AC LINK LOAD FLOW CONTROL IN ELECTRIC POWER SYSTEMS

AC LINK LOAD FLOW CONTROL IN
ELECTRIC POWER SYSTEMS
by
Fernando Mancilla−David
A Dissertation
submitted in partial fulfillment of the
requirements for the degree of
Doctor of Philosophy
(Electrical Engineering)
at the
UNIVERSITY OF WISCONSIN−MADISON
2007
© Copyright by Fernando A. Mancilla−David 2007
All Rights Reserved
i
Abstract
As congestion in
AC
power transmission systems increases due to technical
limitations, environmental pressures, economic necessities and developmental demands,
devices with the ability to control power flow with much greater ease than presently
possible would be highly desirable. Furthermore, devices that channel power flow in the
network to fulfill contractual obligations for purchase agreements between power
producers and consumers would facilitate improved operation of the power system in the
emerging deregulated scenario. Following this trend, several power electronic-based
devices capable of accomplishing these objectives have been developed, unified under
the rubric of Flexible AC Transmission Systems, FACTS.
This thesis proposes a new family of
FACTS
devices: the Π−controller, the
Ξ−controller and the Γ−controller. Based on the vector switching converter (VeSC), these
devices feature similar functional characteristics to the ones encountered in the voltage
source HVDC interties, the Static Synchronous Series Compensator (SSSC) and the Unified
Power Flow Controller (UPFC), respectively. The control function, however, is realized
by direct
AC↔AC
conversion without frequency change, adopting the strategy of pulse
width modulation (PWM).
A comprehensive analysis is performed, including operating principles,
equivalent circuits, design consideration and dynamic modeling and control. Computer
simulations of illustrative examples are executed to validate the approach.
For the series compensation case, this new technology is contrasted against the
state−of−the−art. A detailed comparative study between the Ξ−controller and the
SSSC
shows that the use of VeSC−based power flow controllers could become an alternative for
the next generation of FACTS devices.
In order to fully prove the concept of
VeSC−based
power flow controllers, a
ii
laboratory prototype was built and tested. The Γ−controller was chosen for experimental
verification, as its topology is the most general of all three controllers.
Finally, a system point of view is studied, incorporating the Ξ−controller into an
Optimal Power Flow (OPF) program.
iii
Acknowledgments
I would like to express my sincere gratitude and thanks to my advisor Professor
Giri Venkataramanan for his guidance, counsel and encouragement at every stage of this
research. I also thank my former advisor Professor Fernando Alvarado who gave me the
opportunity to become a graduate student at the University of Wisconsin and guided my
studies towards my M.S. degree.
I acknowledge the Wisconsin Electric Machines and Power Electronics
Consortium (WEMPEC), the Electric Power Research Institute (EPRI), and the Electrical
and Computer Engineering Department at the University of Wisconsin for funding this
research.
I have been fortunate to work with the
WEMPEC
and power area professors,
students and support staff who provided an excellent working environment with all the
resources that made this research possible.
I also thank Professor Subhashish Bhattacharya from North Carolina State
University for his contributions to Chapter 6 of this thesis.
I acknowledge the cooperation and support of Patrick Flannery who helped me
implement my experiments at the WEMPEC laboratory.
Last but not least, I express thanks to my dearest father for his effort,
encouragement and inspiration. ¡Te quiero mucho papá!
iv
Table of Contents
Abstract............................................................................................. i
Acknowledgments ........................................................................... iii
Table of Contents ............................................................................iv
List of Figures .............................................................................. xiii
List of Tables .................................................................................. xx
Chapter 1 Introduction ................................................................22
1.1...... General ...................................................................................................... 22
1.1.1
Congestion in Transmission Systems............................................ 24
1.1.2
Loop Flows ................................................................................... 24
1.2...... Power Flow Control through FACTS Technology...................................... 26
1.3...... Dynamic Compensation of Power System with FACTS Devices............... 28
1.4...... Literature Survey....................................................................................... 28
1.5...... State−of−the−Art....................................................................................... 33
1.5.1
The Convertible Static Compensator ............................................ 33
1.5.2
Distributed FACTS .......................................................................... 34
1.6...... Outline of the Research............................................................................. 36
v
Chapter 2 Vector Switching Converters ..................................... 39
2.1...... Introduction ............................................................................................... 39
2.2...... Three−Phase Vector Switching Converters .............................................. 39
2.2.1
Converter Model ........................................................................... 39
2.2.2
Equivalent Circuit ......................................................................... 43
2.2.3
Switch Realization ........................................................................ 46
2.2.4
VeSC Pulse Width Modulation...................................................... 47
2.3...... Multiple−Secondary Phase−Shifter Transformer ..................................... 50
Chapter 3 AC Link Power Flow Controllers............................... 54
3.1...... Introduction ............................................................................................... 54
3.2...... Π−Controller: An AC Link Back−to−Back Intertie................................... 55
3.2.1
Topology ....................................................................................... 55
3.2.2
Equivalent Circuit ......................................................................... 57
3.2.3
Principle of Operation ................................................................... 58
3.2.4
Control Capabilities ...................................................................... 61
3.2.4.1
Active Power Control............................................................. 62
3.2.4.2
Reactive Power Control ......................................................... 63
3.2.4.3
Angle Control......................................................................... 64
3.2.5
Design Considerations .................................................................. 65
3.2.5.1
Parameter Calculations........................................................... 65
vi
3.2.5.2
Feasible operating range ........................................................ 66
3.2.6
Simulation Results ........................................................................ 67
3.2.7
Summary ....................................................................................... 69
3.3...... Ξ−Controller: An AC Link Series Compensator ....................................... 70
3.3.1
Topology ....................................................................................... 71
3.3.2
Equivalent Circuit ......................................................................... 72
3.3.3
Principle of Operation ................................................................... 74
3.3.4
Control Capabilities ...................................................................... 75
3.3.5
Design Considerations .................................................................. 77
3.3.6
Simulation Results ........................................................................ 78
3.3.7
Summary ....................................................................................... 81
3.4...... Γ−Controller: An AC Link Unified Power Flow Controller...................... 81
3.4.1
Topology ....................................................................................... 82
3.4.2
Equivalent Circuit ......................................................................... 85
3.4.3
Principle of Operation ................................................................... 86
3.4.4
Control Capabilities ...................................................................... 89
3.4.4.1
Transmitted Power Control .................................................... 90
3.4.4.2
Tap−changing/phase−shifting control.................................... 91
3.4.5
Design Considerations .................................................................. 93
3.4.5.1
Transformers turn ratio .......................................................... 93
3.4.5.2
Filter Capacitors / Switching Frequency................................ 93
3.4.6
Simulation Results ........................................................................ 95
vii
3.4.7
Summary ..................................................................................... 101
Chapter 4 Dynamic Modeling of AC Link Power Flow
Controllers 103
4.1...... Introduction ............................................................................................. 103
4.2...... Generalized Averaging Theory ............................................................... 104
4.2.1
Index symmetry........................................................................... 105
4.2.2
Reconstruction............................................................................. 105
