circuit

It is important to note that the slip used in this model is the frequency of the
bus voltage minus the motor speed. Some programs incorrectly use either average
system frequency or 1.0 in place of the bus frequency. As with the frequency-

41
dependent load, such approximations will incorrectly represent damping effects.
Several levels of detail, based on this equivalent circuit, may be available in dynamic
simulation programs, including:

A dynamic model including the mechanical dynamics but not the flux
dynamics.
Addition of the rotor flux dynamics.
Addition of the stator flux dynamics.

Stator flux dynamics are normally neglected in stability analysis and the rotor flux
dynamics may sometimes be neglected, particularly for long-term dynamic analysis.
Low voltage tripping is an important feature for voltage stability analysis and other
studies involving sustained low voltage [2].

2.6.3 Acquisition of Load Model Parameters

There are two basic approaches to the identification of load model parameters, these
are [4]:

Component-Based Approach: This approach was developed by Electric


Power Research Institute (EPRI) under several research projects since 1976. It
involves building up the load model from information of constituent parts as
illustrated in Figure 2.7.
Measurement-Based Approach: In which the load characteristics are
measured at representative substation and feeders at selected time of the day
and season. Alternatively load characteristics are monitored continuously from
naturally occurring system variations. The measurements are then used to
extrapolate the parameters of the load throughout the system. Measurement-
based load modeling provides a closer insight at the real-time power system
loads and their dynamic behavior. Nowadays, static and dynamic power
system behavior can be extracted from a number of data sources including
Power Quality (PQ) monitors, Digital Fault Recorders (DFR), etc.

42
Figure 2.7: Component-Based Modeling Approach [2]

Most dynamic simulation programs allow multiple generators, multiple motor


loads, and a single static load model to be connected to a bus. Generalization of this
capability was recommended in [12] to allow multiple loads of various types on a bus.
Each individual load type (static, induction motor, etc.) may have multiple
representations. For example, a bus load may consist of one or more static models,
one or more induction motors, and a synchronous motor. Each load type may have
load shedding or disconnection logic [12].

2.7 Literature Review on Load Modeling

In the past two decades, load modeling has received a great deal of attention.
Recent findings about the impact of load-to-voltage dependency on power system
voltage stability spurred further interest on the load modeling work. Research results
have shown that the voltage-dependent characteristics of bulk loads not only affect the
voltage secure operating region of a system but also influence the speed of voltage
collapse. A literature review of the key efforts done in Load Modeling during the past
two decades, arranged in chronological order, is given below.

43
In [14], M. K. Pal concluded that motor loads represent a special problem for
system voltage stability. Instability of motors may lead to system collapse and
accordingly, a detailed dynamic analysis is required to identify such instabilities.

In [6], D. J. Hill addressed the modeling of high voltage bus load models for
use in voltage stability studies. The paper concluded that models must account for the
lower level and regulating devices. Thus the models must be dynamic; for instance,
the effect of load recovery following a disturbance is important. The natural
differential equations for induction motors, heating, and tap-changing near a static
load are highly nonlinear and so difficult to parameterize for model estimation. A
somewhat simpler, but still nonlinear first order model was proposed by the author
based on assuming exponential recovery.

In [15], C.J Lin et.al implemented a measurement-based approach for dynamic


load model development. The approach has the advantage of direct measurement of
actual load behavior and can yield load models directly in the form needed for
existing computer program input. The paper described the procedure used for
applying a set of measured data from on-line transient recording system to develop
dynamic load models for the Taipower system. The authors concluded that the
developed second-order and third-order model structures are better in capturing load
behaviors during transients than the first-order load model structure used in earlier
publications. The authors stated that this improvement of capturing load behaviors, at
the expense of complexity in load model, shown up in both real and reactive powers.
They, however, acknowledged that it remains to be seen how many faults need to be
observed in order to derive an accurate dynamic load model.

Reference [2] presents the efforts done by IEEE Task Force on Load
Representation for Dynamic Performance to review the state of art of load
representation for dynamic performance analysis. The task force concluded that
considerable progress has been made in understanding load characteristics and,
particularly, in the methods for determining improved load model data. It was
recommended that serious attention should be given by power system analysts to the
load model and data used in their studies by following the below steps:

44
1. Familiarization with recent literature.
2. Selection of most realistic models for near-term use, based on available data.
3. Investigation of possible existing sources of data on system load characteristics,
such as billing data, load research studies and system disturbance monitors.
4. Development of a plan for acquisition of improved load data, either through
measurements or load mix/composition analysis.

In [16], D. Karisson et al presented a methodology for experimental


determination of aggregate dynamic loads in power systems. The methodology is
oriented to large disturbance studies and motivated towards voltage stability analysis.
The developed models can be expressed as first-order or higher-order differential
equations with a special structure related to steady state and transient load response.
The paper also presented convenient block diagram representation in terms of
nonlinear functions and linear transfer functions. The parameters of the model can be
identified from field measurement data using a combination of closed form formulae
(for step and ramp responses) and least square curve fitting. Identification of a load
model with recovery in the time scale of minutes has been carried out based on field
measurements. Load recovery was originated from thermostat controlled heating
devices; however, the effect of faster recovery in electrical motors has not been dealt
with.

Reference [17] presents the efforts done by W. Xu et al to investigate the


nature of voltage stability from the perspective of load dynamics. The study was
conducted using a generic load model that represents the dynamic behavior of
aggregate loads. The mechanism of voltage collapse was revealed by examining the
interaction of load dynamics with the supply network. The study concluded that:

The dynamic behavior of aggregate loads, which can be characterized by time


constants, transient and steady-state load characteristic parameters is the main
factor for determining the dynamics of voltage collapse.
The rate and the form of voltage instability under a large disturbance are functions
of load time constants, transient load characteristic parameters and the degree of
imbalance between load demand and network supply

45
Network Jacobian matrix and steady-state load characteristics jointly determine
the small-signal voltage stability limit of a power system. The degree of small-
signal voltage stability, however, is also affected by the transient characteristics of
loads.
The generic dynamic load model can easily be expanded to model more complex
forms of load responses.
Modeling loads in an aggregate form simplifies voltage stability analysis.

Reference [12] addresses the efforts done by the IEEE Task Force on Load
Representation for Dynamic Performance as a follow up to their previous work
addressed in [2]. The objective of this publication is to promote better and advanced
load modeling, and to facilitate data exchange among users of various simulation
programs. For transient stability, longer-term dynamics, and small-disturbance
stability programs, the paper recommended the structure of multiple load types
connected to a load bus. The presented static models are suitable for power flow
simulations, and for dynamic simulations at locations where results are not sensitive
to load modeling. Additionally, the paper recommended induction motor models for
use at locations where results are sensitive to load modeling and for longer-term
dynamic simulations. The paper presented also the suitable data to be used with each
model.

Reference [18] presented a nonlinear composite dynamic-static load model


(CDSM) that was developed by P. Ju et al. The main features of CDSM are:

1. It includes the effect of voltage angle on transient active power, which may be
dominative in some cases.
2. It has a parallel structure, which means the total power consists of quasi-
steady-state component, voltage magnitude dependent transient component,
voltage angle dependent transient component, and frequency dependent
transient component.
3. It is applicable, with different time scales, to both angle stability and voltage
stability studies.

The authors also developed a genetic algorithm based parameter estimation (GABPE)
approach. The most attractive performance of the developed approach is its robustness

46
in a complex parameter search space. This approach has been successfully applied to
induction motor models, input-output models and neural network models. The
application results prove that GABPE is simple yet powerful.

In [19], A. Borghetti et al compared between three different simplified models


for describing the behavior of a load consisting of a static load and an aggregate of
induction motors. The different load models have been introduced at the far end of a
longitudinal-structured power system, to account for one of the critical conditions for
voltage stability. The authors concluded that:

When a severe network disturbance occurs, voltage drops at the load side. The
induction motors absorb thus a greater amount of reactive power which leads
to a further decrease of the voltage at the load side and may result in the stall
of the same motors.
In critical load conditions, this phenomenon can be predicted only if an
accurate dynamic model for the motor is employed.
The knowledge of the percentage of asynchronous motors with respect to the
total load, as well as the mechanical load vs. speed dependence is of a primary
importance for selecting the appropriate load model for voltage stability
assessment.

In [20], W. Xu et al performed extensive load tests on the B.C. Hdyro system.


Important factors affecting test design and load parameter estimation were identified
and typical load parameters were determined. Finally, a simple and effective load test
procedure was proposed for future tests. The main conclusions of this paper are:

Load responses due to disturbances caused by transformer tap changing are


sufficient to capture the dynamic and static load characteristics that are critical
to power system voltage security and stability studies. Transformer tapping
disturbances can be easily implemented without significant adverse effects on
the customers. Thus a special instrument may be built for regular load
characteristics monitoring. The instrument relies on the load responses to
voltage disturbances caused by such means as transformer tap changing to
estimate the load real and reactive power demand sensitivities to supply
voltages.

47
Load parameters corresponding to both the summer and the winter seasons
indicate that the loads have significant real and reactive power demand
sensitivities to the supply voltage. This is largely due to the differences in the
seasonal-dependent load composition. It indicates the need for performing
load tests in different seasons.

In [21], L. M. Hajagos et al performed laboratory measurements and derived


models of modern loads subjected to large voltage changes and investigated their
effect on voltage stability studies. Low-voltage, long-time models of different type of
loads such as modern Air Conditioners (ACs), discharge lighting, and devices
containing electronic regulated power supplies were developed. Results of the
measurements showed that modern loads typically have a higher power factor, wider
voltage range, and more constant power characteristics than their predecessors. It was
also found that simulations performed with conventional load models produced both
conservative and optimistic results compared with studies using the accurate load
models derived in this project. Adherence to familiar-looking load model parameters
(e.g. 50% constant current, 50% constant impedance) does not necessarily accurately
represent load behavior. The paper strongly recommended that if detailed studies
involving load models are being performed, the accurate parameters derived in this
project can be used for the most important loads.

In [22], K. Tomiyama et al performed a series of measurements on single-


phase air conditioners in order to understand their voltage and frequency
characteristics and their interaction with the network dynamics. The voltage and
frequency characteristics of two conventional single-phase air conditioners and an
inverter air conditioner have been measured. The measurements indicated that:

1. The P of the conventional air conditioners studied is practically invariant


within a steady-state voltage range of 0.9 p.u. to 1.1 p.u. The Q of the
conventional air conditioners increases significantly when the voltage is high
and the frequency is low.
2. The P and Q of the inverter air conditioner are practically invariant within the
same voltage range and a frequency range of 55 Hz to 65 Hz.

48
3. Both conventional and inverter air conditioners stall around 0.6 p.u. voltage
when the voltage is decreased gradually. The inverter air conditioner drops out
at around 0.5 p.u. voltage, but the conventional air conditioner stays on line
and can aggravate low voltage situations.
4. Starting a conventional air conditioner at the rated voltage requires the
minimum energy and the minimum time integral of Q. The inverter air
conditioner starting characteristics are invariant within the studied voltage range
of 0.9 p.u to 1.1 p.u.
5. With severe voltage dips, the conventional air conditioner can stall and draw
significant amounts of P and Q for a few seconds before it drops out. The
inverter air conditioner drops out at the end of severe voltage dips. The
inverter air conditioner behaves well during voltage collapse situations while
the conventional air conditioner can aggravate the situations.

Simple models have been developed to represent the measured characteristics. The
load modeling approach has been used to correlate the observed bus load
characteristics to the estimated cooling load magnitude. The authors hoped that the
characteristics of the base load component and those of the cooling load component
can be determined separately.

In [23], S.Z Zhu et al analyzed the relation between the mechanism of voltage
stability and load models. Two methods, Instrumental Variable (IV) and Minimum
Sum of the Absolute Residues (MSAR) were used to identify the parameters of
nonlinear and linear load model. Eigenvalue-Structure Analysis (ESA) and two other
sensitivity methods based on ESA were used to analyze the voltage stability
sensitivity on state variables or control variables and sensitivity of total generated
reactive power. The study concluded that the load parameters and models have an
important effect on the study of voltage stability. The calculation results of voltage
stability with constant resistance are more conservative than the results obtained with
load parameters from real field test.

Reference [24] presented an improved composite load model structure that can
be used for measurement-based approach. This model was developed by J.H. Shi et
al. The authors stated that the improved composite load modeling structure was

49
compared with a typical dynamic load model and found that it can describe the
reactive power of the loads more accurately.

In [25], K. Morison et al provided a practical approach to develop models for


use in voltage stability analysis. Based on billing data or load inventory surveys, the
load is classified into broad types such as residential, industrial, and commercial load
and each is subdivided into more specific components including induction motors and
other elements. Time-domain simulations were used to establish the load
characteristics and general purpose models were synthesized from the results for use
in static analysis.

Reference [26] presented a neural network methodology for dealing with static
and dynamic load modeling developed by D. Chen et al. The load patterns were
classified by feed forward neural networks. Based on the static load model and
dynamic load model, either static voltage stability analysis or dynamic voltage
stability analysis can be made. The sensitivities involved in neural-network models
for loads were derived, and were then used in the Jacobian matrix, and further for the
modal analysis. Both static and dynamic voltage stability analyses were performed
which included use of the neural-network-based load models. The neural-network
methodology is tested on both the IEEE-14 bus system and real field data.

In [27], P. Pourbeik et al addressed the application of detailed load modeling


of large power system. The paper illustrated the importance of load modeling
assumptions to the results of a voltage stability study. In particular the following
factors were found to play a major role in the outcome of the study results:

Load Amount: As the amount of load increases it becomes harder and harder
to sustain stability. That is, with more load there are more motors being slowed
down by a fault and accordingly more reactive support is needed.
Fault Duration: The longer the fault lasts, the longer motor loads will
decelerate and thus the more likely motors will stall thus dragging system
voltages down.
Fault Severity: The more severe the fault is, the greater the decelerating torque
on the motors. Also, for multi-phase unbalanced faults, the negative sequence
braking torque will further decelerate motors. Most transmission faults tend to

50
be single-line to ground faults, for which the consequences can be
significantly less severe than a 3-phase bolted fault.
Fault Location: The closer the fault is to the substation, the lower the bus
voltage at the substation and the more severe the problem will be.
Transmission Outages: The outage of a major transmission line may severely
compromise the ability of the transmission system to support voltages. This
will hamper the ability of the system to reaccelerate the affected motor loads.
Motor Loads Amount and Characteristics: The lower the motor inertias are the
quicker the motors will decelerate. Also, the motor torque-speed curve will
dictate the break-down torque of the motors and thus the point at which they
stall.

Reference [28] addresses the efforts done by V. Stewart who deployed a


simple methodology for modeling the effects of stalled motor loads, particularly
residential ACs and refrigeration motor loads, for short-term voltage stability analysis
using a sample utility system. The methodology included using a constant-
impedance/admittance model for stalled motors, based on an assumed percentage of
motor loads and remaining loads using a complex load model with large and small
motor loads, discharge lighting constant power, and other loads. The active and
reactive power demands of stalled motor loads were determined based on basic
equations and assumed values of power factor and locked-rotor current for motors,
and then applied during dynamic simulations. Dynamic simulations using a sample
utility system were used to evaluate the performance of the stalled motor and complex
load models. The dynamic voltage profiles obtained illustrated effectively the impact
of stalled motor loads on system voltage stability.

In [29], B. K. Choi et al identified four dynamic linear and nonlinear models


based on eleven sets of measurement data gathered at different loading conditions in
Taiwan Power System. The process of model validation was performed using the
cross-validation technique. Analysis of parameter variation of each dynamic load
model with respect to different loading conditions was also conducted. From this
analysis, it was observed that the parameter values of each dynamic model can be
quite different for different loading conditions, which suggests that parameters of load
models should be properly updated in order to represent the load characteristic at each

51
loading condition. A simple and efficient method was presented to estimate a
representative parameter set for different loading conditions. The authors emphasized
that accurate dynamic load models can capture dynamic behaviors of reactive power
as well as real power during disturbances. However, the use of dynamic load model
increases the statespace dimension of the system. Considerable care must be taken in
attempting to validate the derived load models for practical use on the basis of time-
domain dynamic simulation. This is due to the fact that several components in power
system, including excitation system model, power system stabilizer model, and load
model, can significantly affect the time-domain simulation results. The authors
concluded finally that the linearized model proposed in this paper and the small
induction motor model appears to be good aggregate dynamic load models for both
real and reactive load behaviors.

In [30], K. Morison et al highlighted that the complexity of loads make it


impossible to represent loads exactly, and therefore, uncertainty is present in any
simplified or aggregate load models developed to represent load dispersed throughout
distribution systems. Sensitivity analyses should be conducted in order to assess the
impact of variations in the load parameters on both the system dynamic response and
also stability limits. Some of the recommended sensitivities which should be
considered include:

The percentage of induction motors in the aggregate load model


The induction motor parameters.
The characteristic of the induction motor driven load.
The effect of motor protections.