4.2.3
Derivative with respect to time ................................................... 106
4.2.4
Average of a product................................................................... 106
4.3...... Dynamics for electric elements............................................................... 106
4.3.1
Resistor........................................................................................ 107
4.3.2
Inductor ....................................................................................... 107
4.3.3
Capacitor ..................................................................................... 108
4.4...... Application to State Space Models for Vector Switching Converters.... 108
4.4.1
Real Excitation ............................................................................ 114
4.4.2
Complex Excitation..................................................................... 117
4.5...... Case Studies ............................................................................................ 121
4.5.1
Generalized Averaging Applied to the Ξ−controller .................. 121
4.5.1.1
The Model ............................................................................ 121
4.5.1.2
Model Accuracy ................................................................... 124
viii
4.5.1.3
4.5.2
Eigenvalues Pattern.............................................................. 126
Generalized Averaging Applied to the Γ−controller .................. 127
4.5.2.1
The Model ............................................................................ 127
4.5.2.2
Model Accuracy ................................................................... 130
4.6...... Summary ................................................................................................. 133
Chapter 5 Feedback Control for AC Link Power Flow
Controllers 134
5.1...... Introduction ............................................................................................. 134
5.2...... Feedback Control for the Ξ−controller ................................................... 135
5.2.1
Proposed Controller .................................................................... 136
5.2.1.1
Small Signal Model.............................................................. 138
5.2.1.2
Controller Design via Bode Plots......................................... 141
5.2.2
Simulation Results ...................................................................... 145
5.2.2.1
Response to a step change in the reference.......................... 145
5.2.2.2
Response to disturbances ..................................................... 146
5.3...... Feedback Control for the Γ−controller.................................................... 148
5.4...... Summary ................................................................................................. 149
ix
Chapter 6 A comparative evaluation of Power Flow Controllers
using AC Link and DC Link Converters........................................ 150
6.1...... Introduction ............................................................................................. 150
6.2...... Candidate Compensators......................................................................... 152
6.2.1
DC
Link Series Compensator....................................................... 152
6.2.2
AC
Link Series Compensator....................................................... 153
6.3...... Test system description ........................................................................... 155
6.4...... Steady state performance evaluation....................................................... 156
6.4.1
Operating Modes......................................................................... 156
6.4.2
Design Considerations ................................................................ 158
6.4.2.1
Choice of power semiconductors ......................................... 159
6.4.2.2
Injection Transformer........................................................... 161
6.4.2.3
Capacitor .............................................................................. 164
Capacitor Selection ............................................... 164
6.4.2.3.1
SSSC
6.4.2.3.2
Ξ−Controller Capacitor Selection ................................. 166
6.4.2.4
Converter.............................................................................. 168
6.4.2.4.1
Semiconductors Ratings................................................ 168
6.4.2.4.2
Device Losses................................................................ 170
6.4.2.4.3
Waveform Quality......................................................... 174
6.5...... Dynamic performance evaluation ........................................................... 175
6.5.1
Dynamic Modeling...................................................................... 175
x
6.5.2
Control Structure......................................................................... 176
6.5.3
Regulator Design......................................................................... 178
6.5.4
Command Response.................................................................... 179
6.5.5
Disturbance Response ................................................................. 180
6.6...... Summary ................................................................................................. 180
Chapter 7 Experimental Verification of the Γ−Controller .......182
7.1...... Introduction ............................................................................................. 182
7.2...... Experiment Layout.................................................................................. 183
7.2.1
Laboratory Prototype of a Transmission Line ............................ 183
7.2.2
Laboratory Prototype of a Γ−Controller ..................................... 185
7.3...... Experimental Waveforms........................................................................ 188
7.3.1
Steady Sate Operation ................................................................. 189
7.3.2
Dynamic Operation ..................................................................... 194
7.4...... Summary ................................................................................................. 195
Chapter 8 Optimal Power Flow with AC Link Power Flow
Controllers Embedded .................................................................197
8.1...... Introduction ............................................................................................. 197
8.2...... Standard OPF ............................................................................................ 199
8.3...... OPF with Ξ−controllers embedded ......................................................... 201
xi
8.3.1
The model.................................................................................... 201
8.3.2
OPF
8.3.3
Amount and Location.................................................................. 204
8.3.4
Case Study................................................................................... 205
Formulation .......................................................................... 202
8.3.4.1
Economic Dispatch .............................................................. 206
8.3.4.2
Interface Power Maximization............................................. 208
8.4...... Summary ................................................................................................. 210
Chapter 9 Conclusions and Future Work................................. 211
9.1...... Contributions........................................................................................... 213
9.1.1
Modeling and control .................................................................. 213
9.1.2
Computer simulation................................................................... 213
9.1.3
Comparison against the state−of−the−art technology for series
compensation........................................................................................... 214
9.1.4
Experimental verification............................................................ 215
9.1.5
System level modeling ................................................................ 215
9.2...... Proposed Continuing Research ............................................................... 215
9.2.1
Feedback Control for Γ−controller ............................................. 215
9.2.2
OPF with
9.2.3
System level dynamics................................................................ 216
9.2.4
Multilevel realization for the Ξ−Controller ................................ 217
9.2.5
Additional topologies .................................................................. 217
embedded Γ−controllers............................................... 216
xii
9.2.6
Field demonstration..................................................................... 218
References ....................................................................................225
xiii
List of Figures
Figure 1-1 Example of a power system. ............................................................... 22
Figure 1-2 Schematic of a 2-bus system example................................................. 24
Figure 1-3 Schematic of a 3-bus system example................................................. 25
Figure 1-4 Schematic Illustrating Power Flow Control through a
Transmission Line. .......................................................................... 26
Figure 1-5 Power Flow Control Evolution............................................................ 29
Figure 1-6 Schematic diagrams of DC link realizations for power flow
control.............................................................................................. 31
Figure 1-7 Schematic of the Marcy Substation CSC.............................................. 34
Figure 1-8 Distributed Static Series Compensator (DSSC) Schematic. ................. 35
Figure 2-1 Circuit schematic of a three-phase VeSC synthesizing an
adjustable pole voltage obtained from N throws.............................. 40