B. WU et al investigated the effects of air conditioner load on voltage stability


of an urban power system based on reactive power-voltage characteristics analysis of
Air Conditioners (ACs). The study which was addressed in [31] concluded that high
percentage of ACs at load bus will reduce the maximum reactive load supplied from
the bus and increase the critical voltage, and therefore push the bus voltage security
into danger. A further analysis of loading margin employing continuation power flow
shows that AC load may reduce the maximum loadability of the system, and plays an
adverse role in the voltage stability of power systems. The paper recommended the

52
installation of sound reactive power compensation measure to enhance voltage
stability of urban power system.

In [32], L. Y. Taylor et al made use of the actual voltage signature captured by


data fault recording equipment of a practical multi-fault multi-breaker failure to depict
an accurate load model capable of representing the incident. An aggregate load model
made up of additional distribution system impedance, 50% small induction motor, and
50% static load resulted in successful replication of the delayed voltage recovery
event through dynamic simulations. The same model has been successfully utilized in
the replication of voltage signatures resulting from milder fault events. However, it
was recognized that the aggregate load model could possibly be improved by
including the embedded load tripping due to induction motor protective relay action.

T. Parveen et al developed an approach for decomposition/separation of the


composite induction motors load from measurement at a system bus [33]. In power
system transmission buses, load is represented by static and dynamic loads where the
induction motor is considered as the main dynamic loads and in the practice for major
transmission buses there will be many and various induction motors contributing to
the load composition. Particularly, at an industrial bus most of the load is dynamic.
Rather than trying to extract models of many machines, the paper seeks to identify
three groups of induction motors to represent the dynamic loads. Three groups of
induction motors were used to characterize the load. These were: the small motors
group (4 kW to 11 kW), the medium motors group (15 kW to 180 kW) and the large
motors group (above 630 kW). At first, these groups with different percentage
contribution of each group were composite. After that from the composite models,
each motor percentage contribution was decomposed by using the least square
algorithm. To apply this theory to other types of buses such as residential and
commercial it is good practice to represent the total load as a combination of
composite motor loads, constant impedance loads and constant power loads. To
validate the theory, the 24hrs of Sydney West data was decomposed according to the
three groups of motor models.

In [34], K. Yamashita et al used measured data to identify static and dynamic


load model parameters. Transfer function-based simple dynamic load model was
applied for identifying static and dynamic load model parameters. The derivation

53
method of static load model parameters, with and without load drop, was developed.
The load model responses calculated by the measurement data (load bus voltage and
frequency) and the identified static load model parameters were consistent with the
measurement data. The derivation method of dynamic load model parameters without
load drop through high frequency measurement data was also developed. The
dispersion of the identified dynamic load model parameters is sufficiently small,
which means the dynamic load model is effective for representing load response with
voltage drop within around 30%. However, the response of the transfer function-
based dynamic load model were not consistent with around 20 % of the measurement
data mainly because the load time constant immediately after the fault was different
from the time constant immediately after the fault clearance. This tendency can be
seen in the response of induction motors.

I. F. Visconti et al proposed a methodology of measurement-based load


modeling for transient stability analysis [35]. The paper presented practical results
compared to power quality recordings. The technique proposed to estimate load
model parameters based on Genetic Algorithms (GA). The following steps summarize
the system identification procedure for load modeling:

Select input-output system data.


Transform, resample and cluster data.
Choose an appropriate model structure.
Estimate load model parameters.
Validate load model.

A second order autoregressive model proved to cover a wide range of situations.


Moreover, the third and fifth order autoregressive models were also tested but the
results were not much different from second order models. The authors
recommended that measurement-based load modeling should be adopted by utilities
interested in taking full advantage of their resources and data storage by power quality
monitors. They concluded that measurement-based load modeling is the state-of-the-
art on load representation and shall improve reliability on bulk power system stability
studies.

54
In [36], S. A. Al Dessi et al developed a detailed dynamic load model for
Dubai Power Grid based on a monitoring campaign of the transmission network and
staged testing of sample Low Voltage Air Conditioning (LV AC) appliances. The
study highlighted the growing dimension of the fast embedded under voltage load
shedding (EUVLS) phenomenon which is driven by the advent of the inverter based
appliance electrical interfaces, and the protection settings of the main components of
the large chillers and cooling plants. The adopted load model has shown that the
EUVLS phenomenon can significantly affect the nature and the design of the
mitigation measures to secure the voltage stability of the power system, especially the
required operating range of the fast acting MVAr reserves such as SVS. Due to load
diversity, the study suggested monitoring the amplitude of the EUVLS phenomenon
with dedicated DFR devices in order to reduce the severity of the assumptions
performed and to remain on the secure side of the voltage stability assessment.

In [37] P. Regulski et al presented a compact solution based on instantaneous


measurements to extract parameters of a load model, which includes estimation of
power components based on an Improved Recursive Newton Type Algorithm and
estimation of parameters of a dynamic Load Model using Genetic Algorithms. The
paper demonstrated the influence of the preprocessing of the instantaneous values on
the final estimation of load parameters. The method has been tested in two stages
using 9-buses P.M.Anderson test system built in DIgSILENT. Firstly, the quality of
extracted RMS values has been presented, followed by the estimation of parameters
of a dynamic load model performed by GA. The procedure returned considerably
good results proving that the combination of both methods is suitable for this task.

M. Tomoda et al proposed a dynamic load model considering a change of flux


inside the induction motor for transient stability analysis and presented a new method
for estimating unknown parameters of the dynamic load model [38]. The model is a
parallel composite of a constant impedance load and an induction motor load behind a
series constant reactance. This model can represent the behavior of actual load by
using appropriate parameters. However, the problem of this model is that it has a large
amount of parameters and it is not easy to estimate a lot of unknown parameters.
Therefore, the authors proposed an estimating method based on Particle Swarm
Optimization (PSO), which is a non-linear optimization method, by using measured

55
instantaneous voltage sag data. The model was tested and it was confirmed that the
proposed method was successful and PSO is effective in parameter estimation.

In [39], B. Ho-Kim et al addressed the composite load model in which the ZIP
model and the induction motor model are used for the static and dynamic load models
respectively. The paper described the parameter estimations needed for the selected
load model. The authors obtained the measurements using the TSAT simulation
program. The parameter estimation using the least square method is then verified
using these measurements. The author concluded that, despite some oscillations, the
results are satisfactory in general and the calculation time is fast. Accordingly, they
recommended using the parameter estimation method for real time applications.

In summary power system loads have a significant impact on the stability of a


system. A proper understanding of load behavior is crucial in voltage stability
analysis, as it can provide information necessary for the accurate assessment of
different scenarios. However, accurate load modeling continues to be a difficult task
due to the complex and changing nature of the load and the difficulty in obtaining
accurate data on its characteristics, hence, uncertainty is encountered in any simplified
or aggregate load model. Therefore, sensitivity studies are recommended to determine
the impact of the load characteristics on the study results of interest. This will help in
guiding the selection of a conservative load model or focusing attention on where load
modeling improvements should be sought.

56
Chapter 3: Updating Dubai Power Grid Load Model

3.1 Overview of Dubai Power Grid

The Dubai Electricity and Water Authority (DEWA) was formed by a decree
issued by His Highness Sheikh Maktoum bin Rashid Al Maktoum on January 1,
1992, to take over and merge the Dubai Electric Company and the Dubai Water
Department that had been operating independently for several years until then. Both
these organizations were established in 1959 through the foresight and initiative of
His Highness Sheikh Rashid bin Saeed Al Maktoum, the former Ruler of Dubai, as
government supported bodies with the objective of making available to the people of
Dubai an adequate and reliable supply of electricity and water [40]. Figure 3.1 is a
geographical map of Dubai showing the existing power system [41].

Figure 3.1: Dubai Power Grid Geographical Map [41]

Dubai Power Grid has been developing rapidly since the foundation of the
utility in the emirate of Dubai in 1959. Initially, the generation and transmission
voltage was at 6.6 kV and gradually a higher voltage of 33 kV was introduced in
1969, which remained as the primary transmission voltage till 1977. During 1978, the
132 kV system was introduced in Dubai and in 1993, 400 kV lines and substations

57
were introduced. There has been a considerable growth in the power system
infrastructure since the introduction of the first 132 kV substations in 1978. The rapid
growth of the system can be seen from the statistics related to the installed capacity
and peak demand as shown in Figure 3.2, as well as the substations and lines statistics
shown in Table 3.1 [41].

Figure 3.2: Electricity Installed Capacity & Peak Demand (2002-2012) [41]

Table 3.1: Substations and Lines Statistics for Dubai Power Grid (2002-2012) [41]

Description Voltage Unit 2002 2003 2004 2005 2006 2007 2008 2009 2010 2011 2012

400 kV Nos 7 7 7 8 10 11 12 14 14 17 18
Number of
substations
132 kV Nos 31 33 37 52 69 88 116 139 153 165 184

400 kV km 291 291 291 344 583 637 766 870 870 875 876
Overhead
Lines
132 kV km 543 543 543 535 529 517 466 466 459 437 437

Underground
132 kV km 228 258 266 369 495 642 867 1002 1137 1250 1486
Cables

The total number of consumers by end of year 2012 is 624445 classified into:
Residential, Industrial, Commercial and Others (Non-Commercial such as Mosques,
Police Stations, Government Hospitals, Government Schools, DEWA Offices & Staff
premises, etc.). The total electricity consumption by year 2012 reached 35124

58
(GWhs). Figure 3.3 shows the percentage of different consumer categories and the
corresponding electricity consumption for 2012 [40].

Figure 3.3: Percentages of Energy Consumption for different Consumer Categories - Year 2012 [40]

3.2 Problem Statement

A significant part of the load in Dubai and the Gulf countries in general is
dominated by AC appliances, especially during summer period. Recently, the
demand side at Dubai Power Grid had incurred enormous penetration of large and
small motor equipment ranging from small home Air Conditioning (AC) appliances to
large District Cooling/Chiller Plants (DCPs). During large system disturbances, the
transmission system voltage can fall below a critical threshold, resulting in induction
motors stalling or tripping depending on several factors such as motor type, size and
control. Voltage collapse can occur when stalled motors remain connected to the
system, or motors previously disconnected are automatically reconnected causing a
large, mostly reactive, starting current. On the other hand, in case of excessive motor
tripping, the system voltages may increase above the specified planning and
operational limits. Consequently, and in case the reactive power compensation
facilities were not adequately sized in the power system, the excess reactive power
will be absorbed by generators driving them operate in an under-excited mode which
may eventually lead to their cascading tripping.

The Dubai Power Grid blackout of the 9th of June 2005 showed that load
variation, which took place after the under frequency load shedding was not in line
with the expected behavior. Moreover, on 18th of June 2006, 37% of the supplied
load was inadvertently shed at customer side following a single phase to ground fault

59
at 'L' Power Station 400 kV bus. Both incidents raised the importance of having
accurate representation of the supplied loads considering frequency and voltage
dependency. Accordingly, in 2008, DEWA decided to develop a more detailed load
model for Dubai Power Grid to be used for dynamic stability studies (especially the
ones leading to severe transients of the power system). A dedicated monitoring
campaign was conducted on five selected 132/11 kV substations at 11 kV side. The
selected monitored substations represent different load classes and mixes including:
residential, commercial, industrial, and district cooling.

The load modeling effort also included laboratory staged tests performed on
representative set of LV AC appliances that are commonly used in Dubai. The overall
monitoring campaign and the laboratory tests gave clear evidence that a portion of the
load is subject to a sudden disconnection following voltage disturbances. The efforts
done were addressed in [36]. In summary, the developed load model adopts the
assumption of a homogenous load composition at all load substations, which is 35%
static load and 65% motor load (of which 35% are large motors and 30% are small
motors) as illustrated in Figure 3.4 .

Figure 3.4: Dubai Power Grid Load Composition (Existing Load Model)
Moreover, five different stages were proposed to represent motor tripping following
the occurrence of voltage dips. The percentage of tripped motors is directly related to
the magnitude of the dip and all the tripping was assumed to occur simultaneously
(after 250 ms delay time) as described in Table 3.2 below.

60
Table 3.2: Load Tripping Rules for Motor Load (Existing Load Model)
Voltage Load Tripping
Stage Sensing Time Tripping Time
Threshold Fraction
1 90% 19% 50 ms 200 ms
2 72.5% 37% 50 ms 200 ms
3 55% 55% 50 ms 200 ms
4 37.5% 77% 50 ms 200 ms
5 20% 77% 50 ms 200 ms

On 20th of October 2009, a single phase to ground fault at the 400 kV bus of
Mushrif (MUSH) 400/132 kV load substation resulted in tripping 8% of the total
supplied load. Figure 3.5 shows the recorded voltage at MUSH 132 kV bus. The short
circuit resulted in a sudden voltage drop. After fault clearance, the voltage recovered
to about 100 kV. After reacceleration of the motors the voltage recovered to about
117 kV within 0.5 seconds demonstrating a fast voltage increase. Thereafter, the
voltage recovered to the pre-fault voltage during a 10 seconds period, demonstrating a
slow and smooth voltage increase. The same conclusions can be deduced from
frequency trend of the same disturbance. Figure 3.6 shows the frequency, system
requirement and interchange flow trend on 20th of October 2009. A zoomed in part of
the frequency trend curve during this incident is shown in Figure 3.7.

Figure 3.5: Recorded Voltage at MUSH 132 kV bus during MUSH Disturbance [42]

61
Figure 3.6: Frequency, System Requirement and Interchange Flow Trend during MUSH
Disturbance [42]

Figure 3.7: Zoomed in part of the Frequency Trend Curve during MUSH Disturbance [42]

It is obvious from Figure 3.6 and Figure 3.7 that there is a temporary
frequency rise due to the short circuit at Mushrif. After fault clearing, the frequency
drops back again and then rise again within 10 seconds indicating reduction in the
system load and hence a slow voltage recovery in the voltage during the same period.

62
The existing load model was used to replicate the disturbance by performing
dynamic simulation. However, there were great discrepancies between the measured
and simulated voltages as shown in Figure 3.8. This is due to the fact that the
developed load model was built on the assumption of simultaneous tripping of motor
loads following voltage dips whereas the actual disturbance indicated that a part of the
load tripped during 0.5 second period after fault clearance, and another part tripped
during 10 second period. Moreover, the simulation resulted in tripping a higher
amount of load than the actual recorded amount as shown in Figure 3.9. Therefore, it
is obvious that the developed load model needs further modification in order to
represent, with accepted accuracy, the actual load behavior of Dubai Power Grid
following system disturbances.

V(p.u.)
1.2

0.8

0.6

0.4

0.2

0
0 1 2 3 4 5 6 7 8 9 10
Time (s)
V_MUSH_132_Recorded V MUSH 132kV - Existing Load Model

Figure 3.8: Recorded vs. Simulated Voltage Trend at MUSH 132 kV Bus MUSH Disturbance
(Existing Load Model)

Total System Load


(MW) Tripped Load = 12%
6000

5500

5000

4500

4000

3500

3000
0 1 2 3 4 5 6 7 8 9 10
Time (s)

Figure 3.9: Simulated Total System Load during MUSH Disturbance (Existing Load Model)

63
3.3 Methodology

Whatever the methodology used by the utility to build the load model, the
developed load model should answer three questions [42]:

1. What is the appropriate load model structure that is capable of representing the
aggregated load?
2. For the selected load model structure, how should the load model parameters be
identified?
3. Does the developed load model have an adequate generalization capability?

Based on the analysis done in [30], the following basic principles of load modeling
were developed:

1. The load model is built mainly for performing stability studies; therefore, it should
be easily integrated into the power system analysis tool.
2. The developed load model should be general and valid under most scenarios. For
special studies, a special load model should be built according to the study
requirement.
3. The load model should have clear physical interpretations in order to be accepted
and easily implemented by the system analysts and operators

The above principles were taken into consideration while formulating the
methodology to update the existing Dubai Power Grid Load Model. The developed
methodology is a combination of the two basic approaches to the identification of
load model parameters addressed in 2.6.3, namely, Component Based Approach and
Measurement-Based Approach. This hybrid methodology is suitable for developing
load models for any power system based on real data and actual system
measurements. The new hybrid methodology is described in details in the following
sections and its different process is illustrated in Figure 3.10

64
Figure 3.10 : Illustration of the Developed Load Modeling Methodology

65
3.3.1 Component Based Approach Processes

It has been found from literature that surveying load composition instead of
load component to develop models for dynamic simulations is more practical to
derive the basic load models for short-term voltage stability studies. This means,
individual devices are not surveyed, but rather load classes [30].