Figure 2-2 Single−phase equivalent circuit of a VeSC. ......................................... 44
Figure 2-3 Region of pole voltage obtainable for different throw
realizations....................................................................................... 45
Figure 2-4 IGBT-Diode realization for the switch.................................................. 47
Figure 2-5 Waveforms Illustrating Pulse Width Modulation for Vector
Switching Converters. ..................................................................... 49
Figure 2-6 A Simulink/DSP strategy to generate the gate pulses described
in Figure 2-5. ................................................................................... 50
Figure 2-7 A four−winding three phase transformer to generate three
secondary vector voltages 120° phase−shifted................................ 51
xiv
Figure 2-8 A five−winding three−phase transformer to generate four
secondary vector voltages 90° phase−shifted.................................. 52
Figure 3-1 Schematic of a VeSC−based AC link intertie. ....................................... 55
Figure 3-2 Topology of a VeSC −based AC link intertie. ....................................... 56
Figure 3-3 Single−Phase averaged equivalent circuit........................................... 57
Figure 3-4 The ideal AC link intertie. .................................................................... 58
Figure 3-5 The “unity” rhombus. Permissible duty ratio values........................... 60
Figure 3-6 Variation of sets of duty ratios, (a), (b) and reactive power at
receiving end, (c) as a function of active power control while
Qs=0pu and θ=─20°. ....................................................................... 62
Figure 3-7 Variation of sets of duty ratios, (a), (b) and reactive power at
receiving end, (c) as a function of reactive power control
while PS =1.5pu and θ=─20°. .......................................................... 63
Figure 3-8 Variation of sets of duty ratios, (a), (b) and reactive power at
receiving end, (c) as a function of phase−shifting control
while PS =1.5pu QS=0pu. ................................................................. 64
Figure 3-9 Complex Power Nomogram at the sending end. (a) Entire
operating range, (b) restricted and more practical operating
range. ............................................................................................... 67
Figure 3-10 Phase−A voltage and current waveforms at the sending end............ 68
Figure 3-11 Phase−A voltage and current waveforms at the receiving end. ........ 69
Figure 3-12 Schematic of a VeSC−based AC link series compensator................... 71
Figure 3-13 Topology of a VeSC−based AC link series compensator.................... 72
Figure 3-14 Single−Phase Averaged Equivalent Circuit. ..................................... 73
Figure 3-15 The ideal AC link series compensator................................................ 74
Figure 3-16 Equivalent impedance (a) and power at the receiving end (b)
as a function of the duty ratio. ......................................................... 76
xv
Figure 3-17 Phase−A Current and Injected Voltage shown over a 60Hz
period. .............................................................................................. 79
Figure 3-18 Simulation results for a step−change of the duty ratio (a) from
20 to 80%. (b) and (c) show the response of the line current
and the power at the receiving end, respectively............................ 80
Figure 3-19 Simulation results for a ramp−change of the duty ratio (a)
from 20 to 80%. (b) and (c) show the response of the line
current and the power at the receiving end, respectively. .............. 80
Figure 3-20 Schematic of a VeSC−based AC link Unified Power Flow
Controller......................................................................................... 82
Figure 3-21 Topology of a VeSC−based AC link Unified Power Flow
Controller......................................................................................... 83
Figure 3-22 Single−Phase Averaged Equivalent Circuit. ..................................... 86
Figure 3-23 The ideal AC link Unified Power Flow Controller. ........................... 87
Figure 3-24 Receiving−end power control. .......................................................... 91
Figure 3-25 Tap−changing, phase−shifting control. ............................................. 92
Figure 3-26 Resonance frequency as a function of the capacitor value................ 94
Figure 3-27 Phase−A Voltage and Current at the receiving−end. ........................ 96
Figure 3-28 Phase−A Capacitor voltages.............................................................. 97
Figure 3-29 Phase−A SPT transformer currents. .................................................. 98
Figure 3-30 Duty Ratio Commands. Stepped from d1=33%, d2=35%,
d3=32% d4=0% to d1=0%, d2=18%, d3=57% d4=25%. ................... 99
Figure 3-31 Response to a duty ratio step change of the (a) line current,
(b) active power and (c) reactive power, all at the receiving
end. .................................................................................................. 99
Figure 3-32 Duty Ratio Commands. Ramped from d1=33%, d2=35%,
d3=32% d4=0% to d1=0%, d2=18%, d3=57% d4=25%. ................. 100
xvi
Figure 3-33 Response to a duty ratio ramp change of the (a) line current,
(b) active power and (c) reactive power, all at the receiving
end. ................................................................................................ 101
Figure 4-1 Schematic of a generic power system with a vector switching
converter embedded....................................................................... 109
Figure 4-2 Waveforms corresponding to the switching function, index {0}
approximation (duty ratio) and index {0, -N, N}
approximation. All with d=50%. .................................................. 113
Figure 4-3 Exact model for the Ξ−controller. ..................................................... 122
Figure 4-4 Waveform of the switched capacitor current along with the
traditional and augmented approximations. FP=60(Hz),
FS=1020(Hz), d=50%. ................................................................... 125
Figure 4-5 Eigenvalues distribution for the different models. FR=30(Hz),
FP=60(Hz), FS=1020(Hz)............................................................... 127
Figure 4-6 Exact model for the Γ−controller. ..................................................... 128
Figure 4-7 Capacitor voltage waveforms with FR=1862(Hz) and
FS=2400(Hz).................................................................................. 131
Figure 4-8 Capacitor voltage waveforms with FR=1862(Hz) and
FS=1800(Hz).................................................................................. 132
Figure 5-1 Configuration for classical control design......................................... 134
Figure 5-2 Block diagram of the control system................................................. 137
Figure 5-3 Block diagram for the small signal model......................................... 140
Figure 5-4 Bode Plot for the partial loop gain βg(jω)......................................... 143
Figure 5-5 Bode Plot for the PI controller h(jω). ................................................ 144
Figure 5-6 Gain and Phase margins from the loop gain
T(jω)= βh(jω)g(jω)........................................................................ 144
Figure 5-7 Response to a 15% step change in the reference power. (a)
shows the step change applied to the reference power while
(b) and (c) show the response on the line current and the
measured power at the receiving end, respectively. ...................... 146
xvii
Figure 5-8 Response to a 5% step change in the sending end voltage. (a)
shows the step change applied to the magnitude of the
sending end voltage while (b) and (c) show the response on
the line current and the measured power at the receiving end,
respectively.................................................................................... 147
Figure 5-9 Response to a 3° step change in the transmission angle. (a)
shows the step change applied to the transmission angle
while (b) and (c) show the response on the line current and
the measured power at the receiving end, respectively. ................ 147
Figure 5-10 Control system for the Γ−controller. ............................................... 148
Figure 6-1 Schematic of candidate compensators. (a) SSSC (b)
Ξ−controller................................................................................... 153
Figure 6-2 Converter realizations. (a) SSSC (b) Ξ−controller.............................. 154
Figure 6-3 Waveforms illustrating AC voltage synthesis. (a) SSSC. (b)
Ξ−controller................................................................................... 154
Figure 6-4 One line diagram of the candidate series compensation system. ...... 155
Figure 6-5 Steady sate equivalent circuits. (a) SSSC (b) Ξ−controller................. 157
Figure 6-6 Switch realization in terms of real semiconductors. For the
SSSC, 4 series IGCT with antiparallel diode (ABB 5SHX
08F4510) are used. The clamping diode is also realized
using 4 diodes in series (ABB 5SDF 03D4502). For the
Ξ−controller, 3 parallel IGBT with antiparallel diode (EUPEC
FZ600R65KF1) are used. .............................................................. 169
Figure 6-7 Pole current breakup. (a) SSSC, (b) Ξ−controller............................... 173
Figure 6-8 Voltage/Current injected into the transmission line. (a) SSSC,
(b) Ξ−controller. ............................................................................ 175
Figure 6-9 Dynamic equivalent circuits. (a) SSSC, (b) Ξ−controller. .................. 176
Figure 6-10 Feedback control loop. TS=5 (ms), β=10-6. (a) SSSC, (b)
Ξ−controller................................................................................... 177
Figure 6-11 Dc voltage equalizing algorithm ..................................................... 177
xviii
Figure 6-12 Small signal frequency response of loop gain of (a) SSSC, (b)
Ξ−controller................................................................................... 178
Figure 6-13 Computer simulation waveforms illustrating response to
command changes. (a) SSSC, (b) Ξ−controller............................... 179
Figure 6-14 Computer simulation waveforms illustrating response to bus
voltage and angle disturbances. (a) SSSC, (b) Ξ−controller. ......... 180
Figure 7-1 Schematic of the Γ−controller system implemented in the
laboratory....................................................................................... 182
Figure 7-2 Laboratory prototype of a transmission line...................................... 184
Figure 7-3 Laboratory prototype of a transmission line connecting two
stiff voltage busbars....................................................................... 185
Figure 7-4 Laboratory prototype of a Γ−controller............................................. 186
Figure 7-5 Detailed laboratory prototype of a Γ−controller system. ................. 187
Figure 7-6 Picture of hardware built at the laboratory....................................... 188
Figure 7-7 Attainable pole voltage (a) and the corresponding power
diagram (b). ................................................................................... 189
Figure 7-8 Experimental waveforms showing the gate voltage applied to
the IGBTs. ....................................................................................... 191
Figure 7-9 Waveforms showing the voltage at location V2. (a)
Experimental (b) Simulated........................................................... 191
Figure 7-10 Waveforms showing the line current. (a) Experimental (b)
Simulated....................................................................................... 192
Figure 7-11 Waveforms showing the injected voltage. (a) Experimental
(b) Simulated. ................................................................................ 193
Figure 7-12 Waveforms showing the fundamental component of the
injected voltage. (a) Experimental (b) Simulated.......................... 193
Figure 7-13 Waveforms showing the injected voltage at the switching
frequency timescale. (a) Experimental (b) Simulated. .................. 194
xix
Figure 7-14 Experimental waveforms illustrating the transition from one
operating point to another by means of duty ratio ramping. ......... 195
Figure 8-1. A 3−area electric power system. ...................................................... 198
Figure 8-2 Simplified Equivalent Circuit for the Ξ−controller........................... 201
Figure 8-3 The IEEE 39−bus system. ................................................................... 205
xx
List of Tables
Table 1-1 Complex Power Sensitivities................................................................ 27
Table 6-1 System Data ........................................................................................ 156
Table 6-2 Operating Modes ................................................................................ 158
Table 6-3 Injection Transformer Rating ............................................................. 163
Table 6-4 Capacitor Requirements Summary ..................................................... 167
Table 6-5 Devices Stress..................................................................................... 169
Table 6-6 Devices Capability.............................................................................. 170
Table 6-7 Device Count ...................................................................................... 170
Table 6-8 Converter Losses ................................................................................ 172
Table 6-9 Active state for semiconductors.......................................................... 173
Table 6-10 Analytical description for IT1A .......................................................... 173
Table 6-11 Converter Losses for PL=160(MW), QINJ=7.7(MVAr) .................... 174
Table 6-12 Total Harmonic Distortion (THD).................................................... 175
Table 6-13 Regulator Design Results ................................................................. 179
Table 7-1 Γ−controller component rating ........................................................... 186
Table 7-2 Operating Point................................................................................... 190
Table 8-1 Active Constraints .............................................................................. 206
Table 8-2 Interface Flows ................................................................................... 207
Table 8-3 Active Constraints .............................................................................. 207
Table 8-4 Interface Flows ................................................................................... 209
xxi
Table 8-5 Interface Flows ................................................................................... 209
22
Chapter 1 Introduction
1.1
General
The electric power industry has traditionally been divided into three sectors:
generation, transmission and distribution. At the generation stage, electricity is produced
from sources such as fossil fuel, nuclear fuel, or waterfalls. Then, the electric power is
injected into the transmission system which is configured as an interconnected network
of transmission lines and control devices. Finally, at the distribution stage, the power is
taken from the transmission system and delivered to consumers at specific standards of
quality and reliability. These three stages together −generation, transmission and
distribution− are generically known as a power system. Figure 1-1 illustrates a possible
realization of such a system.
Distribution
Transmission Network
Figure 1-1 Example of a power system.
23
Since the ultimate goal of a power system is to deliver power to consumers, a
perfectly valid question is why an interconnected transmission system is needed. Why
not deliver power via radial lines from local generators to local consumers without being
part of a meshed network? The real need for an interconnected transmission system is
twofold: (i) it allows for the pooling of power plants and load centers in order to
minimize the total power generation capacity and fuel cost and (ii) it improves the
reliability of power supply through redundancy.
Following this trend, transmission networks throughout the world have
developed into enormous, complex systems containing thousands of transmission line
segments interconnecting large geographical areas.
However, during the last few decades the electricity industry has had to face two
new issues. On the one hand, due to environmental pressures and right−of−way matters,
it is becoming increasingly difficult, if not impossible, to build new transmission lines in
order to satisfy the always increasing demand for electric power. As a result of this,
existing transmission systems are being pushed to operate close to their capacity limits
and transmission line congestion is being observed in several regions particularly those
with large load growth at large populated centers. On the other hand, due to the
deregulation of the electricity industry that has taken place, channeling power flows
through pre−established contractual paths has become a priority for the system
participants. This issue is known as the loop flow problem. These issues are briefly
reviewed in the following sections.
24
1.1.1
Congestion in Transmission Systems
The concept of congestion in transmission networks can be understood by
considering a simple case of a generator serving a load through two parallel transmission
lines, as suggested in Figure 1-2. Say the transmission line l1 has an impedance X and a
maximum capacity to carry power given by Pm. The second line, l2, has exactly the same
impedance as l1 but its maximum capacity is only 0.5Pm. Since both lines have the same
impedance, the power sent from the generator G to the load L equally splits into the two
transmission lines, as suggested in the figure. When the power demanded by L is less
than Pm, the system operates within its capacity margins since each line is carrying less
power than 0.5Pm. However, when the load L demands more power than Pm, say 1.2Pm,
each line carries 0.6Pm, which overloads the line l2. Notice that the combined capacity of
both lines is 1.5Pm, but the system is unable to carry such an amount of power due to the
congestion in one of the lines. Situations like this, in a more complex context, are
frequently encountered in modern power systems [1].
G
P
l1: X, Pm
L
P l2: X, 0.5Pm
Figure 1-2 Schematic of a 2-bus system example.
1.1.2
Loop Flows
The concept of loop flows, also known as parallel flows, is formally discussed in
Chapter 8, but here it can be described considering a simplified 3−bus system as
25
suggested in Figure 1-3. Say the generator G1 has a contractual obligation of serving the
load L2 by an amount of Pc. For that purpose, the generator “reserves” capacity on the
line 1−2 for that amount. However, since power flows obey Kirchhoff’s laws and not
contractual paths, only a fraction of such a flow (κPc) will actually follow the contractual
path and the remaining amount, (1−κ)Pc, will flow through the bus 3, creating an
undesired loop flow.
Figure 1-3 Schematic of a 3-bus system example.
Thus, the philosophy for modern transmission system operation can be
summarized as follows: “Optimal utilization of the existing transmission network subject
to both system limits and security requirements”. “Optimal” can mean, among other
things, minimum fuel cost, maximum loadability, or minimum loop flows, depending on
the goals pursued. System limits are threefold: thermal limits on transmission lines,
voltage profile limits, and stability limits. The security requirement, known as (n−1)
contingency criterion, is that the system must satisfy the same quality standard even
when any of its facilities has an outage. In order to satisfy these new needs, power
engineers have developed the concept of Flexible AC Transmission Systems, FACTS [1-3].