3.3.1.1. Classification of Load Classes and Mixes

Classification of load into classes and mixes can be done at either transmission
or distribution system level, depending on the nature of the study that shall be
performed using the developed load model. For global system studies, classification
of load at transmission system level is adequate. However, for special load zone
studies, load classification needs to be done at the distribution system level. In this
thesis, the purpose of the developed load model is to assess the overall system voltage
stability; therefore, the load classification is done at the transmission system level,
which means at the end of the process, the identified load composition will be unified
for the whole system. Another reason behind this choice of classification level is that
Dubai Power Grid is compact and almost every load zone has similar load
composition, i.e. each load zone is feeding all different types of load classes with very
close percentages. Moreover, the system is growing quickly requiring shifting load
substations from a zone to a nearby one resulting in higher uncertainty of load
composition for each zone. Based on this, and with the help of the concerned
departments in DEWA, the existing 132/11 kV load substations were classified into
classes and mixes including: Residential, Commercial, Industrial,
Residential/Commercial etc. Table 3.3 is an excerpt of the obtained information.

Table 3.3: Sample of Load Classes/Mixes in Dubai Power Grid System Level

Load Class/Mix Percentage (%)


Load Voltage
No. Substation Level District
Name (kV) Residential Commercial Industrial Cooling Total
Plant
1 BDXB 132/11 50 50 100
2 DMAL 132/11 50 50 100
3 OTWN 132/11 15 15 5 65 100
4 FRDH 132/11 50 50 100
5 HLAL 132/11 25 30 45 100
6 KHLS 132/11 15 10 75 100

66
3.3.1.2. Identification of Load Composition for Each Load Classes and
Mixes

Resistive, Small Motor, Large Motor, Discharge Lighting, and other


components are identified based on the load classification (percentage of Residential,
Commercial and Industrial) and composition (percentage of each component in each
load class) obtained from published typical models. Unless better information is
available, this typical data can be used in deriving the overall load model. Table 3.4
is an example of the available typical load compositions that utilities serve [30].

Table 3.4: Typical Load Composition for Different Load Classes [30]
Load Composition (%)
Load Class
Residential Commercial Industrial
Resistive 25 14 5
Small Motor 75 51 20
Large Motor 0 0 56
Discharge Lighting 0 35 19

Table 3.4 was modified based on the conducted demand side survey to reflect
the load compositions for each load class in Dubai Power Grid. The fact that some
residential and commercial areas are being cooled by District Cooling/Chiller Plants
was behind adding a percentage of large motors to these two classes. On the other
hand, the discharge lighting was aggregated with the resistive load and given a more
general name, Static. The modified load compositions are shown in Table 3.5.

Table 3.5: Modified Load Composition for Dubai Load Composition


Load Composition (%)
Load Class District Cooling
Residential Commercial Industrial
Plant
Static 25 49 24 10
Small Motor 75 51 20 20
Large Motor 0 0 56 70

Moreover, based on the conducted demand side survey small motors were
divided into two categories: Conventional (non-inverter based) and Non-Conventional
(inverter based). The result of the implemented component based approach processes

67
is illustrated in Figure 3.11 which shows Average Load Compositions at system level.
As explained earlier, the identified load composition can be used for global system
voltage stability studies, which is in line with the scope of this thesis. For local
voltage stability studies concerning a certain load zone, detailed load compositions for
the zone under study may be required.

Figure 3.11: Result of Implementing Component Based Approach on Dubai Power Grid

3.3.2 Measurement Based Approach Processes

System identification is considered as the theoretical foundation of


measurement-based approach. It has developed into a mature engineering discipline,
which is widely applied in many branches of modern engineering. The process of
system identification involves finding a mathematical model (suitable model
structure) with a set of parameters that is capable of accurately replicating the
dynamic response of the system [13]. Once the model structure is selected, the model
parameters can be estimated using parameter estimation techniques. The objective
function for parameter estimation is usually defined as the minimization of the
difference between model outputs and actual measurements. Generally, measurement
based approach for load modeling consists of three steps; selection of a suitable
aggregate load model structure, parameter estimation and load model validation [43].

68
3.3.2.1. Selection of Aggregate Load Model Structure

The selection of the load model structure depends on various aspects including
the following [30]:

1. The load model structure should not be very simple. A very simple load model
will result in a big bias error, and hence great discrepancy between the
measured and estimated quantities.
2. The load model structure should not be very complex. A very complex load
model will result in a high model variance error. Therefore, while a complex
load model may fit the training data well, nevertheless, it may have poor
generalization capability.

Besides that, the applicability of the selected load model structure in power system
simulation software is essential. Most power system simulation software package,
such as Siemens PTI Power System Simulator for Engineers (PSS/E), include
various aggregate load model structures that are user friendly and covers wide load
compositions. For more information, refer to PSS/E user manual [10]. In this thesis,
the aggregate load model structure selected for parameter estimation consists of a
static part (Constant Impedance, Constant Current and Constant Power (ZIP)) in
parallel with dynamic part (induction motor) as shown in Figure 3.12. This load
model structure is suitable for reflecting the load composition identified for Dubai
Power Grid.

Figure 3.12: Equivalent Circuit of the Selected Aggregate Load Model Structure [29]

69
The static part of this aggregate load model structure is the polynomial (ZIP) load
model described earlier in 2.6.2 and represented by equations 2.9 and 2.10. On the
other hand, the dynamic part of the selected aggregate load model structure is the
induction motor model. In this thesis, the fifth order model of induction motor is used
for parameter estimation and is represented by the following five differential
equations [44]:


= + ( ) (3.1)


= + + ( ) (3.2)

( )
= + ( ) (3.3)

( )
= + + ( ) (3.4)

1
= [ ] (3.5)
2

Where


= + (3.6)


= + (3.7)

1 1 1
= 1 + + (3.8)

= ( ) (3.9)

= (2 + + ) (3.10)

The currents in equation (3.9) can be calculated as follows:

1
= (3.11)

1
= [ ] (3.12)

70
The active and reactive power of the induction motor are calculated as follows:

= + (3.13)

= (3.14)

Therefore, the total power consumed by the selected load model structure can be
written as follows:

= 1 + 1 (3.15)

= 2 + 2 (3.16)

With 1 + 1 = 1 2 + 2 = 1

Where
d: direct access,
q: quadratic access,
s: stator variable,
r: rotor variable,
and : direct and quadratic access stator flux linkage,
and : direct and quadratic access rotor flux linkage,
and : direct and quadratic access magnetizing flux linkage,
and : direct and quadratic access stator voltages,
and : direct and quadratic access rotor voltages,
and : direct and quadratic access stator currents,
Rs and Rr : stator and rotor resistance,
Xls and Xlr : stator and rotor leakage reactance,
Xm: magnetizing leakage reactance,
: stator angular electrical frequency,
: motor angular electrical base frequency,
: rotor angular electrical speed,
H: inertia constant,
Te : electrical output torque,
To : initial load torque,
TL: load torque,
A, B and C: torque coefficients that obey the following equality A+B+C=1,
PIM and QIM: active and reactive power consumed by the induction motor.
PL and QL: active and reactive power consumed by the aggregate load.
1 , 2 , 1 and 2: static and dynamic load percentages.

71
3.3.2.2. Load Model Parameter Estimation

The purpose of parameter estimation techniques is to find a set of model


independent parameters that results in the best fit between the recorded measurement
and model output. Parameter estimation techniques are well established for linear
models, however, for nonlinear systems are still a relatively open field. In general
there are three approaches for parameter estimation of nonlinear dynamic models
[13]:

1. Analytical Based Approach: This kind of approach derives parameters


determinately and can be used for special tests such as step/staged/controlled test.
However, it is extremely sensitive to measurement error.
2. Optimization Based Approach: This kind of approach searches the best
parameters, which minimizes an error function between the measured output
variables and simulated ones. Traditional search algorithms have been applied in
this type of approach such as :
a. Statistical Search Algorithms: such as least square and gradient search, etc.
b. Heuristic Search Algorithms: such as genetic algorithm, simulated
annealing, etc.
3. Stochastic Based Approach: These approaches may be limited to the
assumptions of the error function (noise).

In search of best-fit parameters, there is an important issue to consider, that is,


the data are subject to measurement errors (noise). Therefore, typical data never
exactly fit the model that is being used, even when that model is correct [13]. The
least squares method was selected for estimating load parameters, since it is used in
many applications and easy to implement. Furthermore, it is robust, and on the
contrary, the real application of Artificial Intelligence (AI) based approaches is
limited partly due to lack of theoretical and practical guidance on selection of various
parameters that affect these algorithms [13]. The selected aggregate load model
structure was built in MATLAB Simulink for the purpose of parameter estimation as
shown in Figure A.1 (Appendix A).

72
Based on the selected load model structure, load parameters that need to be identified
are summarized in Table 3.6.

Table 3.6: Parameters to be identified for the Aggregate Load Model

Load Model Structure Parameters to be Estimated


ZIP + Induction Motor Load Composition 1 , 2 , 1 and 2
ZIP model 1 , 2, , 3 , 1 , 2 3
Induction Motor Model Rs, Rr, Xls, Xlr , Xm, H, A, B and C

The Objective function for parameter identification is defined as the difference


between estimated load model outputs and actual measurements. The mathematical
formulation of the objective function is as follows [43]:


1 1
( ) = 2 () = [ () ()]2 (3.17)

=1 =1

1 1
= 2 () = [ () ()]2 (3.18)

=1 =1

Where are the measured active and reactive powers, are the
estimated load model active and reactive powers and () represents the load model
output error.

As mentioned earlier, DEWA have installed DFRs on five selected 132/11 kV


substations at 11 kV side representing different load classes and mixes for the purpose
of load modeling and continuous updating. These DFRs are capable of continuously
recording the normal variations in frequency as well as instantaneous voltages and
currents of the three phases. The DFRs were programed to capture system
disturbances such as voltage dips and frequency transients. The available data from
the DFRs can be easily exported in (*.csv) format. This facilitates plotting curves of
interesting quantities and extracting useful information using MS EXCEL or
MATLAB. All the measured and estimated active and reactive power are converted
to positive sequence to match the quantities adopted in transient stability analysis
[45]. In order to calculate positive sequence active and reactive power, the

73
instantaneous samples of voltage and current signals are first converted into phasor
form, corresponding to their fundamental frequency, using Discrete Fourier
Transform (DFT). Then, the positive sequence voltage and currents are calculated
Using Fortescue transformation. The positive sequence active and reactive power are
calculated as:

+ + + = 3 . + . + (3.19)

Where V+ is the positive sequence voltage phasor and I+* is the complex conjugated
positive sequence current phasor. P+ and Q+ are the positive sequence active and
reactive power. The overall procedure of measurement based load modeling is
illustrated in Figure 3.13.

Figure 3.13: Parameter Estimation Procedures

A set of four interesting recorded voltage dip incidents have been carefully
selected for parameter estimation. Two of the selected incidents are three phase
voltage dip and the other two are single phase voltage dip. Measurement-based
approach generally assumes the load characteristics do not change during the time
interval that measurements were taken. This is because the time interval of the load
measurement data used for load parameter estimation is from 1s to 10s, which is

74
enough to capture induction motor dynamics. Therefore, for the selected set of
incidents, it is assumed that the percentages of static and dynamic load for, each
incident remain constant during the period that measurement data were recorded. A
description of each incident and the corresponding measured quantities is given
thereafter:

Incident No.1:

This incident has resulted in three phase voltage dip of 850 ms duration. Figure 3.14
shows the measured instantaneous three phase voltage and current signals and
Figure 3.15 shows the calculated positive sequence voltage and current.

Figure 3.14: Measured Instantaneous Voltage and Current Signals Incident No.1

Figure 3.15: Calculated Line Positive Sequence Voltage and Current Incident No.1

75
Incident No.2:

This incident has resulted also in three phase voltage dip of 800 ms duration.
Figure 3.16 shows the measured instantaneous three phase voltage and current signals
and Figure 3.17 shows the calculated positive sequence voltage and current.

Figure 3.16: Measured Instantaneous Voltage and Current Signals Incident No.2

Figure 3.17: Calculated Line Positive Sequence Voltage and Current Incident No.2

76
Incident No.3:

This incident has resulted in single phase voltage dip of 250 ms duration. Figure 3.18
shows the measured instantaneous three phase voltage and current signals and
Figure 3.19 shows the calculated positive sequence voltage and current.

Figure 3.18: Measured Instantaneous Voltage and Current Signals Incident No.3

Figure 3.19: Calculated Line Positive Sequence Voltage and Current Incident No.3

77
Incident No.4:

This incident has resulted in single phase voltage dip of 250 ms duration. Figure 3.20
shows the measured instantaneous three phase voltage and current signals and
Figure 3.21 shows the calculated positive sequence voltage and current.

Figure 3.20: Measured Instantaneous Voltage and Current Signals Incident No.4

Figure 3.21: Calculated Line Positive Sequence Voltage and Current Incident No.3

In order to improve the estimation process, the parameter estimation task was
initiated with typical induction motor parameters. The search interval for each
parameter was selected carefully based on the physical characteristic of each

78
parameter. Our objective is to identify a set of induction motor parameters that is
capable of representing the aggregated induction motor load in the system. That
means the difference between the estimated load parameters for each incident from
the other should result from the different ZIP Load Model (static load) parameters and
the load composition parameters (percentage of induction motors), i.e. the estimated
induction motor parameters remain unchanged when the load changes. Therefore,
parameter estimation was repeated several times for the selected incidents. After
several trials, the static load parameters and the percentage of induction motor were
identified for each incident and carried out for the remaining of the parameter
estimation task. After identifying the static load parameters and the percentage of
induction motors for each incident, the induction motor parameters were identified by
starting the estimation of an incident using the induction motor parameters found from
the previous estimation trial of another incident, until the difference between the
estimated induction motor parameters for each incident became negligible. The
estimated versus measured active and reactive power for Incident No.1, Incident
No.2, Incident No.3 and Incident No.4 are shown in Figure 3.22 through Figure 3.25
respectively. Table 3.7 shows the estimated load composition and ZIP Load Model
parameters for each Incident and Table 3.8 shows the estimated unified aggregate
motor load parameters.

Figure 3.22: Measured and Estimated Active and Reactive Power Incident No.1

79
Figure 3.23: Measured and Estimated Active and Reactive Power Incident No.2

Figure 3.24: Measured and Estimated Active and Reactive Power Incident No.3

Figure 3.25: Measured and Estimated Active and Reactive Power Incident No.4

80
Table 3.7: Estimated Load Composition and ZIP Load Model Parameters for each Incident
Incident Name
Estimated Parameters
Incident No.1 Incident No.2 Incident No.3 Incident No.4

Load Composition
0.28 0.31 0.41 0.35
0.35 0.47 0.46 0.37
0.72 0.69 0.59 0.65
0.65 0.53 0.54 0.63
ZIP Load model
0.1 0.09 0 0.01
0.05 0.03 0 0.03
0.85 0.88 1 0.96
0.92 0.78 0.5 0.8
0.04 0.2 0.20 0.03
0.04 0.02 0.3 0.17

Table 3.8: Estimated Unified Aggregate Motor Load Parameters

Rs () Rr () Xls () Xlr () Xm () H A B C

1.1154 0.4586 0.7074 0.0253 42.0565 7.2534 1.4954 x 10-6 0.0019 0.9981

The Torque, Speed and Currents of the estimated aggregated motor parameters were
plotted for a step load torque input to ensure that these parameters represent an
aggregate motor load as shown in Figure 3.26 through Figure 3.28.

Figure 3.26: Aggregated Motor Currents

81
Figure 3.27: Aggregated Motor Torque

Figure 3.28: Aggregated Motor Speed

3.3.2.3. Validation of the Aggregate Motor Load Parameters

To check the validity of the estimated unified aggregate motor load parameters
presented in Table 3.8 for different incidents other than those used during parameter
estimation, the parameters were tested for randomly selected incident. Sample test
incident is given in Figure 3.29, which shows the measured instantaneous three phase
voltage and current signals. Figure 3.30 shows the calculated positive sequence
voltage and current. Figure 3.31 shows the measured versus simulated active and
reactive power using the estimated unified aggregate motor load parameters presented
in Table 3.8. Only the load composition and static load model parameters were
slightly changed to match the steady state (pre-incident) conditions. It is found that
the estimated unified aggregate motor load parameters have a good generalization

82
capability, i.e., that they are not incident-dependent. Therefore, these parameters can
be used to represent the aggregated motor loads for Dubai Power Grid. However, it is
recommended to test the validity and generalization capability of these parameters
periodically, for example once annually, in order to account for load variations and
network changes (especially at demand side) that take place in a year time.

Figure 3.29: Measured Instantaneous Voltage and Current Signals

Figure 3.30: Calculated Line Positive Sequence Voltage and Current

83
Figure 3.31: Measured and Simulated Active and Reactive Power

3.3.2.4. Identification of Load Tripping Rules

While analyzing the recorded incidents, it was observed that load tripping
takes place for positive sequence voltage as high as 90% of the nominal voltage for
dip duration of 150 ms and above. Two load tripping patterns were observed, Fast
Tripping and Slow Tripping. The two load tripping patterns were analyzed in order to
extract load tripping rules as follows:

1. Fast Tripping:
It was observed that in some of the recorded incidents, there is a reduction in the
active power immediately after the voltage recovers to its nominal value, then the
active power stabilizes at the new value indicating that portion of the load tripped in a
fast manner as illustrated in Figure 3.32 and Figure 3.33.