26
1.2
Power Flow Control through FACTS Technology
The basics of power flow control can be understood by considering a purely
inductive transmission line connecting two buses, as suggested in Figure 1-4(a).
(a) A purely inductive transmission line
(b) A transmission line with an embedded FACTS device
Figure 1-4 Schematic Illustrating Power Flow Control through a Transmission Line.
Using elementary phasor ideas it can be shown that the complex power flow
measured at the bus 1 of the line is given by,
S = P + j⋅ Q =
V1V2
V2 V V
sin θ + j ⋅ ( 1 − 1 2 cos θ)
X
X
X
(1-1)
where P and Q are the active and reactive power, respectively. It can be seen from
Equation (1-1) that value of the power flow depends upon four parameters, namely, the
magnitude of the voltage at each bus, V1 and V2, the transmission angle, θ, and the line
reactance, X. Consider now the insertion of a
FACTS
device as shown in Figure 1-4(b)1.
By means of a power electronics−based control, FACTS devices are able to modify one (or
1
The symbol shown in the figure has been introduced in [14] to represent a generic FACTS device.
27
more) of the parameters that define the power flow equation, and therefore realize power
flow control. This simple but powerful idea is the core of FACTS technology.
It is evident from Equation (1-1) that the modification of any of the parameters
affects both the active and reactive power simultaneously. Table 1-1 shows the sensitivity
functions of the complex power with respect to the system parameters for a typical set of
values encountered in a power system.
Table 1-1 Complex Power Sensitivities.
V1=V2=1pu, θ=10°, X=0.5pu.
∂→
∂↓
θ
X
V1
P
Q
1.97 0.35
-0.69 -0.06
0.35 2.03
As can be seen, the transmission angle and the line reactance have a much bigger
relative influence on the active power while the value of the bus voltage mostly affects
the value of the reactive power. Thus, the control of either the transmission angle or the
line reactance can be considered as active power compensation whereas the control of the
bus voltage can be considered as reactive power compensation. This decoupling effect
has been successfully used in other areas of power engineering as well, such as the
DC
power flow method [4, 5].
As will be discussed in the Literature Review Section, the power
electronics−based control can take different forms, with semiconductors with no gate
turn−off (only gate turn−on), or with power devices with gate turn−off capability,
depending on the application pursued [6].
28
1.3
Dynamic Compensation of Power System with
FACTS
Devices
Although the focus of this research is on steady state power flow control, it is
important to mention that
FACTS
devices can also enhance the various types of stability
limits of a power system [7]. With respect to the large−signal stability, it can be shown
by means of the equal area criterion [4], that
FACTS
devices can enlarge the stability
margin of a system upon a short−circuit fault. In the small−signal stability arena, it can
be demonstrated that
FACTS
are able to damp power oscillations that occur as a result of
minor system disturbances. Finally, voltage stability can also be improved as
FACTS
devices are able to reshape the nose−curve [5, 8].
1.4
Literature Survey
Although the concept of
FACTS
was introduced by Dr. Narain G. Hingorani as a
response of modern power systems needs, the aim of controlling power flows in
transmission network is not new. Ever since power systems were conceived, there was an
interest in controlling the power flows through the transmission network.
The literature review can be performed by following the evolution that the power
flow control technology has featured throughout the years. Figure 1-5 shows the various
technologies that have been applied to effectively realize power flow control.
29
Figure 1-5 Power Flow Control Evolution.
The earliest devices conceived for power flow control, well before the
introduction of
FACTS,
are the mechanically switched phase−shifter and tap−changer
transformers. The former control the transmission angle and therefore realize active
power flow control whereas the latter control the bus voltage to realize reactive power
compensation [4]. These devices, although still broadly used, have two major drawbacks.
Since they are mechanically switched, they are difficult to operate and have a slow speed
of response. Moreover, since they have a small number of tap positions, they can only act
in a discrete way. With the development of semiconductor devices, however, proposals
have been made to realize the switching with for example, back−to−back thyristors, in
30
order to improve the speed of response [9, 10]. In addition, hybrid approaches embedding
DC−AC
inverters have been proposed for the purpose of achieving continuous control [11,
12].
The invention of the mercury valve allowed the development of the high voltage
direct current (HVDC) transmission technology [7, 13]. Modern semiconductors such as
the thyristor, GTOs and
IGBTs
have also been use to realize HVDC systems. The basic idea
behind HVDC technology is to cascade a rectifier, a DC−link, and an inverter. The DC−link
can be either a capacitor or an inductor, leading to a voltage or current sourced system,
respectively. The control takes place in the rectification and inversion process, as the
firing angle (or the pulse duration) can be externally manipulated. These systems can
continuously control the active power and in the case of converters with gate turn−off
capability, they can realize independent reactive power flow control as well. In addition,
HVDC
interties feature the unique property of decoupling the AC dynamics of the systems
being interconnected and, as a result of this, they can be used to interconnect systems
operated at different frequency, as it is done, for example, in the Itaipu interconnection,
that joins the 60Hz Brazilian system with the 50Hz Paraguayan system. The major
disadvantage of these systems is the high rating needed for the semiconductors. They
have to carry the entire power throughput and consequently the cost is high.
Phase controlled devices correspond to the next technology that came about and
are consider the first generation of FACTS controller [14]. Based on thyristors, they come
in two flavors. The shunt connected device, called Static Var Compensator (SVC) [7, 14],
injects an adjustable amount of reactive power in to the system and as a result of the
31
sensitivities explained in Table 1-1 is credited as a voltage regulator. The series
connected device called Thyristor Controlled Series Capacitor (TCSC) [7, 14], can
continuously adjust the reactance of the transmission line hence realizing active power
flow control. Moreover, as these controllers do not involve any change of frequency to
realize the power flow function, they may be considered AC↔AC devices.
The development of high−power semiconductors with turn−off capability paved
the way to the next generation of
engineers from
SIEMENS
FACTS
devices. Led by Dr. Laszlo Gyugyi, a group of
have developed
the Static Synchronous Compensator
(STATCOM) [15], the Static Synchronous Series Compensator (SSSC) [16], the Unified
Power Flow Controller (UPFC) [17], and the Interline Power flow Controller (IPFC) [18].
Together, these devices constitute the family of voltage source
DC
link power flow
controllers that have been unified under the Convertible Static Compensator (CSC) [19].
Figure 1-6 shows the various ways these devices are embedded into a transmission line.
(a) The shunt connected STATCOM
(b) The series connected SSSC
Transmission line 1
DC
link
Transmission line 2
(c) The shunt−series connected UPFC
(d) The series−series connected IPFC
Figure 1-6 Schematic diagrams of DC link realizations for power flow control.
32
The
and
TCSC
STATCOM
i. e., the
and the
STATCOM
SSSC
have, respectively, an analogous function to the
provides voltage support and the
SSSC
SVC
injects quadrature
voltage into the line hence realizing active power flow control. The
UPFC
can
simultaneously and selectively provide voltage support, and realize active and reactive
power through the line. Lastly, the IPFC can provide series compensation to two (or more)
transmission lines and also allow transfer of real power between the cross−connected
lines. It is important to notice that for each of these devices the inverter realization as
well as the modulation strategy can take different forms [20].