84
Figure 3.32: Fast Tripping Pattern-Ilustration-1

Figure 3.33: Fast Tripping Pattern-Ilustration-2

The amount of tripped load of the fast pattern was correlated to the positive voltage
dip. The dip duration was ignored, since for the majority of incidents, the dip duration
was around 300 ms. The amount of tripped load was plotted against the voltage dip,
then, three fast tripping schemes were deduced, the first is linear, the second is
exponential and the third is staircase shape as shown in Figure 3.34. Table 3.9 shows
the details of the three extracted Load Tripping Schemes of the Fast Tripping Pattern.

85
The proper tripping scheme shall be selected through the load model validation
process later in section 3.2.3.

Figure 3.34: Different Tripping Schemes (Fast Tripping Pattern)

Table 3.9: Extracted Load Tripping Scheme (Fast Tripping Pattern)


Voltage Sensing Tripping Load Tripping Fraction (%)
Stage Threshold Time Time
(%) (ms) (ms) Linear Exponential Staircase
1 90% 50 200 3.0% 2.0% 5.0%
2 80% 50 200 7.0% 3.0% 10.0%
3 70% 50 200 11.0% 6.0% 15.0%
4 60% 50 200 15.0% 11.0% 20.0%
5 50% 50 200 19.0% 23.0% 25.0%

1. Slow Tripping:
In some of the recorded incidents another slower pattern of load tripping was
observed. There is immediate active power reduction after the voltage recovers to its
nominal value like in the previous incidents. However, the active power does not
stabilize; it rather continues to reduce in linear slow manner indicating slow form of
load tripping as illustrated in Figure 3.35 and Figure 3.36. Figure 3.37 shows the slow

86
linear tripping pattern during 10 seconds time extracted from the available
measurements with respect to the triggering voltage dip magnitude (p.u). The average
curve was implemented in this thesis.

Figure 3.35: Slow Tripping Pattern-Ilustration-1

Figure 3.36: Slow Tripping Pattern-Ilustration-2

87
Figure 3.37: Extracted Slow Tripping Pattern related to the Triggering Voltage Dip

3.3.3 Validation of the Developed Load Model

In any modeling task, irrespective of the modeling approach, a model


verification and validation should be performed to ensure its accuracy [11]. Now,
after identifying the load composition, aggregated load model parameters and fast and
slow load tripping schemes, the updated load model of Dubai Power Grid is ready for
validation. For this purpose, the model shall be validated against two recorded global
system disturbances (i.e., seen by the whole system). The first disturbance is the
aforementioned single phase to ground fault encountered at the 400 kV bus of
Mushrif (MUSH) 400/132 kV load substation on the 20th of October 2009 which
resulted in tripping 8% of the total supplied load. The second disturbance is a single
phase to ground fault at the 400 kV bus of Warsan (WRSN) 400/132 kV generation
and load substation on the 21st of November 2012 and resulted in tripping 10% of the
supplied load.

PSS/E software package was used for validating the updated load model
against these two recorded disturbances. For each disturbance, PSS/E load flow
model of Dubai Power Grid was tuned to reflect the recorded system conditions
during these two disturbances (including Network topology, total generation, total
load, etc). For each of the two disturbances, the system events, including fault

88
occurrence, fault clearance, tripping of power system components, etc., were
simulated dynamically in their chronological order. The simulated voltage trends
were compared with the measured ones. Also, the total amount of tripped load during
the simulation was monitored for the sake of comparison with the recorded figure.
Mushrif disturbance was simulated using the three fast tripping schemes deduced
earlier in section 3.3.1.1, namely linear, exponential and staircase tripping schemes, in
order to select the best one among the three. The selected tripping scheme was
carried out for Warsan disturbance.

3.3.3.1. Validation Dubai Power Grid Updated Model against Mushrif


Disturbance

This disturbance was initiated by a single phase to ground fault encountered at


the 400 kV bus of Mushrif (MUSH) 400/132 kV load substation. The fault was
cleared within 60 ms (which is within the normal fault clearance time criteria of
Dubai Power Grid). The fault clearance was accompanied by tripping one 400 kV
feeder circuit (out of three) at MUSH substation bus and three 400/132 kV
transformers (out of four) at MUSH substation. The disturbance had resulted in
tripping 8% of the total supplied load. This disturbance was dynamically simulated
using the updated load model and the simulated voltage trend at MUSH 132 kV bus
was compared with the recorded one as shown in Figure 3.38. It is obvious from
Figure 3.38 that the fast linear load tripping schemes give the best results for voltage
trends. The results of the fast linear load ripping scheme were compared to the load
tripping scheme of the old load model as shown in Figure 3.39. It is clear that the
new load model is much more accurate than the old load model. Figure 3.40 shows
the total tripped load during simulation of this disturbance using the fast linear load
tripping scheme is around 9% which is very close to actual figures.

89
Figure 3.38: Recorded vs. Simulated Voltage Trend at MUSH 132 kV bus MUSH Disturbance
(Different Fast Load Tripping Scheme)

Figure 3.39: Recorded vs. Simulated Voltage Trend at MUSH 132 kV bus MUSH Disturbance
(New and Old Load Model)

90
Figure 3.40: Simulated Total System Load during MUSH Disturbance (New Load Model)

3.3.3.2. Validation Dubai Power Grid Updated Model against Warsan


Disturbance

This disturbance was initiated by a single phase to ground fault encountered at


the 400 kV bus of Warsan (WRSN) 400/132 kV load substation. The fault was
cleared within 66 ms (which is within the normal fault clearance time criteria of
Dubai Power Grid). The fault clearance was accompanied by tripping one 400 kV
feeder circuit (out of nine) at WRSN substation. The disturbance had resulted in
tripping 10% of the total supplied load. This disturbance was dynamically simulated
using the updated load model and the simulated voltage trend at MUSH 132 kV bus
was compared to the recorded one as shown in Figure 3.41 and the simulated voltage
trend at BKRA 400 kV bus was compared with the recorded one as shown in
Figure 3.42. Figure 3.43 shows that the total load lost during simulation of this
disturbance using the fast linear load tripping scheme is around 10% which is very
close to reality.

The validation effort of the updated Dubai Power Grid Load Model against the
recorded global system disturbances shows that the model is capable, with high
accuracy, of replicating the system behavior (especially voltage trends) during the
recorded system disturbances. Therefore, this model shall be used in the next chapter
to perform voltage stability assessment of Dubai Power Grid.

91
Figure 3.41: Recorded vs. Simulated Voltage Trend at MUSH 132 kV bus WRSN Disturbance

Figure 3.42: Recorded vs. Simulated Voltage Trend at BKRA 400 kV bus - WRSN Disturbance

92
Figure 3.43: Simulated Total System Load during WRSN Disturbance

93
Chapter 4: Voltage Stability Assessment of Dubai Power Grid

4.1 Overview of Dubai Power Grid Planning Standards

The Dubai Electricity and Water Authority (DEWA) meets its responsibility to
supply electricity demanded by its customers with a high degree of reliability. This is
being done through the carefully planned development of power generating sources,
transmission, and distribution systems. A reliable supply of electricity involves two
elements adequacy and security. Planning a reliable transmission system needs to be
done taking into account adequate reliability/continuity of supply and due
consideration to the various constraints (environmental, operational, regulatory, etc.),
all in line with the best international standards and practices. This involves several
assessment processes including evaluation of existing system capabilities and
identification of system deficiencies. Analysis of system performance should include
thermal, voltage, short circuit, stability and frequency assessments [41]. The focus of
this thesis is voltage stability assessment due to the reasons mentioned earlier in
Section 3.2.

The 400 kV Network is the backbone of the system and N-2 on-line reliability
criterion is used. This means that the sudden loss of two 400 kV circuits (even during
summer peak) should not result in system collapse nor lead to a widespread loss of
power supply to consumers. The Expanded Planning Criteria for Dubai Power Grid
was prepared based on North American Electric Reliability Council NERC
Planning Standard Document. These criteria define the reliability of Dubai Power
Grid using the following two terms:

Adequacy: The ability of the electric systems to supply the aggregate electrical
demand and energy requirements of their customers at all times, taking into
account scheduled and reasonably expected unscheduled outages of system
elements.
Security: The ability of the electric systems to withstand sudden disturbances
such as electric short circuits or unanticipated loss of system elements.

The main objective of the Expanded Planning Criteria is to maintain Dubai Power
Grid within the normal state. Examples of system conditions that could cause
departure from the normal state are: capacity deficiencies, energy deficiencies, loss of

94
generation or transmission facilities, transmission facility overloads and high or low
voltages and abnormal power system frequency. Dubai Power Grid is being planned
with a sufficient transmission capability to withstand the loss of specified,
representative and reasonably foreseeable contingencies at projected customer
demand and anticipated power transfer levels. The application of these credible
contingency planning criteria should not result in any criteria violations, or the loss of
a major portion of the system, or unintentional separation of a major portion of the
system. The planning criteria, however, recognize that extreme non-credible events
might also occur. These events are considered to be high impact low probability
(HILP) events and, therefore, require appropriate assessment [41].

Voltage stability of Dubai Power Grid shall be maintained at all times


according to the following criteria:

a. Pre-Contingency Voltage Criteria

For both normal and emergency transfers, no bus voltage shall be below its pre-
contingency low voltage limit nor be above its pre-contingency high voltage limit.

b. Post-Contingency Voltage Criteria

No bus voltage shall fall below its post-contingency low voltage limit nor rise above
its post-contingency high voltage limit.

Table 4.1 shows the steady state voltage levels for 400 kV and 132 kV levels under
normal and contingency operational conditions.

Table 4.1: Steady State Voltage Levels for 400 kV and 132 kV levels
(Normal and Contingency Operational Conditions)

Normal Operation Contingency


Nominal
Voltage Min > Max< Min > Max <

~kV ~% ~kV ~% ~kV ~% ~kV ~%

400 kV 380 -5% 420 +5% 360 -10% 420 +5%

132 kV 125 -5% 138 +5% 120 -10% 145 +10%

95
4.2 Voltage Stability Assessment Methodology

The comprehensive methodology for voltage stability assessment of power


systems using modern analytical tools addressed in [8] was implemented to assess the
voltage stability of Dubai Power Grid. The methodology endorses performing both
steady state and dynamic voltage stability analyses, with high emphasis that the latter
becomes very critical especially for a power system with a significant share of motor
loads. As mentioned earlier, dynamic voltage stability analysis describes the time-
dependent voltage performance of the system by revealing the transient and/or the
longer-term voltage stability of a power system under study following system
disturbances. In this thesis, the time frame of interest is 10 seconds after the inception
of the disturbance, since this period is enough to capture the induction motor
dynamics, which are the dominant load type of Dubai Power Grid. Therefore, the use
of dynamic voltage stability analysis term denotes short-term or transient voltage
stability analysis. Long term voltage stability analysis is not addressed.

4.2.1 Software Tool

PSS/E Software Package was used to conduct both steady state and dynamic
voltage stability analysis. PSS/E Software Package is comprised of a comprehensive
suite of programs for studies of power system transmission network and generation
performance in both steady state and dynamic conditions [10].

4.2.2 Load Model Representation

For steady state voltage stability analysis, short term transients and dynamics
can be ignored; therefore, static load models (such as exponential and polynomial or
ZIP load models) are exclusively employed as their effects dominate during a voltage
disturbance under study. As explained earlier in Section (2.6.1), static load model may
consist of a combination of three load components: Constant Impedance, Constant
Current and Constant Power loads. Constant Power loads maintain a constant power
draw from the system regardless the change in voltage. Therefore, constant power
loads tend to aggravate a voltage collapse condition. On the other hand, the power
drawn by constant current and constant impedance loads decreases with voltage drop
providing load relief and hence, better voltage recovery [9]. In this thesis, for steady

96
state voltage stability analysis constant power load is assumed, at all load buses, as a
worst case scenario. For dynamic voltage stability analysis, the detailed load model
developed in Chapter-3 is implemented to incorporate the impact of the large
proportion of induction motors and reveal if there is any associated transient or quasi-
steady state voltage problems that were not captured by steady state voltage analysis.

4.2.3 Study Considerations and Scenarios

Steady state and dynamic voltage stability analyses were performed for Dubai
Power Grid considering Year 2014 Network Topology and Forecasted Peak Load
Conditions. Figure B.1 (Appendix B) shows the planned 400 kV network topology of
Dubai Power Grid for year 2014. The available generation capacity for year 2014 is
around 9650 MW and the forecasted load 6995 MW. Four main load substations
(400/132 kV level) were selected for voltage stability analysis as these substations are
most likely the weakest from the voltage stability point of view due to the following
reasons:

Two of the four substations are currently feeding heavy load compared to
other load substations in the system. The other two are expected to feed heavy
load in the future.
The four substations are located in highly loaded and developing area that is
Deira Side of Dubai City.
The remoteness of these four substations from the main generation complex
in Jebel Ali. Figure B.2 (Appendix B) shows the location of the four
substations on Dubai City Map.

These substations are Nahda (NHDA) 400/132 kV substation, Mushrif (MUSH)


400/132 kV substation, Car Complex (CARX) 400/132 kV substation and Mamzar
Beach (MBCH) 400/132 kV substation. Typical 400/132 kV substation layout is
shown in Figure 4.1.

97
Figure 4.1: Typical Substation Layout a 400/132 kV Substation

Various disturbance scenarios were selected for assessment according to the


contingency criteria for credible and secured events identified in Dubai Power Grid
Expanded Planning Criteria. These disturbances cover:

N-1 contingencies including loss of one circuit or loss of one transformer as


shown in Figure 4.2.
N-2 contingencies including loss of two circuits, or loss of one transformer and
one circuit (equivalent to the loss of one bus section) as shown in Figure 4.3.
N-3 contingencies including loss of two circuits and one transformer as shown in
Figure 4.4.

A comprehensive list of all the above contingency types was prepared for each
substation. The total number of contingencies of all the three categories for each
substation ranges from 13 to 21 contingencies. The comprehensive contingency list
for each substation is presented in Tables B.1, B.2, B.3 and B.4 (Appendix B) for
NHDA, MUSH, CARX and MBCH respectively. For steady state analysis, the
analysis was performed for each substation using the full contingency list. The steady

98
state analysis results were used to identify the worst contingency of each contingency
category for each substation i.e. three contingency scenarios for each substation.
Flowingly, the dynamic analysis was performed using the identified worst
contingencies for each substation.

Figure 4.2: Example of N-1 Contingencies

Figure 4.3: Example of N-2 Contingencies

Figure 4.4: Example of N-3 Contingencies

99
4.2.4 Steady State Voltage Analysis Results

The steady-state voltage stability analysis was performed based on the criteria
established by Western Electricity and Coordinating Council (WECC) of North
America. The document that was used as a guide for this study was approved by
WECC in May of 1998 [46]. The WECC voltage stability criteria are specified in
terms of active and reactive power margins, as shown in Table B.5 (Appendix B).
PSS/E software package was used for performing PV and VQ analysis.

The active power transfer limit for each contingency scenario is calculated
from the nose of the corresponding PV curve. Then, for each contingency category
the active power transfer limit of the worst contingency among this category is
considered as the maximum active power transfer margin. According to WECC
voltage stability criteria, a safety margin of 5% is taken for N-0 and N-1 contingency
conditions and 2.5% for N-2 and N-3 contingency conditions. Figure 4.2 illustrates
the process of calculating the maximum active power transfer margin.

Figure 4.5: Calculating Active Power Transfer Margin from PV Curves

The Reactive Power Reserve Margin (RRM) for each contingency scenario is
calculated from the knee of the corresponding QV curve. For each contingency
scenario, the reactive power margin for the base load and a load increase of 5% is
calculated. The change in reactive power margin between base load and base load
+5% is calculated, and the contingency having the maximum margin change is
considered the worst contingency of its category. This means, under this contingency,

100
the system is very sensitive to load change, i.e. prone to voltage instability. This
process is illustrated in Figure 4.3.

Figure 4.6: Calculating Reactive Power Reserve Margin from VQ Curves

4.2.4.1 PV Analysis Results

PV curves for NHDA, MUSH, CARX and MBCH 400/132 kV substation are
plotted for intact network condition (Base Case or N-0) as well as the comprehensive
list of N-1, N-2 and N-3 contingencies addressed in Tables B.1, B.2, B.3 and B.4
respectively (Appendix B). PV analysis of NHDA 400/132 kV substation is
presented in details. For MUSH, CARX, and MBCH, the results are summarized in
this section and the remaining curves and tables are provided in Appendix C.