Finally, a few devices that follow the approach of fast switching direct
AC↔AC
conversion without frequency change to realize power flow control have recently
appeared in the literature. The research has been conducted in terms of matrix converter
[21] based power flow controllers [22, 23] and in terms of more restricted configurations
such as the vector switching converter (VeSC) [24-28].
As may be observed from Figure 1-5 (left part), there has been a continuous push
and pull between
DC
link and
AC
link technology, driven mostly by the availability of
switching power semiconductors. This research focuses on the next leap introducing fast
switching FACTS devices with the potential of better performance and controllability.
33
1.5
State−
−of−
−the−
−Art
The current state−of−the−art for
FACTS
controllers may be conveniently divided
into two categories. The Convertible Static Compensator installed in the Marcy
Substation in New York may qualify as the state−of−the−art already available technology
while the concept of distributed
FACTS
devices may be eligible for the state−of−the−art
technology still in a research stage.
1.5.1
The Convertible Static Compensator
The Convertible Static Compensator (CSC), installed at the Marcy 345 kV
substation in Central New York, has been developed through a joint effort between the
New York Power Authority (NYPA), SIEMENS Power Transmission & Distribution, and
the Electric Power Research Institute (EPRI) [19]. It can be considered as the unification
of the various DC link FACTS devices described in Figure 1-6, as it can adapt to operate as
a
STATCOM, SSSC, UPFC
or
IPFC.
The general
CSC
configuration shown in Figure 1-7
consists of two 100MVA inverters, two 100MVA series transformers, and a single
200MVA shunt transformer with two identical secondary windings. The low voltage
disconnect switches allow utilization of the two inverters in various shunt and series
configurations, resulting in eleven different configurations.
34
Line 1
To New Scotland
Line 2
To Coopers Corners
Figure 1-7 Schematic of the Marcy Substation CSC.
This device is already operative and has proven its versatility in controlling the
Marcy 345kV bus voltage as well as in regulating the power flow on the Marcy−New
Scotland and Marcy−Coopers Corners lines.
1.5.2
Distributed FACTS
The concept of distributed
FACTS (D−FACTS)
has very recently been introduced
[29]. The idea is to use low rated power converters that can be (i) mass−manufactured,
(ii) massively dispersed, and (ii) remotely operated to realize power flow control.
The proposal, shown in Figure 1-8, consists of Distributed Static Series
Compensator (DSSC) modules realized by means of a small rated (~10 kVA) single phase
35
DC
link voltage source inverter and a single turn transformer, along with associated
controls, power supply circuits and built−in communications capability.
Figure 1-8 Distributed Static Series Compensator (DSSC) Schematic.
As shown in the above figure, the module consists of two parts that can be
physically clamped around a transmission conductor. The transformer and mechanical
parts of the module form a complete magnetic circuit only after the module is clamped
around the conductor. The weight and size of the
DSSC
module is low, allowing the unit
to be suspended mechanically from the power line. The module is self−excited from the
power line itself. The unit normally sits in bypass mode until the inverter is activated.
Once the inverter is turned on, the
DSSC
module can inject a quadrature voltage into the
line, allowing a “small” power flow control action. The overall system control function is
achieved by using a large number of modules coordinated through communications and
36
smart controls.
As can be seen, the state−of−the−art corresponds to
FACTS
devices based on
DC
link voltage source converters. The research proposed in this thesis represents an
alternative to this technology.
1.6
Outline of the Research
The background necessary to understand
AC
link power flow controllers is
provided in Chapter 2. The concept of the vector switching converter (VeSC) is
introduced, along with the corresponding converter model, equivalent circuit, switch
realization, and modulation strategy. In some of the applications pursued in this thesis,
VeSCs
commute between phase−shifted voltage sources of equal magnitude. For that
purpose, multiple−secondary phase−shifter transformers are also discussed.
Chapter 3 is the core of this thesis proposal. It introduces a new family of
FACTS
devices: the Π−controller, the Ξ−controller and the Γ−controller. Based on the vector
switching converter, these devices feature functional characteristics similar to the ones
encountered in the voltage source
HVDC
interties, the Static Synchronous Series
Compensator (SSSC) and the Unified Power Flow Controller (UPFC), respectively. The
control function, however, is realized by direct
AC↔AC
conversion without frequency
change, adopting the strategy of pulse width modulation (PWM).
In Chapter 4, generalized averaging theory is used to develop the dynamical
model for the controllers proposed in Chapter 3. Averaging theory is a generalization of
the concept of dynamic phasors and becomes especially useful when the natural modes of
37
the physical system are located in the vicinity to the switching frequency. It that case, it
gives a more accurate description of the system behavior, as it is able to capture the
dynamics of the waveforms’ ripple.
Using the dynamical model developed in Chapter 4, closed loop regulation is
assessed in Chapter 5. In the case of the Ξ−controller, the regulator design is performed
analyzing the frequency response of the small−signal transfer function. The ideas of
security margins for stability are used as design criteria. The Γ−controller regulator
design is briefly discussed and proposed as a part of the future work.
For the series compensation case, this new technology is contrasted against the
state−of−the−art in Chapter 6. A detailed comparative study between the Ξ−controller
and the SSSC shows that the use of VeSC−based power flow controllers may become a real
alternative for the next generation of FACTS devices.
In order to fully prove the concept of
VeSC−based
power flow controllers, a
laboratory prototype was built and tested. The Γ−controller was chosen for experimental
verification, as its topology is the most general of all three controllers. These results are
presented in Chapter 7.
Chapter 8 gives a system point of view of the power flow control problem,
incorporating the Ξ−controller into an optimal power flow (OPF) program. A case study
of the
IEEE
39−bus system shows that the economic dispatch as well as control of the
power exchange between control areas can be enhanced by embedding the Ξ−controller
into the transmission lines. Optimal power flow with Γ−controllers is proposed as a part
of the continuing research.
Finally, the conclusions and continuing work for this dissertation are given in
38
Chapter 9.
All computations throughout this thesis were executed using the
MATLAB
mathematical analysis software package. Simulations were developed using
MATLAB’s
simulator package SIMULINKTM.
39
Chapter 2 Vector Switching
Converters
2.1
Introduction
This chapter introduces the three−phase vector switching converters as well as
multiple−secondary phase−shifter transformers. The former represent the foundations to
realize
AC
link power flow control while the latter is an auxiliary tool needed in some
applications.
2.2
Three−
−Phase Vector Switching Converters
The principles of operation of three−phase vector switching converters (VeSC)
have been described in detail in [26]. Although their realization and applications for
power flow control can take various forms, herein the attention is focused on
point−to−point applications. More generalized configurations and operations may be
found in [26].
2.2.1
Converter Model
A schematic of an
N
throw single pole three−phase VeSC is illustrated in Figure
40
2-1. The system synthesizes an adjustable three phase pole voltage (VP(A-C)) by switching
among N stiff three−phase voltage sources (VT1(A-C), VT2(A-C)… VTN(A-C)).
VT1A IT1A
VT1B IT1B
VT1C IT1C
VT2A IT2A
t1A
SA
t2A
tNA
t1B
SB
VPA
IPA
VT2B IT2B
t2B
VPB
IPB
VT2C IT2C
tNB
VPC
IPC
VTNA ITNA
t1C
SC
t2C
VTNB ITNB
tNC
VTNC ITNC
Figure 2-1 Circuit schematic of a three-phase VeSC synthesizing an adjustable pole voltage obtained from N
throws.