Figure 4.4 shows all the PV curves plotted for NHDA 400/132 kV substation
covering the Base Case (N-0) and all the studied N-1, N-2 and N-3 contingency
scenarios. For comparison purposes and in order to identify the worst contingency
among each category, Figure 4.5 segregates the PV curves plotted of all N-1
contingencies, Figure 4.6 segregates the PV curves plotted of all N-2 contingencies,
and Figure 4.7 segregates the PV curves plotted of all N-3 contingencies. The worst
contingency of each category is the one that corresponds to the lowest active power
transfer among the others. After identifying the worst contingency of each category,
the corresponding PV curves are segregated with the PV curve of Base Case (N-0) in
Figure 4.8. The normal loading point and the substation firm capacity are shown in

101
Figure 4.8. The steady state voltage limits for normal operation and contingency
conditions are also displayed.

V (p.u.)
1.10
BASE CASE
CONTINGENCY-1
CONTINGENCY-2
CONTINGENCY-3
1.05 CONTINGENCY-4
CONTINGENCY-5
CONTINGENCY-6
CONTINGENCY-7
CONTINGENCY-8
1.00
CONTINGENCY-9
CONTINGENCY-10
CONTINGENCY-11
CONTINGENCY-12
0.95 CONTINGENCY-13
CONTINGENCY-14
CONTINGENCY-15
CONTINGENCY-16
CONTINGENCY-17
0.90 CONTINGENCY-18
CONTINGENCY-19
CONTINGENCY-20
CONTINGENCY-21
0.85
500 700 900 1100 1300 1500 1700 Power Transfer (MW)

Figure 4.7: PV Curves for NHDA 400/132 kV Substation Base Case and All Contingencies

V (p.u.)

1.10

1.05

CONTINGENCY-1
1.00
CONTINGENCY-2
CONTINGENCY-3
CONTINGENCY-4
0.95
CONTINGENCY-5

0.90

0.85
500 700 900 1100 1300 1500 1700 Power Transfer (MW)

Figure 4.8: PV Curves for NHDA 400/132 kV Substation All N-1 Contingencies

102
V (p.u.)

1.10

CONTINGENCY-6
1.05
CONTINGENCY-7

CONTINGENCY-8

1.00 CONTINGENCY-9

CONTINGENCY-10

CONTINGENCY-11

0.95 CONTINGENCY-12

CONTINGENCY-13

CONTINGENCY-14
0.90 CONTINGENCY-15

0.85
500 700 900 1100 1300 1500 1700 Power Transfer (MW)

Figure 4.9: PV Curves for NHDA 400/132 kV Substation All N-2 Contingencies

V (p.u.)

1.10

1.05
CONTINGENCY-16
CONTINGENCY-17
CONTINGENCY-18
1.00
CONTINGENCY-19
CONTINGENCY-20
CONTINGENCY-21
0.95

0.90

0.85
500 700 900 1100 1300 1500 1700 Power Transfer (MW)

Figure 4.10: PV Curves for NHDA 400/132 kV Substation All N-3 Contingencies

103
V (p.u.)

1.10
Normal Loading (Year 2014) Substation Firm Capacity

High Voltage Limit


1.05 (Normal &
Contingency)

1.00

Low Voltage Limit


0.95
(Normal)

Low Voltage
0.90 Limit
(Contingency)

0.85
500 600 700 800 900 1000 1100 1200 1300 1400 1500 1600 Power Transfer (MW)

Base Case (N-0) Worst N-1 Contingency


Worst N-2 Contingency Worst N-3 Contingency

Figure 4.11: PV Curves for NHDA 400/132 kV Substation Worst Contingencies

It is obvious from Figure 4.8 that all PV curves for Base Case and the worst
contingencies fall within the acceptable voltage limits. It is also found that NHDA
400/132 kV substation can be loaded to near its firm capacity during the worst
contingency scenario. Table 4.2 presents the calculated maximum active power
transfer margin for the Base Case and all the contingency scenarios. The worst
contingency of each category, highlighted in yellow, is considered as the limiting
contingency of its category, and hence, the corresponding active power transfer limit
is considered the maximum active power transfer margin for this category. The
figures shown are after implementing the 5% safety margin for N-0 and N-1
contingencies, and 2.5% safety margin for N-2 and N-3 contingencies.

104
Table 4.2: Maximum Active Power Transfer Margin for NHDA 400/132 kV Substation Base
Case and All Contingencies
Active Power Maximum Active
Contingency Identifier Category Transfer Limit Power Transfer
(MW) Margin (MW)
BASE CASE N-0 1519 1443
CONTINGENCY-1 1138
CONTINGENCY-2 1256
CONTINGENCY-3 N-1 1468 1081
CONTINGENCY-4 1431
CONTINGENCY-5 1514
CONTINGENCY-6 1138
CONTINGENCY-7 1181
CONTINGENCY-8 1213
CONTINGENCY-9 1425
CONTINGENCY-10 1425
N-2 1060
CONTINGENCY-11 1425
CONTINGENCY-12 1088
CONTINGENCY-13 1131
CONTINGENCY-14 1131
CONTINGENCY-15 1125
CONTINGENCY-16 1044
CONTINGENCY-17 1056
CONTINGENCY-18 1069
N-3 1018
CONTINGENCY-19 1119
CONTINGENCY-20 1113
CONTINGENCY-21 1138

Similarly, PV analyses were performed for MUSH, CARX and MBCH


400/132 kV substation. PV Curves corresponding to the worst contingencies are
shown in Figure 4.9, Figure 4.10 and Figure 4.11 for MUSH, CARX and MBCH
400/132 respectively. Table 4.3 summarizes PV analyses results for the four
substations. It was found that NHDA and MUSH 400/132 kV substation can be
loaded to near their firm capacity during the worst contingency scenario. However,
for CARX and MBCH 400/132 kV substation the maximum active power transfer
margin of the worst contingency is approximately equivalent to the load forecast of
year 2023. This means that detailed analyses of these two substations are required
before 2023.

105
V (p.u.) Normal Loading (Year 2014) Substation Firm Capacity
1.10

High Voltage Limit


1.05 (Normal &
Contingency)

1.00

Low Voltage Limit


0.95
(Normal)

Low Voltage Limit


0.90
(Contingency)

0.85
500 600 700 800 900 1000 1100 1200 1300 1400 1500 Power Transfer (MW)
Base Case (N-0) Worst N-1 Contingency
Worst N-2 Contingency Worst N-3 Contingency

Figure 4.12: PV Curves for MUSH 400/132 kV Substation Worst Contingencies

V (p.u.)
1.10
Normal Loading (Year 2014) Substation Firm Capacity

High Voltage Limit


1.05 (Normal &
Contingency)

1.00

Low Voltage Limit


0.95
(Normal)

Low Voltage Limit


0.90 (Contingency)

0.85
200 300 400 500 600 700 800 900 1000 1100 1200 1300 1400 Power Transfer (MW)
Base Case (N-0) Worst N-1 Contingency
Worst N-2 Contingency Worst N-3 Contingency

Figure 4.13: PV Curves for CARX 400/132 kV Substation Worst Contingencies

106
V (p.u.)
1.10
Normal Loading (Year 2014) Substation Firm Capacity

High Voltage Limit


1.05 (Normal &
Contingency)

1.00

Low Voltage Limit


0.95
(Normal)

Low Voltage Limit


0.90
(Contingency)

0.85
200 300 400 500 600 700 800 900 1000 1100 1200 1300 1400 Power Transfer (MW)
Base Case (N-0) Worst N-1 Contingency
Worst N-2 Contingency Worst N-3 Contingency

Figure 4.14: PV Curves for MBCH 400/132 kV Substation Worst Contingencies

Table 4.3: Maximum Active Power Transfer Margin for NHDA, MUSH, CARX and MBCH
400/132 kV Substation Base Case and Worst Contingencies

Maximum Power
Contingency
Substation Transfer Margin Remarks
Type
(MW)
Base Case (N-0) 1443
Worst N-1 1081
NHDA Approaching Firm Capacity
Worst N-2 1060
Worst N-3 1018
Base Case (N-0) 1358
Worst N-1 1186
MUSH Approaching Firm Capacity
Worst N-2 1199
Worst N-3 1168
Base Case (N-0) 849
Worst N-1 707 Approaching Year 2023
CARX
Worst N-2 713 Forecasted Load
Worst N-3 695
Base Case (N-0) 785
Worst N-1 642 Approaching Year 2023
MBCH
Worst N-2 610 Forecasted Load
Worst N-3 549

107
4.2.4.2 VQ Analysis Results

VQ curves for NHDA, MUSH, CARX and MBCH 400/132 kV substation are
plotted for intact network condition (Base Case or N-0) as well as the comprehensive
list of N-1, N-2 and N-3 contingencies addressed in Tables B.1, B.2, B.3 and B.4
respectively (Appendix B). VQ analysis of NHDA 400/132 kV substation is presented
in details. For MUSH, CARX, and MBCH, the results are summarized in this section
and the remaining curves and tables are provided in Appendix C.

Figure 4.12 shows all the QV curves plotted for NHDA 400/132 kV substation
covering the Base Case (N-0) all the studied N-1, N-2 and N-3 contingency scenarios.
For comparison purposes and to identify the worst contingency among each category,
Figure 4.13 segregates the QV curves plotted of all N-1 contingencies, Figure 4.14
segregates the QV curves plotted of all N-2 contingencies, and Figure 4.15 segregates
the VQ curves plotted of all N-3 contingencies. After identifying the worst
contingency of each category, the corresponding QV curves are segregated with the
QV curve of Base Case (N-0) in Figure 4.16.

Q (MVAr)
1500
BASE CASE
CONTINGENCY-1
CONTINGENCY-2
1000 CONTINGENCY-3
CONTINGENCY-4
CONTINGENCY-5
CONTINGENCY-6
500
CONTINGENCY-7
CONTINGENCY-8
CONTINGENCY-9
0 CONTINGENCY-10
CONTINGENCY-11
CONTINGENCY-12
CONTINGENCY-13
-500 CONTINGENCY-14
CONTINGENCY-15
CONTINGENCY-16
-1000 CONTINGENCY-17
CONTINGENCY-18
CONTINGENCY-19
CONTINGENCY-20
-1500 CONTINGENCY-21
0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20
V(p.u.)

Figure 4.15: QV Curves for NHDA 400/132 kV Substation Base Case and All Contingencies

108
Q (MVAr)
1500

1000

500
CONTINGENCY-1
CONTINGENCY-2
0 CONTINGENCY-3
CONTINGENCY-4
CONTINGENCY-5
-500

-1000

-1500
0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 V(p.u.)

Figure 4.16: QV Curves for NHDA 400/132 kV Substation All N-1 Contingencies

Q (MVAr)
1500

1000

CONTINGENCY-16
500 CONTINGENCY-17
CONTINGENCY-18
CONTINGENCY-19
0 CONTINGENCY-20
CONTINGENCY-21

-500

-1000
0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 V(p.u.)

Figure 4.17: QV Curves for NHDA 400/132 kV Substation All N-2 Contingencies

Q (MVAr)
1500

1000

CONTINGENCY-6
500 CONTINGENCY-7
CONTINGENCY-8
CONTINGENCY-9
0
CONTINGENCY-10
CONTINGENCY-11
-500 CONTINGENCY-12
CONTINGENCY-13
CONTINGENCY-14
-1000 CONTINGENCY-15

-1500
0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 V(p.u.)

Figure 4.18: QV Curves for NHDA 400/132 kV Substation All N-3 Contingencies

109
Q (MVAr)
1500

1000

500

-500

-1000

-1500
0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 V(p.u.)

Base Case (N-0) Worst N-1 Contingency


Worst N-2 Contingency Worst N-3 Contingency

Figure 4.19: QV Curves for NHDA 400/132 kV Substation Worst Contingencies

The curves are plotted for base load and a load increase of 5% in order to calculate the
reactive power reserve margin for Base Case (N-0) and all contingency scenarios as
shown in Table 4.4. The worst contingency of each category, highlighted in yellow,
is the ones with the highest change in reactive power reserve margin with respect to a
load increment of 5%. Hence, the corresponding reactive power reserve margin (for
base load +5% situations) is considered as the available reactive power margin of
each contingency category. It is found that the calculated reactive power reserve
margin for each category is compliant with WECC voltage stability criteria addressed
in Table B.5 (Appendix B).

110
Table 4.4: Available Reactive Power Reserve Margin for NHDA 400/132 kV Substation Base
Case and All Contingencies
Reactive Power
Reserve Margin
Change Available
(RRM) (MVAr)
Contingency Identifier Category in RRM RRM
Base (MVAr) (MVAr)
Base
Load
Load
+5%
BASE CASE N-0 1012 971 42 971
CONTINGENCY-1 931 865 66
CONTINGENCY-2 921 871 50
CONTINGENCY-3 N-1 943 901 42 865
CONTINGENCY-4 930 888 42
CONTINGENCY-5 981 939 42
CONTINGENCY-6 775 712 62
CONTINGENCY-7 759 695 64
CONTINGENCY-8 794 744 49
CONTINGENCY-9 904 862 42
CONTINGENCY-10 894 852 42
N-2 726
CONTINGENCY-11 844 801 43
CONTINGENCY-12 814 726 88
CONTINGENCY-13 862 797 66
CONTINGENCY-14 900 834 66
CONTINGENCY-15 849 783 66
CONTINGENCY-16 652 509 144
CONTINGENCY-17 637 514 122
CONTINGENCY-18 687 594 93
N-3 509
CONTINGENCY-19 823 757 66
CONTINGENCY-20 813 747 66
CONTINGENCY-21 763 696 67

Similarly, VQ analyses were performed for MUSH, CARX and MBCH


400/132 kV substation. VQ Curves corresponding to the worst contingencies are
shown in Figure 4.17, Figure 4.18 and Figure 4.19 for MUSH, CARX and MBCH
400/132 respectively. Table 4.5 summarizes VQ analyses results for the four
substations. It is found that the calculated reactive power reserve margin for each
contingency category is compliant with WECC voltage stability criteria addressed in
Table B.5 (Appendix B).

111
Q (MVAr)
2000

1500

1000

500

-500

-1000

-1500
0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 V(p.u.)

Base Case (N-0) Worst N-1 Contingency


Worst N-2 Contingency Worst N-3 Contingency

Figure 4.20: VQ Curves for MUSH 400/132 kV Substation Worst Contingencies

Q (MVAr)
1500

1000

500

-500

-1000

-1500
0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 V(p.u.)

Base Case (N-0) Worst N-1 Contingency


Worst N-2 Contingency Worst N-3 Contingency

Figure 4.21: VQ Curves for CARX 400/132 kV Substation Worst Contingencies

112
Q (MVAr)

1500

1000

500

-500

-1000

-1500
0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 V(p.u.)

Base Case (N-0) Worst N-1 Contingency


Worst N-2 Contingency Worst N-3 Contingency

Figure 4.22: VQ Curves for MBCH 400/132 kV Substation Worst Contingencies

Table 4.5: Available Reactive Power Reserve Margin for NHDA, MUSH, CARX and MBCH
400/132 kV Substation Base Case and All Contingencies
Available
Contingency
Substation RRM WECC Criteria
Type
(MVAr)
Base Case (N-0) 971 Reference RRM
Worst N-1 865 > 50% of RRM of N-0
NHDA Compliant
Worst N-2 726 > 50% of RRM of N-0
Worst N-3 509 >0
Base Case (N-0) 961 Reference RRM
Worst N-1 796 > 50% of RRM of N-0
MUSH Compliant
Worst N-2 637 > 50% of RRM of N-0
Worst N-3 517 >0
Base Case (N-0) 1002 Reference RRM
Worst N-1 710 > 50% of RRM of N-0
CARX Compliant
Worst N-2 683 > 50% of RRM of N-0
Worst N-3 553 >0
Base Case (N-0) 977 Reference RRM
Worst N-1 769 > 50% of RRM of N-0
MBCH Compliant
Worst N-2 596 > 50% of RRM of N-0
Worst N-3 512 >0

113
4.2.5 Dynamic Voltage Analysis Results

Transient dynamic voltage analysis was performed for each of the four
selected substations. The studied disturbance scenarios reflected the worst case
contingencies of each contingency category (N-1, N-2 and N-3) revealed from the
steady state analysis. The disturbances were simulated to reflect the worst
contingencies considering practical scenarios as illustrated in Table 4.6. The different
fault locations are illustrated in .The simulated fault is a solid single phase to ground
fault. Three phase fault is not simulated due to the very low occurrence probability.
During the past 23 years, Dubai Power Grid has not encountered any Three Phase
Fault on the high voltage network (400 and 132 kV levels). Moreover, the load model
was not validated for three phase faults due to unavailability of records of such
incidents.

Table 4.6: Details of Simulated Disturbances


Simulated Disturbance Scenario
Contingency Total Fault
Category Fault Location Protection Action Tripped Equipment Clearance
Time

One Transformer One Transformer


Normal Fault
N-1 Or Or 100ms
Clearance
One 400 kV Circuit One 400 kV Circuit

Two 400 kV circuits on


Two 400 kV Circuits
the same tower
Or Normal Fault Or
Bus Section 100 ms
Clearance
Bus Section (Feeding (One 400 kV Circuit
Two Elements) +
N-2 One Transformer)
Breaker Protection
One Transformer Bus Section
Failure
(One 400 kV Circuit
Or Followed by 280 ms
+
Activation of One Transformer)
One 400 kV Circuit
Backup Protection
Bus Section
Bus Section (Feeding Normal Fault (Two 400 kV Circuit
100ms
Three Elements) Clearance +
One Transformer)
Breaker Protection
One 400 kV Circuit
N-3 Failure
Two 400 kV circuits on Followed by and
Bus Section 280 ms
the same tower
Activation of (One 400 kV Circuit
Backup Protection +
One Transformer)

114
Figure 4.23: Illustration of Different Fault Locations

PSS/E software package was used for simulating these disturbances. The
simulation starts with normal system operation for 1 second duration, then the fault is
applied at t = 1.0 second. The fault is then cleared after 100 ms (for normal fault
clearance time scenarios) or 280 ms (for breaker failure scenarios) as described in
Table 4.10. After fault clearance the elements outages according to the study
scenarios take place. The simulation then is carried out for 20 seconds to monitor
system behavior following the simulated disturbance. During the simulation, the
voltage trends at the 400 kV and 132 kV buses of the substation under study were
plotted to screen any oscillatory response, overvoltage and or voltage recovery
problem resulting from the applied disturbance. The voltage trends were plotted for
20 seconds duration. The results of dynamic voltage stability analysis for each of the
selected substations are presented thereafter.

115
The disturbances presented in Table 4.6 were simulated for all the four
selected 400/132 kV substation. For all the disturbances, the voltage trends at the 400
kV and the 132 kV buses of each substation are plotted. The voltage trends for NHDA
400/132 kV substation are shown in details, however for MUSH, CARX and MBCH
400/132 kV substations the voltage trends figures are provided in Appendix C.

Figure 4.20 shows the voltage trends at NHDA 400 kV Bus for the worst
contingencies for normal fault clearance time. Figure 4.21 shows the voltage trends at
NHDA 132 kV Bus for the worst contingencies for normal fault clearance time.
Figure 4.22 shows the voltage trends at NHDA 400 kV Bus for the worst
contingencies for backup protection clearance time and Figure 4.23 shows the voltage
trends at NHDA 132 kV Bus for the worst contingencies for backup protection
clearance time.

V(p.u.)
1.2
1.1
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0 5 10 15 20
Time (s)
V-NHDA-400 - Worst N-1 - Normal Fault ClearaNormal Fault Clearance Time
V-NHDA-400 - Worst N-2 - Normal Fault ClearaNormal Fault Clearance Time
V-NHDA-400 - Worst N-3 - Normal Fault ClearaNormal Fault Clearance Time

Figure 4.24: Voltage Trends at NHDA 400 kV Bus for the Worst Contingencies
(Normal Fault Clearance Time)

116
V(p.u.)
1.2
1.1
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0 5 10 15 20
Time (s)
V-NHDA-132 - Worst N-1 - Normal Fault ClearaNormal Fault Clearance Time
V-NHDA-132 - Worst N-2 - Normal Fault ClearaNormal Fault Clearance Time
V-NHDA-132 - Worst N-3 - Normal Fault ClearaNormal Fault Clearance Time
Figure 4.25: Voltage Trends at NHDA 132 kV Bus for the Worst Contingencies
(Normal Fault Clearance Time)

V(p.u.)
1.2
1.1
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0 5 10 15 20
Time (s)
V-NHDA-400 kV - Worst N-2 -Breaker Failure- Backup Protection Time
V-NHDA-400 kV - Worst N-3 -Breaker Failure- Backup Protection Time

Figure 4.26: Voltage Trends at NHDA 400 kV Bus for the Worst Contingencies
(Breaker Failure- Backup Protection Time)

117
V(p.u.)
1.2
1.1
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0 5 10 15 20
Time (s)
V-NHDA-132 kV - Worst N-2 -Breaker Failure- Backup Protection Time
V-NHDA-132 kV - Worst N-3 -Breaker Failure- Backup Protection Time

Figure 4.27: Voltage Trends at NHDA 132 kV Bus for the Worst Contingencies
(Breaker Failure- Backup Protection Time)

For all simulated single phase to ground faults cleared in normal fault
clearance times, there is a fast voltage recovery and a slight transient voltage rise at
NHDA 400 kV and 132 kV buses. However, the voltage returns to the acceptable
limit within 10 seconds. On the other hand, for all the simulated single phase to
ground faults cleared in backup protection time, the voltage recovery at NHDA 400
kV and 132 kV buses is relatively slow. However, the voltage returns to the
acceptable limit within maximum time duration of 10 seconds at NHDA 400 kV bus.
For NHDA 132 kV bus, although the voltage recovery is slow, the final voltage after
20 seconds is above acceptable limits for both N-2 and N-3 contingency scenarios.
The same voltage behavior is observed for MUSH, CARX and MBCH 400/132 kV
substations with slight differences.

In summary, for all the simulated contingencies resulting from single phase to
ground faults cleared in normal fault clearance times, there is a fast voltage recovery
at 400 kV and 132 kV buses. In some cases, there is a slight transient voltage rise (up
to 1.08 p.u. and 40 ms duration for the worst case). However, the voltage returns to
the acceptable limit within 10 seconds. On the other hand, for all the simulated
contingencies resulting from single phase to ground faults cleared in backup
protection time, the voltage recovery at 400 kV and 132 kV buses is relatively slow.

118
This is probably due to the extended fault duration which is sufficient to cause motor
stalling at the low voltage side. Though the voltage recovery is slow for extended
fault durations, the voltage returns to the acceptable limit within maximum time
duration of 10 seconds for 400 kV buses, and 20 seconds for few 132 kV buses. In
some cases, the final voltage after 20 seconds is above acceptable limits for both N-2
and N-3 contingency scenarios.

It is obvious that the updated load model revealed three problems related to
voltage stability, the first one, is the transient and/or quasi steady state voltage rise
that was encountered for normal cleared faults. The second problem is the slow
voltage recovery after extended fault durations and/or severe contingencies, which
was also evidenced in the slow voltage recovery during Mushrif incident in 2009. And
the third problem is the final voltage rise encountered following extended fault
durations and/or severe contingencies, this also was witnessed during Mushrif
incident in 2009. These three problems could be worsened with higher motor load
proportion, more severe contingencies, and more severe fault conditions. Moreover,
the dynamic simulation proved that steady state voltage stability analysis is not
sufficient to assess the voltage stability of a system with high proportion of motor
loads.

Based on the above, it is recommended to perform a comprehensive reactive


power assessment study for Dubai Power Grid, especially in view of the anticipated
high penetration of new load types including inverter based appliances. The study
may investigate the potential benefits of installing dynamic reactive power
compensation devices such as Static Var Compensators (SVCs).

119
Chapter 5: Conclusions, Recommendations and Future Work

5.1 Conclusions

Motivated by the increasing concern of power utilities (including Dubai


Electricity and Water Authority) in voltage stability problem and its potential adverse
impact on system reliability and security, this thesis presented the efforts performed to
assess the voltage stability of Dubai Power Grid against the increasing use of
transmission system resources, growing demand and associated stress on available
active and reactive power resources.

In order to investigate practical problems, as such, and propose appropriate


solutions, power system utilities depend on static and dynamic simulations. Power
system load characteristics have a major impact on voltage stability of power systems.
Load model uncertainty is recognized as the major source for simulation inaccuracy.
Accordingly, in any system stability study, especially voltage stability studies, it is
very important to have accurate load models capable of reflecting load behavior
during system disturbances. This enhances the ability of power system planners to
anticipate possible risks and design power systems more accurately.

The main two objectives of this thesis are:

To update Dubai Power Grid load model, to become capable of capturing load
behavior during system disturbances and enhance the accuracy of voltage
stability studies.
To evaluate the voltage performance of Dubai Power Grid against the
increasing use of transmission system resources, growing demand and
associated stress on available active and reactive power resources.

The first objective (Load Modeling) was accomplished through implementing a


hybrid methodology consisting of a combination of two load modeling approaches:

Component Based Approach, which was accomplished by:


o Surveying the supplied load at all main load buses and classifying
them into classes including: Residential, Commercial, Industrial and
District Cooling Load.

120
o Identifying Load Composition for each Load Class including: Resistive
Loads, Small Motor Loads, and Large Motor Loads.
o Selecting a proper aggregate load model representing the identified
load composition. This step is the last step in component-based
approach and the first step in measurement-based approach, in which
the parameters of the aggregated load model shall be estimated using
actual system measurements.
Measurement-Based Approach: which was achieved by:
o Collecting extensive set of monitored natural load versus voltage
variations from Digital Fault Recorders DFRs installed at five different
load buses.
o Estimating aggregate load model parameters from the collected
measurements using parameter estimation techniques.
o Extracting load tripping (self-disconnection) scheme from the collected
measurements.

The updated load model is validated against recorded system disturbances and
found capable, with high accuracy, of replicating the system behavior during the
recorded system disturbances. Several literatures have dealt with aggregate load
modeling and validating the developed load models against small voltage variations
disturbances. However, few studies investigated the validity of these models during
large voltage variation disturbances. In this thesis, the load model parameters were
estimated using actual measurements of small voltage variations disturbances and the
developed load model was validated against recorded large voltage variation
disturbances, however the availability of large voltage vibration disturbances data has
been limited.

The second objective (Voltage Stability Assessment) was accomplished via


performing comprehensive voltage stability studies for Dubai Power Grid, using the
developed load model, under normal operational conditions and various contingency
scenarios. The assessment comprised of two assessment tools:
Static Voltage Stability Analysis using PV and VQ curves to determine the
steady state maximum power transfer limits and reactive power margin for
main load buses.

121
Dynamic Voltage Stability Analysis using Time Domain Simulation to capture
the transient behavior of system voltages during and after disturbances.

The results of voltage stability assessment of Dubai Power Grid shows that the
maximum power transfer margins and reactive power reserve margins for the
substations under study (which represent the worst case scenario) are compliant with
WECC voltage stability criteria. On the other hand, the updated load model revealed
two problems related to voltage stability, the first one, is the transient and/or quasi
steady state voltage rise that was encountered for normal cleared faults. The second
problem is the slow voltage recovery after extended fault durations and/or severe
contingencies, which was also evidenced in the slow voltage recovery during Mushrif
incident in 2009. Both problems could be worsened with higher motor load
proportion, more severe contingencies, and more severe fault conditions. Moreover,
the dynamic simulation proved that steady state voltage stability analysis is not
sufficient to assess the voltage stability of a system with high proportion of motor
loads.

5.2 Recommendations

In view of the escalating stress on most of the power systems around the world
and the associated growing concern in voltage stability problems, it is, therefore,
recommended for all power system utilities, to develop accurate load model based on
real system data and measurements, in order to enhance the accuracy of system
stability studies, especially voltage stability. Power systems that are characterized by
high proportion of air conditioner load are exposed to short-term voltage stability, fast
voltage collapse, and delayed voltage recovery problems. Therefore, the dynamic
characteristic of air conditioning load and their protection must be adequately
represented in voltage stability studies. Furthermore, the new generation of electric
appliances that employ the advent of power electronics (such as inverter-based
appliances) are easily disconnected from the power system following voltage dips.
The self-disconnection of such loads causes a significant rise in post-disturbance
voltage, which is a major concern in voltage stability studies. Therefore, the self-
disconnection behavior of new load types need to be incorporated in power system
load models.

122
The following consideration must be taken while developing load models:

1. The load model will be mainly used for performing stability studies; therefore, it
should be easily integrated into the power system analysis tool.
2. The developed load model should have a generalization capability in order to be
valid under most scenarios. For special system studies, a special load model
should be built according to the study requirement.
3. The developed load model should have clear physical interpretations in order to be
accepted and easily implemented by the system analysts and operators.
4. The developed load model must be verified against actual system disturbances.

Bearing in mind that different load models may give completely opposing conclusions
on system voltage stability, load modeling should be considered as a continuous task.
A developed load model is never considered final and therefore, the developed load
model need to be updated as frequent as possible, every three or five years for
example.

5.3 Future Work

Extensive efforts have been made to extract load model parameters by


analyzing the load-voltage characteristic during small voltage variation disturbances.
However, there is still a lot of w to be done to analyze the load-voltage characteristic
during large voltage variation disturbances and develop a more generalized aggregate
load model. A more comprehensive study would include the analysis of different
sized voltage dips for different load compositions and / or different operating
conditions. This thesis investigated the load-voltage characteristic during short time
scales. Long term time scales were not studied. Also, load frequency sensitivity has
not been analyzed.

It has been found that the estimation of load parameters from sinusoidal signal
is a very challenging task, especially in the presence of harmonic distortion and/or
noise. The high nonlinearity of the load model, the large number of parameters to be
estimated and the huge searching space are the main sources of difficulties faced
during the load model parameter estimation task. Therefore, a great deal of work
needs to be done in this area.

123
Following the recent economic recovery and the resulting real estate business
revival, the demand side at Dubai Power Grid is expected to incur enormous
penetration of more efficient new Air Conditioning technologies based on inverter
driven motors ranging from small home Air Conditioning appliances to large District
Cooling/Chiller Plants. These load appliances are characterized by a higher
sensitivity to sudden voltage variations, and accordingly the amount of self-
disconnected loads due to sudden voltage variations will increase. Indeed, the load-
voltage characteristics of these new appliances need to be represented in the load
model. This would require gathering more information about these appliances from
their manufacturers.

This thesis has dealt with voltage stability assessment, however, voltage
stability enhancement using static and dynamic reactive power compensation such as
Static Var Systems (SVS) were not considered. It is worth to perform a
comprehensive reactive power assessment study for Dubai Power Grid, especially in
view of the anticipated high penetration of new load types including inverter based
appliances.

124
References

[1] T. V. Cutsem, C. Vournas, Voltage Stability of Electric Power Systems, 2nd. ed.,
USA: Kluwer Academic Publishers, 2001.

[2] Force, IEEE Task, "Load Representation for Dynamic Performance Analysis,"
IEEE Transactions on Power Systems, vol. 8, no. 2, pp. 472-482, May 1993.

[3] P. Kundur, J. Paserba, V. Ajjarapu, G. Andersson, A. Bose, C. Canizares, N.


Hatziargyriou, D. Hill, A. Stankovic, C. Taylor, T. V. Cutsem , V. Vittal,
"Definition and Classification of Power System Stability - IEEE/CIGRE Joint
Task Force on Stability Terms and Definitions," IEEE Transactions on Power
Systems, vol. 19, no. 2, pp. 1387-1401, May 2004.

[4] Kundur, P., Power System Stability and Control, USA: Mc Grow Hill Inc., 1994.

[5] Taylor, C. W., Power System Voltage Stability, USA: Mc Grow Hill Inc., 1994.

[6] Hill, D. J., "Non Linear Dynamic Load Models with Recovery for Voltage
Stability Studies," IEEE Transactions on Power Systems, vol. 8, no. 1, pp. 166-
176, February 1993.

[7] G.K. Morison, B. Gao, P. Kundur, "Voltage Stability Analysis Using Static and
Dynamic Approaches," IEEE Transactions on Power Systems, vol. 8, no. 3, pp.
1159-1170, August 1993.

[8] S. A. Al Dessi, A. I. Ibrahim, A. H. Osman, "A Comprehensive Methodology for


Voltage Stability Assessment of Power Systems Using Modern Analytical
Tools," in International Conference on Electric Power and Energy Conversion
Systems (EPECS), Sharjah, 2011.

[9] (RRWG), Reactive Reserve Working Group, "Guide to WECC/NERC Planning


Standards I.D: Voltage Support and Reactive Power," Western Electricity
Coordinating Council, 2006.

[10] Siemens- Power Technologies, Inc., PSS(R)E Version 32.0.5 Program


Application Guide-Volume II, 2010.

[11] Ajjarapu, Venkataramana, Computational Techniques for Voltage Stability


Assessment and Control, Springer, 2006.

[12] Force, IEEE Task, "Standard Load Models for Power Flow and Dynamic
Performance Simulation," IEEE Transactions on Power Systems, vol. 10, no. 3,
pp. 1302-1313, August 1995.

125
[13] Electric Power Research Institute, "Measurement-Based Load Modeling," EPRI,
California, USA, 2006.

[14] Pal, M. K., "Voltage Stability Conditions Considering Load Characteristics,"


IEEE Transactions on Power Systems, vol. 7, no. 1, pp. 243-249, February 1992.

[15] C. J. Lin, Y.T. Chen, C.Y. Chiou, C.H. Huang, H.D. Chiang, J. C. Wang, L. F.
Ahmed, "Dynamic Load Models in Power Systems using the Measurement
Approach," IEEE Transactions on Power Systems, vol. 8, no. 1, pp. 309-315,
February 1993.

[16] P. Ju, E. Handschin, D. Karlsson, "Nonlinear Dynamic Load Modelling: Model


and Parameter Estimation," IEEE Transactions on Power Systems, vol. 11, no. 4,
pp. 1689-1697, November 1996.

[17] W. Xu, Y. Mansour, "Voltage Stability Analysis Using Generic Load Models,"
IEEE Transactions on Power Systems, vol. 9, no. 1, pp. 479-493, 1994.