The throws of the switches are assumed ideal for the purpose of the discussion
herein, as is common in preliminary functional analysis of switching power converters.
These assumptions include: (i) negligible forward voltage drop of the switch throws in
their on−state; (ii) sufficient on−state current carrying capacity and off−state voltage
blocking capacity commensurate and compatible with the voltage and current ratings of
the system; and (iii) negligible transition periods between open and closed of the switch
throws that permit repetitive high frequency switching.
Furthermore, the nominal
frequency of the three−phase voltages and currents (power frequency) are assumed
41
identical, in the absence of which there could be no net steady state power transfer among
them.
The voltages at the throw terminals of the switch are assumed stiff such that their
variations during a switching period can be neglected. Similarly, the switch pole currents
are assumed stiff such that their variations over a switching period can be neglected.
These assumptions essentially allow the focus to be on the power transfer process and the
functional features. In practical power converters, filter elements appropriately applied at
the input and output ports of the system would ensure that these assumptions are valid.
Snubber circuits may further be applied to provide for appropriate commutation features
during the finite transition periods between switch states. It is important to notice that in
AC
systems, the designation of the stiff voltage and stiff currents is arbitrary. Stiff voltage
sources and stiff current sources can be transformed into either by adding series inductors
or shunt capacitors, respectively, in order to fit application considerations.
As indicated in Figure 2-1, the three poles are ganged together i.e., they switch
concurrently between the throws to which they are connected. This can be
mathematically expressed as tiA = tiB = tiC = ti, for i = 1, 2…N.
The transfer properties can be represented using the vector equations consisting of
three components for each of
the terminal quantities. In this case, the vectors
representing the throw voltages, throw currents, pole voltages and pole currents become
VTi = [VTiA VTiB VTiC]T, ITi = [ITiA ITiB ITiC]T, VP = [VPA VPB VPC]T, IP = [IPA IPB IPC]T
respectively, for i=1,2…N. Thus, the pole voltage and throw currents can be expressed
as,
42
N
VP (t) = ∑ q i (t) ⋅ VTi
(2-1)
i =1
I Ti (t) = q i (t) ⋅ I P
for i = 1, 2,…N
1
where q i (t) = 
0
if t i is closed
otherwise
(2-2)
is the switching function of a throw connecting the
voltage VTi to the current IP. When the repetition frequency of the switching function (or
simply the switching frequency) is large with respect to the power frequency, net power
transfer between the voltage ports and the current ports arises from the average value (DC
component) of the switching function. The DC component of the switching function may
be readily represented by the duty ratio of the particular throw. Thus, under proper
assumptions (discussed in Chapter 4), the transfer relationships (2-1) and (2-2) may be
simplified as
N
VP (t) = ∑ d i (t) ⋅ VTi
(2-3)
i =1
I Ti (t) = d i (t) ⋅ I P
for i = 1, 2,…N
(2-4)
where the duty ratio of the ith throw is defined as
1
d i (t ) =
TS
t + TS
∫ q ( τ ) ⋅ dτ
i
for i = 1, 2...N
t
where TS is the switching period. As may evident from the definition of q(t),
(2-5)
43
N
∑ d i (t) = 1
(2-6)
i =1
Since the operating principle of these converters is based on controlling the
connectivity between various three−phase AC voltage and/or current vectors by switching
among their components concurrently, these converters have been termed vector
switching converters [26]. The distinguishing features of vector switching converters as
compared to matrix switching converters have been explored also in [26].
2.2.2
Equivalent Circuit
Relationships (2-3) and (2-4) indicate a reciprocal input↔output transfer property
similar to that of a transformer. Therefore, the fundamental component averaged vector
(or single−phase) equivalent circuit of the converter system may be represented as shown
in Figure 2-2. Figure 2-2(a) describes the input↔output reciprocal relationships
explicitly as dependent sources, while Figure 2-2(b) implicitly represents them in the
form of a magnetically coupled transformer with winding turns ratios equivalent to the
duty ratios of the throws.
44
VT1
VT1
d1IP
IT1
1/d1
d1VT1
+
VT2
d2IP
+
dNVT1
VTN
VT2
d2VT1
IP
+
dNIP
VP
IT2
1/d2
VTN
1
IP
ITN
1/dN
(a) Controlled source coupling
(b) Magnetic coupling
Figure 2-2 Single−phase equivalent circuit of a VeSC.
As is evident in either realization, the net pole voltage, VP, depends on the value
(magnitude and phase) of the throw voltages as well as the corresponding duty ratios.
These equivalent circuits may be conveniently utilized to derive and study the application
of three−phase vector switching converter for
AC
power flow control by incorporating
them in power system load flow and dynamic analysis programs.
Figure 2-3 shows the realizable pole voltage values, represented by the gray
shaded region in the phasor plane, for different combinations of throw voltages and
number of throws.
45
VT3
VT1
VT3
VT4
VT1
VT2
VˆP
VT4
VT2
(a) Quadruple−throw switch with random
magnitude/phase throw voltages
(b) Quadruple−throw switch with random
magnitude/phase throw voltages and one redundant
throw voltage.
VT2
VT2
VˆP
VˆP
VT3
VT1
VT3
VT1
VT4
(c) Triple−throw switch with identical
magnitude, 120° phase−shifted throw voltages
(d) Quadruple−throw switch with identical magnitude,
90° phase−shifted throw voltages
Figure 2-3 Region of pole voltage obtainable for different throw realizations.
Figure 2-3(a) and Figure 2-3(b) show realizable pole voltage with arbitrary
voltages (VT1–VT4) available for each of the throws (1−4). It may be observed that the
region of the realizable pole voltages is given by the largest polygon whose vertices are
the locations of the corresponding throw voltages and the phasor origin. For power flow
control applications [14, 19], one is usually interested in synthesizing voltages within
symmetric regions in all directions about the origin, which would be given by the largest
circle that can be inscribed inside the polygon about the origin.
It can be seen in Figure 2-3(a) that there is a large unused area outside the
circumference. Figure 2-3(b) shows that not only is it impossible to trace a complete
46
circle circumference but also that there is a redundant voltage source, namely VT2. On the
other hand, Figure 2-3(c) and Figure 2-3(d) show symmetric realizations with three
throws and four throws, respectively. The adaptation of these realizations for power flow
control will be explored in Section 2.3.
2.2.3
Switch Realization
Although the operation of the VeSC converter presented in Figure 2-1 does not
require the three−phase
AC
sources and loads to be balanced, the analytical development
presented herein is limited to the case of balanced sources and loads. Since balanced
three−phase
AC
systems may operate as three wire systems, the reference terminal that
provides the return path for the three−phase currents carries no current (shown in gray
line on Figure 2-1) and may be eliminated. As a result, due to inherent symmetry in
three−phase voltage and current waveforms, when all the three−phase
AC
ports are three
wire systems, the throws may be realized using bi−directional current conducting, but
unidirectional voltage blocking capability as illustrated in Figure 2-4 in terms of
IGBT−diode
pairs. Other devices such as GTOs may be used instead of IGBTs in order to fit
rating considerations.