[18] P. Ju, E. Handschin, D. Karlsson, "Nonlinear Dynamic Load Modelling: Model


and Parameter Estimation," IEEE Transactions on Power Systems, vol. 11, no. 4,
pp. 1689-1697, November 1996.

[19] A. Borghetti, R. Caldon, A. Mari, C. A. Nucci, "On Dynamic Load Models for
Voltage Stability Studies," IEEE Transactions on Power Systems, vol. 12, no. 1,
pp. 293-303, February 1997.

[20] w. Xu, E.Vaahedi, Y. Mansour, J.Tamby, "Voltage Stability Load Parameter


Determination from Field Tests on B.C. Hydro's System," IEEE Transactions on
Power Systems, vol. 12, no. 3, pp. 1290-1297, August 1997.

[21] L. M. Hajagos, B. Danai, "Laboratory Measurements and Models of Modem


Loads and Their Effect on Voltage Stability Studies," IEEE Transactions on
Power Systems, vol. 13, no. 2, pp. 584-592, May 1998.

[22] K.i Tomiyama, J. P. Daniel, S. Ihara, "Modeling Air Conditioner Load for Power
System Studies," IEEE Transactions on Power Systems, vol. 13, no. 2, pp. 414-
421, May 1998.

[23] S.Z. Zhu, J.H. Zheng, S.D. Shen, G.M. Luo, "Effect of Load Modeling on
Voltage Stability," Power Engineering Society Summer Meeting, 2000. IEEE,
vol. 0, no. c, pp. 395-400, 2000.

[24] H. R. J. H. Shi, "Measement-based Load Modeling - Model Structure," in IEEE


Bologna Power Tech Conference, Bologna, Italy, 2003.

126
[25] K. Morison, H. Hamadani, L. Wang, "Practical Issues in Load Modeling for
Voltage Stability Studies," 2003 Power Engineering Society General Meeting,
vol. 3, pp. 1392-1397, 2003.

[26] D. Chen, R. R. Mohler, "Neural-Network-Based Load Modeling and Its Use in


Voltage Stability Analysis," IEEE Transactions on Contorl Systems Technology,
vol. 11, no. 11, pp. 460-469, July 2003.

[27] P. Pourbeik, D. Wang, K. Hoang, "Load Modeling in Voltage Stability Studies,"


in 2005 IEEE Power Engineering Society General Meeting, 2005.

[28] V. Stewart, E. H. Camm, "Modeling of Stalled Motor Loads for Power System
Short-Term Voltage Stability Analysis," in 2005 IEEE Power Engineering
Society General Meeting, 2005.

[29] B.K. Choi, H.D. Chiang, Y. Li, H. Li, "Measurement-Based Dynamic Load
Models: Derivation, Comparison, and Validation," IEEE Transactions on Power
Systems, vol. 21, no. 3, pp. 1276-1283, August 2006.

[30] K. Morison, H. Hamadani, L. Wang, "Load Modeling for Voltage Stability


Studies," in 2006 IEEE Power Systems Conference and Exposition, 2006.

[31] B. WU, Y. ZHANG, M. CHEN, "The Effects of Air Conditioner Load on


Voltage Stability of Urban Power Systems," in 6th WSEAS International
Conference on Power Systems, Lisbon, Portugal, 2006.

[32] L. Y. Taylor, R. A. Jones, S. M. Halpin, "Development of Load Models for Fault


Induced Delayed Voltage Recovery Dynamic Studies," in 2008 IEEE Power and
Energy Society General Meeting - Conversion and Delivery of Electrical Energy
in the 21st Century, 2008.

[33] T. Parveen, G. Ledwich, "Decomposition of Aggregated Load: Finding Induction


Motor Fraction in Real Load," in 2008 Australasian Universities Power
Engineering Conference (AUPEC'08), 2008.

[34] K. Yamashita, M. Asada, K. Yoshimura, "A Development of Dynamic Load


Model Parameter Derivation Method," in Power & Energy Society General
Meeting, Komae, Japan, 2009.

[35] Visconti, I.F., de Souza, L.F.W., Costa, J.M.S.C., Sobrinho, N.R.B.C., "From
Power Quality Monitoring to Transient Stability Analysis: Measurement-based
Load Modeling for Dynamic Simulations," in 14th International Conference on
Harmonics and Quality of Power (ICHQP), 2010.

[36] S. A. Al Dessi, Y. A. Jebril, A. I. Ibrahim, S. A. Shaban, K. Karoui, F.

127
Depierreux, A. Szekut, F. Villella, "Development of a detailed dynamic load
model and implementation staged testing of DEWA network," in GCC-CIGRE
Conference and Exhibition, Doha, Qatar, 2010.

[37] P. Regulski, F. Gonzalez-Longatt, V. Terzija, "Estimation of Load Model


Parameters from Instantaneous Voltage and Current," in 2011 IEEE International
Conference on Fuzzy Systems, Taipei, Taiwan, 2011.

[38] M. Tomoda, J. Matsuk, Y. Hayashi, "Parameter Estimation of Dynamic Load


Model in Power System by using Measured Data," in Journal of International
Council on Electrical Engineering, 2011.

[39] B.H. Kim, H. Kim, B. Lee, "Parameter Estimation for the Composite Load
Model," Journal of International Council on Electrical Engineering, vol. 2, no.
2, pp. 215-218, 2012.

[40] Dubai Electricity and Water Authority, "Dubai Electricity and Water Authority,"
[Online]. Available: http://www.dewa.gov.ae/aboutus/dewahistory.aspx.

[41] Power Transmission Planning Department, "Power Transmission Plan Master


Plan," Dubai Electricity and Water Authority , Dubai, 2012.

[42] Transmission Operation Department-DEWA, "Report on October 20, 2009


Incident," Dubai, 2009.

[43] H. Renmu, M.Jin, D.J. Hill, "Composite Load Modeling via Measurement
Approach," IEEE Transaction on Power Systems, vol. 21, no. 2, pp. 663-672,
2006.

[44] P. Z. V. H.Bai, "A Novel Parameter Identification Approach via Hybrid Learning
for Aggregate Load Modeling," IEEE Transactions on Power Systems, vol. 24,
no. 3, pp. 1145-1154, 2009.

[45] P. C. Krause, O. Wasynczuk, S. D. Sudhoff, Analysis of Electric Machinery and


Drive Systems, IEEE Press, A John Wiley & Sons, Inc, 2002.

[46] I. F. Visconti, L. F. W. de Souza, J. M. S. C. Costa, N. R. B. C. Sobrinho, "From


Power Quality Monitoring to Transient Stability Analysis: Measurement-based
Load Modeling for Dynamic Simulations," in International Conference on
Harmonics and Quality of Power (ICHQP), Rio de Janeiro, 2010.

[47] Reactive Power Reserve Work Group (RRWG), "Voltage Stability Criteria,
Under Voltage Load Shedding Strategy and Reactive Power Reserve Monitoring
Methodolgy," Wester Electricity Coordinating Counci l(WECC), 1998.

128
129
Appendix A

Figure A.1: Simulink Model for Load Model Parameter Estimation


130
Appendix B

Figure B.1: Dubai Power Grid Planned Network Topology - Year 2014 [41]
Jebel Ali Power Complex

MBCH

NHDA

MUSH
CARX

131

Figure B.2: Dubai Power Grid Geographical Map [41]
Table B.1: Comprehensive Contingency List for NHDA 400/132 kV Substation

Contingency Type Lost Elements Contingency Identifier Contingency Description


Loss of One 400/132 kV Transformer CONTINGENCY-1 Loss of One 400/132 kV Transformer
CONTINGENCY-2 Loss of NHDA - MBCH 400 kV Circuit
N-1 CONTINGENCY-3 Loss of NHDA - HSTA 400 kV Circuit
Loss of One 400 kV Circuit
CONTINGENCY-4 Loss of NHDA - TCGY 400 kV Circuit
CONTINGENCY-5 Loss of NHDA - WRSN 400 kV Circuit
Loss of NHDA - MBCH 400 kV Circuit
CONTINGENCY-6
Loss of NHDA - HSTA 400 kV Circuit
Loss of NHDA - MBCH 400 kV Circuit
CONTINGENCY-7
Loss of NHDA - TCGY 400 kV Circuit
Loss of NHDA - MBCH 400 kV Circuit
CONTINGENCY-8
Loss of NHDA - WRSN 400 kV Circuit
Loss of Two 400 kV Circuit
Loss of NHDA - WRSN 400 kV Circuit
CONTINGENCY-9

132
Loss of NHDA - HSTA 400 kV Circuit
Loss of NHDA - WRSN 400 kV Circuit
CONTINGENCY-10
Loss of NHDA - TCGY 400 kV Circuit
N-2 Loss of NHDA - HSTA 400 kV Circuit
CONTINGENCY-11
Loss of NHDA - TCGY 400 kV Circuit
Loss of One 400/132 kV Transformer
CONTINGENCY-12
Loss of NHDA - MBCH 400 kV Circuit
Loss of One 400/132 kV Transformer
CONTINGENCY-13
Loss of One 400 kV Circuit Loss of NHDA - HSTA 400 kV Circuit
+ Loss of One 400/132 kV Transformer
One 400/132 kV Transformer CONTINGENCY-14
Loss of NHDA - TCGY 400 kV Circuit
Loss of One 400/132 kV Transformer
CONTINGENCY-15
Loss of NHDA - WRSN 400 kV Circuit
Contingency Type Lost Elements Contingency Identifier Contingency Description

Loss of One 400/132 kV Transformer


CONTINGENCY-16 Loss of NHDA - MBCH 400 kV Circuit
Loss of NHDA - HSTA 400 kV Circuit

Loss of One 400/132 kV Transformer


CONTINGENCY-17 Loss of NHDA - MBCH 400 kV Circuit
Loss of NHDA - TCGY 400 kV Circuit

Loss of One 400/132 kV Transformer


CONTINGENCY-18 Loss of NHDA - MBCH 400 kV Circuit
Loss of Two 400 kV Circuit Loss of NHDA - WRSN 400 kV Circuit
N-3 +
One 400/132 kV Transformer Loss of One 400/132 kV Transformer

133
CONTINGENCY-19 Loss of NHDA - WRSN 400 kV Circuit
Loss of NHDA - HSTA 400 kV Circuit

Loss of One 400/132 kV Transformer


CONTINGENCY-20 Loss of NHDA - WRSN 400 kV Circuit
Loss of NHDA - TCGY 400 kV Circuit

Loss of One 400/132 kV Transformer


CONTINGENCY-21 Loss of NHDA - HSTA 400 kV Circuit
Loss of NHDA - TCGY 400 kV Circuit
Table B.2: Comprehensive Contingency List for MUSH 400/132 kV Substation

Contingency Type Lost Elements Contingency Identifier Contingency Description


Loss of One 400/132 kV Transformer CONTINGENCY-1 Loss of One 400/132 kV Transformer
CONTINGENCY-2 Loss of MUSH - CARX 400 kV Circuit
N-1
Loss of One 400 kV Circuit CONTINGENCY-3 Loss of MUSH - WRSN 400 kV Circuit
CONTINGENCY-4 Loss of MUSH - MBCH 400 kV Circuit
Loss of MUSH - CARX 400 kV Circuit-ckt 1
CONTINGENCY-5
Loss of MUSH - CARX 400 kV Circuit-ckt 2
Loss of MUSH - CARX 400 kV Circuit
CONTINGENCY-6
Loss of MUSH - WRSN 400 kV Circuit
Loss of Two 400 kV Circuit
Loss of MUSH - CARX 400 kV Circuit
CONTINGENCY-7
Loss of MUSH - MBCH 400 kV Circuit
Loss of MUSH - WRSN 400 kV Circuit
N-2 CONTINGENCY-8
Loss of MUSH - MBCH 400 kV Circuit
Loss of One 400/132 kV Transformer
CONTINGENCY-9

134
Loss of MUSH - CARX 400 kV Circuit
Loss of One 400 kV Circuit
Loss of One 400/132 kV Transformer
+ CONTINGENCY-10
Loss of MUSH - WRSN 400 kV Circuit
One 400/132 kV Transformer
Loss of One 400/132 kV Transformer
CONTINGENCY-11
Loss of MUSH - MBCH 400 kV Circuit
Loss of One 400/132 kV Transformer
CONTINGENCY-12 Loss of MUSH - CARX 400 kV Circuit-ckt 1
Loss of MUSH - CARX 400 kV Circuit-ckt 2
Loss of One 400/132 kV Transformer
CONTINGENCY-13 Loss of MUSH - CARX 400 kV Circuit
Loss of Two 400 kV Circuit Loss of MUSH - WRSN 400 kV Circuit
N-3 +
Loss of One 400/132 kV Transformer
One 400/132 kV Transformer
CONTINGENCY-14 Loss of MUSH - CARX 400 kV Circuit
Loss of MUSH - MBCH 400 kV Circuit
Loss of One 400/132 kV Transformer
CONTINGENCY-15 Loss of MUSH - WRSN 400 kV Circuit
Loss of MUSH - MBCH 400 kV Circuit
Table B.3: Comprehensive Contingency List for CARX 400/132 kV Substation

Contingency Type Lost Elements Contingency Identifier Contingency Description


Loss of One 400/132 kV Transformer CONTINGENCY-1 Loss of One 400/132 kV Transformer
CONTINGENCY-2 Loss of CARX - MUSH 400 kV Circuit
N-1
Loss of One 400 kV Circuit CONTINGENCY-3 Loss of CARX - TCGY 400 kV Circuit
CONTINGENCY-4 Loss of CARX - HRSR 400 kV Circuit
Loss of CARX - MUSH 400 kV Circuit-ckt 1
CONTINGENCY-5
Loss of CARX - MUSH 400 kV Circuit-ckt 2
Loss of CARX - MUSH 400 kV Circuit
CONTINGENCY-6
Loss of CARX - TCGY 400 kV Circuit
Loss of Two 400 kV Circuit
Loss of CARX - MUSH 400 kV Circuit
CONTINGENCY-7
Loss of CARX - HRSR 400 kV Circuit
Loss of CARX - TCGY 400 kV Circuit
N-2 CONTINGENCY-8
Loss of CARX - HRSR 400 kV Circuit
Loss of One 400/132 kV Transformer
CONTINGENCY-9
Loss of CARX - MUSH 400 kV Circuit

135
Loss of One 400 kV Circuit
Loss of One 400/132 kV Transformer
+ CONTINGENCY-10
Loss of CARX - TCGY 400 kV Circuit
One 400/132 kV Transformer
Loss of One 400/132 kV Transformer
CONTINGENCY-11
Loss of CARX - HRSR 400 kV Circuit
Loss of One 400/132 kV Transformer
CONTINGENCY-12 Loss of CARX - MUSH 400 kV Circuit-ckt 1
Loss of CARX - MUSH 400 kV Circuit-ckt 2
Loss of One 400/132 kV Transformer
CONTINGENCY-13 Loss of CARX - MUSH 400 kV Circuit
Loss of Two 400 kV Circuit
Loss of CARX - TCGY 400 kV Circuit
N-3 +
One 400/132 kV Transformer Loss of One 400/132 kV Transformer
CONTINGENCY-14 Loss of CARX - MUSH 400 kV Circuit
Loss of CARX - HRSR 400 kV Circuit
Loss of One 400/132 kV Transformer
CONTINGENCY-15 Loss of CARX - TCGY 400 kV Circuit
Loss of CARX - HRSR 400 kV Circuit
Table B.4: Comprehensive Contingency List for CARX 400/132 kV Substation

Contingency Type Lost Elements Contingency Identifier Contingency Description


Loss of One 400/132 kV Transformer CONTINGENCY-1 Loss of One 400/132 kV Transformer
CONTINGENCY-2 Loss of MBCH - MUSH 400 kV Circuit
N-1
Loss of One 400 kV Circuit CONTINGENCY-3 Loss of MBCH - NHDA 400 kV Circuit
CONTINGENCY-4 Loss of MBCH - WRSN 400 kV Circuit
Loss of MBCH - MUSH 400 kV Circuit
CONTINGENCY-5
Loss of MBCH - NHDA 400 kV Circuit
Loss of MBCH - MUSH 400 kV Circuit
Loss of Two 400 kV Circuit CONTINGENCY-6
Loss of MBCH - WRSN 400 kV Circuit
Loss of MBCH - NHDA 400 kV Circuit
CONTINGENCY-7
Loss of MBCH - WRSN 400 kV Circuit
N-2
Loss of One 400/132 kV Transformer
CONTINGENCY-8
Loss of MBCH - MUSH 400 kV Circuit

136
Loss of One 400 kV Circuit
Loss of One 400/132 kV Transformer
+ CONTINGENCY-9
Loss of MBCH - NHDA 400 kV Circuit
One 400/132 kV Transformer
Loss of One 400/132 kV Transformer
CONTINGENCY-10
Loss of MBCH - WRSN 400 kV Circuit
Loss of One 400/132 kV Transformer
CONTINGENCY-11 Loss of MBCH - MUSH 400 kV Circuit
Loss of MBCH - NHDA 400 kV Circuit
Loss of Two 400 kV Circuit Loss of One 400/132 kV Transformer
N-3 + CONTINGENCY-12 Loss of MBCH - MUSH 400 kV Circuit
One 400/132 kV Transformer Loss of MBCH - WRSN 400 kV Circuit
Loss of One 400/132 kV Transformer
CONTINGENCY-13 Loss of MBCH - NHDA 400 kV Circuit
Loss of MBCH - WRSN 400 kV Circuit
7DEOH%WECC9oltagH6WDELOLW\&ULWHULD>@

137
Appendix C

V (p.u.) BASE CASE


1.10 CONTINGENCY-1
CONTINGENCY-2
CONTINGENCY-3
1.05 CONTINGENCY-4
CONTINGENCY-5
CONTINGENCY-6
1.00 CONTINGENCY-7
CONTINGENCY-8
CONTINGENCY-9
CONTINGENCY-10
0.95
CONTINGENCY-11
CONTINGENCY-12
CONTINGENCY-13
0.90
CONTINGENCY-14
CONTINGENCY-15

0.85
500 700 900 1100 1300 1500 1700 Power Transfer (MW)

Figure C.1: PV Curves for MUSH 400/132 kV Substation Base Case and All Contingencies
V (p.u.)
1.10

1.05

CONTINGENCY-1
1.00
CONTINGENCY-2
CONTINGENCY-3
0.95 CONTINGENCY-4

0.90

0.85
500 700 900 1100 1300 1500 1700 Power Transfer (MW)

Figure C.2: PV Curves for MUSH 400/132 kV Substation All N-1 Contingencies
V (p.u.)
1.10

1.05

CONTINGENCY-5
1.00 CONTINGENCY-6
CONTINGENCY-7
CONTINGENCY-8
0.95 CONTINGENCY-9
CONTINGENCY-10
CONTINGENCY-11
0.90

0.85
500 700 900 1100 1300 1500 1700 Power Transfer (MW)
Figure C.3: PV Curves for MUSH 400/132 kV Substation All N-2 Contingencies

138
V (p.u.)
1.10

1.05

1.00 CONTINGENCY-12
CONTINGENCY-13
CONTINGENCY-14
0.95 CONTINGENCY-15

0.90

0.85 Power Transfer (MW)


500 700 900 1100 1300 1500 1700
Figure C.4: PV Curves for MUSH 400/132 kV Substation All N-3 Contingencies

Table C.1: Maximum Power Transfer Margin for MUSH 400/132 kV Substation Base Case and
All Contingencies
Active Power Maximum Active
Contingency Identifier Category Transfer Limit Power Transfer
(MW) Margin (MW)
BASE CASE N-0 1429 1358

CONTINGENCY-1 1248

CONTINGENCY-2 1429
N-1 1186
CONTINGENCY-3 1424

CONTINGENCY-4 1378

CONTINGENCY-5 1429

CONTINGENCY-6 1419

CONTINGENCY-7 1429

CONTINGENCY-8 N-2 1368 1217

CONTINGENCY-9 1248

CONTINGENCY-10 1242

CONTINGENCY-11 1229

CONTINGENCY-12 1229

CONTINGENCY-13 1236
N-3 1168
CONTINGENCY-14 1229

CONTINGENCY-15 1198

139
Q (MVAr)
2000
BASE CASE
CONTINGENCY-1
1500 CONTINGENCY-2
CONTINGENCY-3
1000 CONTINGENCY-4
CONTINGENCY-5
CONTINGENCY-6
500 CONTINGENCY-7
CONTINGENCY-8
0 CONTINGENCY-9
CONTINGENCY-10
CONTINGENCY-11
-500
CONTINGENCY-12
CONTINGENCY-13
-1000 CONTINGENCY-14
CONTINGENCY-15
-1500
0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 V(p.u.)

Figure C.5: QV Curves for MUSH 400/132 kV Substation Base Case and All Contingencies

Q (MVAr)
2000

1500

1000
CONTINGENCY-1
500 CONTINGENCY-2
CONTINGENCY-3
0 CONTINGENCY-4

-500

-1000

-1500
0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 V(p.u.)

Figure C.6: QV Curves for MUSH 400/132 kV Substation All N-1 Contingencies

Q (MVAr)
2000

1500

1000
CONTINGENCY-5
500 CONTINGENCY-6
CONTINGENCY-7
CONTINGENCY-8
0
CONTINGENCY-9
CONTINGENCY-10
-500
CONTINGENCY-11

-1000

-1500
0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 V(p.u.)

Figure C.7: QV Curves for MUSH 400/132 kV Substation All N-2 Contingencies

140
Q (MVAr)
2000

1500

1000
CONTINGENCY-12
CONTINGENCY-13
500 CONTINGENCY-14
CONTINGENCY-15
0

-500

-1000
0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 V(p.u.)

Figure C.8: QV Curves for MUSH 400/132 kV Substation All N-3 Contingencies

Table C.2: Available Reactive Power Margin for MUSH 400/132 kV Substation Base Case and
All Contingencies
Reactive Power
Reserve Margin
Change Available
(RRM) (MVAr)
Contingency Identifier Category in RRM RRM
Base
Base (MVAr) (MVAr)
Load
Load
+5%
BASE CASE N-0 1014 961 53 1014

CONTINGENCY-1 889 796 93

CONTINGENCY-2 994 941 53


N-1 796
CONTINGENCY-3 985 932 53

CONTINGENCY-4 872 819 52

CONTINGENCY-5 776 710 67

CONTINGENCY-6 956 904 53

CONTINGENCY-7 852 799 53

CONTINGENCY-8 N-2 820 755 65 637

CONTINGENCY-9 870 760 110

CONTINGENCY-10 860 750 110

CONTINGENCY-11 748 637 111

CONTINGENCY-12 625 564 61

CONTINGENCY-13 832 723 109


N-3 517
CONTINGENCY-14 727 617 111

CONTINGENCY-15 669 517 152

141
V (p.u.)
1.10 BASE CASE
CONTINGENCY-1
CONTINGENCY-2
1.05 CONTINGENCY-3
CONTINGENCY-4
CONTINGENCY-5
1.00 CONTINGENCY-6
CONTINGENCY-7
CONTINGENCY-8
CONTINGENCY-9
0.95
CONTINGENCY-10
CONTINGENCY-11
CONTINGENCY-12
0.90 CONTINGENCY-13
CONTINGENCY-14
CONTINGENCY-15
0.85
200 400 600 800 1000 1200 1400 Power Transfer (MW)
Figure C.9: PV Curves for CARX 400/132 kV Substation Base Case and All Contingencies

V (p.u.)
1.10

1.05

1.00 CONTINGENCY-1
CONTINGENCY-2
CONTINGENCY-3
0.95 CONTINGENCY-4

0.90

0.85
200 400 600 800 1000 1200 1400 Power Transfer (MW)
Figure C.10: PV Curves for CARX 400/132 kV Substation All N-1 Contingencies

V (p.u.)
1.10

1.05

1.00 CONTINGENCY-5
CONTINGENCY-6
CONTINGENCY-7
0.95 CONTINGENCY-8
CONTINGENCY-9
CONTINGENCY-10
0.90 CONTINGENCY-11

0.85
200 400 600 800 1000 1200 1400 Power Transfer (MW)
Figure C.11: PV Curves for CARX 400/132 kV Substation All N-2 Contingencies

142
V (p.u.)
1.10

1.05

1.00 CONTINGENCY-12
CONTINGENCY-13
CONTINGENCY-14
0.95 CONTINGENCY-15

0.90

0.85
200 400 600 800 1000 1200 1400 Power Transfer (MW)
Figure C.12: PV Curves for CARX 400/132 kV Substation All N-3 Contingencies

Table C.3: Maximum Power Transfer Margin for CARX 400/132 kV Substation Base Case and
All Contingencies
Active Power Maximum Active
Contingency Identifier Category Transfer Limit Power Transfer
(MW) Margin (MW)
BASE CASE N-0 894 849

CONTINGENCY-1 744

CONTINGENCY-2 888
N-1 707
CONTINGENCY-3 881

CONTINGENCY-4 856

CONTINGENCY-5 888

CONTINGENCY-6 894

CONTINGENCY-7 850

CONTINGENCY-8 N-2 894 713

CONTINGENCY-9 744

CONTINGENCY-10 738

CONTINGENCY-11 731

CONTINGENCY-12 738

CONTINGENCY-13 738
N-3 695
CONTINGENCY-14 731

CONTINGENCY-15 713

143
Q (MVAr)
1500
BASE CASE
CONTINGENCY-1
1000 CONTINGENCY-2
CONTINGENCY-3
CONTINGENCY-4
500 CONTINGENCY-5
CONTINGENCY-6
CONTINGENCY-7
0 CONTINGENCY-8
CONTINGENCY-9
CONTINGENCY-10
-500 CONTINGENCY-11
CONTINGENCY-12
CONTINGENCY-13
-1000
CONTINGENCY-14
CONTINGENCY-15
-1500
0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 V(p.u.)

Figure C.13: QV Curves for CARX 400/132 kV Substation Base Case and All Contingencies

Q (MVAr)
1500

1000

500
CONTINGENCY-1
CONTINGENCY-2
0 CONTINGENCY-3
CONTINGENCY-4

-500

-1000

-1500
0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 V(p.u.)

Figure C.14: QV Curves for CARX 400/132 kV Substation All N-1 Contingencies

Q (MVAr)
1500

1000

500 CONTINGENCY-5
CONTINGENCY-6
CONTINGENCY-7
0
CONTINGENCY-8
CONTINGENCY-9
-500 CONTINGENCY-10
CONTINGENCY-11
-1000

-1500
0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 V(p.u.)

Figure C.15: QV Curves for CARX 400/132 kV Substation All N-2 Contingencies

144
Q (MVAr)
1500

1000

500
CONTINGENCY-12
CONTINGENCY-13
0 CONTINGENCY-14
CONTINGENCY-15

-500

-1000

-1500
0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 V(p.u.)

Figure C.16: QV Curves for CARX 400/132 kV Substation All N-3 Contingencies

Table C.4: Available Reactive Power Margin for CARX 400/132 kV Substation Base Case and
All Contingencies
Reactive Power
Reserve Margin
Change in Available
(RRM) (MVAr)
Contingency Identifier Category RRM RRM
Base
Base (MVAr) (MVAr)
Load
Load
+5%
BASE CASE N-0 1029 1002 28 1002
CONTINGENCY-1 1011 978 32

CONTINGENCY-2 1026 998 28


N-1 710
CONTINGENCY-3 958 930 28

CONTINGENCY-4 744 710 34

CONTINGENCY-5 1029 1002 27

CONTINGENCY-6 954 926 28

CONTINGENCY-7 737 705 31

CONTINGENCY-8 N-2 613 581 31 683


CONTINGENCY-9 1007 974 32

CONTINGENCY-10 939 907 33

CONTINGENCY-11 720 683 37

CONTINGENCY-12 1010 978 32

CONTINGENCY-13 935 903 32


N-3 553
CONTINGENCY-14 715 678 37

CONTINGENCY-15 591 553 37

145
V (p.u.)
1.10
BASE CASE
CONTINGENCY-1
1.05 CONTINGENCY-2
CONTINGENCY-3
CONTINGENCY-4
1.00 CONTINGENCY-5
CONTINGENCY-6
CONTINGENCY-7
0.95 CONTINGENCY-8
CONTINGENCY-9
CONTINGENCY-10
0.90 CONTINGENCY-11
CONTINGENCY-12
CONTINGENCY-13
0.85
200 400 600 800 1000 1200 1400 Power Transfer (MW)

Figure C.17: PV Curves for MBCH 400/132 kV Substation Base Case and All Contingencies

V (p.u.)
1.10

1.05

1.00 CONTINGENCY-1
CONTINGENCY-2
CONTINGENCY-3
CONTINGENCY-4
0.95

0.90

0.85 Power Transfer (MW)


200 400 600 800 1000 1200 1400
Figure C.18: PV Curves for MBCH 400/132 kV Substation All N-1 Contingencies

V (p.u.)
1.10

1.05

1.00 CONTINGENCY-5
CONTINGENCY-6
CONTINGENCY-7
0.95 CONTINGENCY-8
CONTINGENCY-9
CONTINGENCY-10
0.90

0.85
200 400 600 800 1000 1200 1400 Power Transfer (MW)

Figure C.19: PV Curves for MBCH 400/132 kV Substation All N-2 Contingencies

146
V (p.u.)
1.10

1.05

CONTINGENCY-11
1.00
CONTINGENCY-12
CONTINGENCY-13
0.95

0.90

0.85
200 400 600 800 1000 1200 1400 Power Transfer (MW)

Figure C.20: PV Curves for MBCH 400/132 kV Substation All N-3 Contingencies

Table C.5: Maximum Power Transfer Margin for MBCH 400/132 kV Substation Base Case
and All Contingencies
Maximum Active Maximum Active
Contingency Identifier Category Power Transfer Power Transfer
Limit (MW) Margin (MW)
BASE CASE N-0 826 785

CONTINGENCY-1 676

CONTINGENCY-2 807
N-1 642
CONTINGENCY-3 807

CONTINGENCY-4 807

CONTINGENCY-5 626

CONTINGENCY-6 701

CONTINGENCY-7 807
N-2 610
CONTINGENCY-8 688

CONTINGENCY-9 645

CONTINGENCY-10 645

CONTINGENCY-11 563

CONTINGENCY-12 N-3 626 549


CONTINGENCY-13 638

147
Q (MVAr)
2000
BASE CASE
CONTINGENCY-1
1500
CONTINGENCY-2
CONTINGENCY-3
1000
CONTINGENCY-4
CONTINGENCY-5
500 CONTINGENCY-6
CONTINGENCY-7
0 CONTINGENCY-8
CONTINGENCY-9
-500 CONTINGENCY-10
CONTINGENCY-11
-1000 CONTINGENCY-12
CONTINGENCY-13
-1500
0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 V(p.u.)

Figure C.21: QV Curves for MBCH 400/132 kV Substation Base Case and All Contingencies

Q (MVAr)
1500

1000

500
CONTINGENCY-1
CONTINGENCY-2
0 CONTINGENCY-3
CONTINGENCY-4

-500

-1000

-1500
0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 V(p.u.)

Figure C.22: QV Curves for MBCH 400/132 kV Substation All N-1 Contingencies

Q (MVAr)
1500

1000

500 CONTINGENCY-5
CONTINGENCY-6
0
CONTINGENCY-7
CONTINGENCY-8
-500
CONTINGENCY-9

-1000

-1500
0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 V(p.u.)

Figure C.23: QV Curves for MBCH 400/132 kV Substation All N-2 Contingencies

148
Q (MVAr)
1500

1000

CONTINGENCY-11
500
CONTINGENCY-12
CONTINGENCY-13

-500

-1000
0.70 0.75 0.80 0.85 0.90 0.95 1.00 1.05 1.10 1.15 1.20 V(p.u.)

Figure C.24: QV Curves for MBCH 400/132 kV Substation All N-3 Contingencies

Table C.6: Available Reactive Power Margin for MBCH 400/132 kV Substation Base Case and
All Contingencies

Reactive Power
Reserve Margin
(RRM) (MVAr) Change Available
Contingency Identifier Category in RRM RRM
(MVAr) (MVAr)
Base Base Load
Load +5%

BASE CASE N-0 1006 977 29 977


CONTINGENCY-1 978 941 36
CONTINGENCY-2 802 769 33
N-1 769
CONTINGENCY-3 949 920 29
CONTINGENCY-4 923 896 28
CONTINGENCY-5 641 596 45
CONTINGENCY-6 685 652 33
CONTINGENCY-7 857 828 29
N-2 596
CONTINGENCY-8 767 723 45
CONTINGENCY-9 920 884 36
CONTINGENCY-10 895 860 35
CONTINGENCY-11 579 512 66
CONTINGENCY-12 N-3 650 606 45 512
CONTINGENCY-13 829 792 36

149
Vita

Salha Ali Al Disi was born on September 14, 1974, in Al Ain, UAE. She
started her education in local public schools, and then graduated from Al Ain
Secondary School in 1992. She was ranked 9th among the 1991/92 UAE Secondary
Students Batch. She then enrolled in UAE University in Al Ain, from which she
received her Bachelor Degree, with Second Class Honors, in Electrical Engineering in
1998. She had secured a GPA of 3.95/4 and was ranked the 1st among the 1997/98
UAE University Students Batch.

Ms. Al Disi has been working at Dubai Electricity and Water Authority since
2005. She is presently holding the position of Manager System Analysis in Power
Transmission Planning Department. In 2010, Ms. Al Disi was awarded a scholarship
from Dubai Electricity and Water Authority to attain her Masters Degree in Electrical
Engineering at the American University of Sharjah. She was awarded the Master of
Science Degree in Electrical Engineering in Spring 2013.

150

You might also like