47
Figure 2-4 IGBT-Diode realization for the switch.
2.2.4
VeSC Pulse Width Modulation
Pulse Width Modulation (PWM) is widely used to create the train of pulses that
define the converter switches
output waveform [30, 31].
ON
times, which create a desired low−frequency target
The basic strategy is to switch at the intersection of a
reference signal and a high−frequency carrier signal. The reference signal carries the
information that defines the properties of the target waveform whereas the carrier signal
defines the switching frequency. The main advantage of
PWM
is that it moves the
harmonics generated from the switching process to frequencies high enough to make
filtering possible with smaller components [32]. All PWM algorithms have to address the
following three fundamental issues [30]:
•
Switch pulse width calculation
•
Switch pulse position within a carrier interval
•
Switch pulse sequence within a carrier interval
48
Thus, depending on the strategy selected, the target output waveform will feature the
properties needed for the specific application.
For power flow control applications using the vector switching converter, both
input and output waveforms have the same frequency, therefore there is no frequency
change and, similarly to the case of
PWM DC/DC
converters, a simple algorithm that
compares the duty ratio against a sawtooth carrier signal can be used to generate the
pulses.
Figure 2-5 graphically shows the process to determine the width of the pulses for
a triple throw single pole switch. As can be seen, it uses a predetermined sequence (d1, d2
d3 in this case) and turns a switch on every time the accumulated references intercept the
carrier signal. In addition, every time a switch is turned on, the previous switch in
sequence in turned off.
The practical implementation of this strategy is presented in Figure 2-6 in a
Simulink/DSP environment. It takes the duty ratios as inputs and compares the
corresponding summations against the carrier signal. The output of these comparisons
are transformed into Boolean variables and then manipulated as shown in the figure in
order to fit the restriction of turning off the previous switch in the sequence when the
current switch in turned on. Finally, the various signals are transformed back into real
numbers to generate the pulses.
This process can be readily generalized to any number of throws by simply
adding more duty ratios in Figure 2-5 and Figure 2-6. In addition, in order to fit practical
applications, adequate dead times should be added to avoid switches
ON−states
49
overlapping.
Figure 2-5 Waveforms Illustrating Pulse Width Modulation for Vector Switching Converters.
The performance of this strategy clearly depends on the sequence selected to
aggregate the duty ratios. Any arrangement would produce the same fundamental
50
frequency waveforms, but the harmonics content would vary depending on the sequence
selected. Optimal sequence arrangement can be designed for a specific application but
such a topic is beyond the scope of the present study.
1
d1
2
d2
3
d3
>
+
+
+
+
+
>
>
True=1
False=0
NOT
NOT
AND
double
1
P1
double
2
P2
AND
Boolean Stage
double
Back to Real
Numbers
3
P3
Carrier
Figure 2-6 A Simulink/DSP strategy to generate the gate pulses described in Figure 2-5.
2.3
Multiple−
−Secondary Phase−
−Shifter Transformer
The synthesis of multiple throws is an essential requirement for using vector
switching converters as power flow controllers some applications. Indeed, as can be seen
in Figure 2-3, the adjustable pole voltage is realized by switching among different stiff
voltage sources.
Three−phase systems feature the property that, given one set of three−phase
voltages, it is possible to synthesize multiple electrically isolated sets of voltages using
multi−winding transformers.
Consider the four−winding three−phase transformer presented in Figure 2-7(a).
The windings drawn parallel are physically located in the same magnetic path. All
51
windings are Y−connected and one of them corresponds to the primary while the
remaining three make a multiple−secondary set. Since there are three sets of three−phase
windings at the secondary side, three isolated voltage sets are created, say {A1, B1, C1},
{A2, B2, C2} and {A3, B3, C3}. Moreover, as a result of the convenient phase designation
given to the various windings, the corresponding phase−to−neutral voltage of each set is
120° phase−shifted with respect to each other, as shown in the corresponding phasor
diagram presented in Figure 2-7(b). Thus, using vector notation, the various sets can be
expressed as VT1=[A1 B1 C1]T, VT2=[A2 B2 C2]T and VT3=[A3 B3 C3]T. It readily follows
that the phasor diagram for the vectors VT1, VT2 and VT3 is as shown if Figure 2-3(c) and
therefore this four−winding three−phase transformer can be used to generate such throws.
n2
n1
n2
n2
Figure 2-7 A four−winding three phase transformer to generate three secondary vector voltages 120°
phase−shifted.
52
The major drawback of the topology introduced above is the relatively large
unused area outside the circle in Figure 2-3(c). Specifically, for a given magnitude of the
throw voltages, say VT, the maximum pole voltage that can be obtained is V̂P = VT / 2 .
This limitation can be improved considering the topology sketched in Figure
2-8(a). The transformer in this case has five three−phase windings. The primary side
winding is ∆−connected while in the secondary side there are two windings ∆−connected
and two Y−connected. In addition, as suggested in Figure 2-8(a), the turn ratio of the
∆−connected windings is 3 times the turn ratio of the Y−connected windings. Thus,
with the phase designation given in Figure 2-8(a), it is possible to create four vector
throws phase−shifted 90° with respect to each other, as shown in the phasor diagram of
Figure 2-8(b).
n2
n2
3
n1
n2
n2
3
Figure 2-8 A five−winding three−phase transformer to generate four secondary vector voltages 90°
phase−shifted.
53
Similar to the case with three secondary windings, by defining the vector voltages
as VT1=[A1 B1 C1]T, VT2=[A2 B2 C2]T, VT3=[A3 B3 C3]T and VT4=[A4 B4 C4]T, it can
readily be seen that they resemble the phasor diagram shown in Figure 2-3(d). For this
topology, the maximum obtainable pole voltage is V̂P = VT / 2 , which represents a
significant improvement with respect to the topology with three secondary windings.
In order to obtain larger useful pole voltage, other configurations with six or more
secondary winding might be developed but they lack practicality due to the large number
of windings needed.
54
Chapter 3
3.1
AC
Link Power Flow
Controllers
Introduction
This chapter introduces the family of power flow controllers proposed in this
research work: the Π−controller [24], the Ξ−controller and the Γ−controller [25]. These
devices are based on the vector switching converter (VeSC) and feature similar functional
properties as some of the
DC
link controllers discussed Chapter 1. Specifically, as will
become evident through this chapter, there is an intimate correlation between the
Π−controller, the Ξ−controller and the Γ−controller and the
HVDC
intertie, the Static
Synchronous Series Compensator (SSSC) and the Unified Power Flow Controller (UPFC).
Although functionally AC link controllers are similar to their DC link counterparts,
the control function is realized by direct
AC↔AC
conversion without frequency change,
adopting the strategy of pulse width modulation (PWM).
As stated in Chapter 1, the
HVDC
intertie, the
SSSC
and the
UPFC
appeared in the
literature chronologically in that order, so in this chapter the same tendency is chosen,
introducing first the Π−controller, followed by the Ξ−controller and the Γ−controller.
55
3.2
Π−Controller:
An
AC
Link
Back−
−to−
−Back
Intertie
DC link interties have been extensively used to interconnect
AC
systems
throughout the world. They feature the unique property of decoupling the dynamics of
the systems being interconnected and, as a result, they can be used to interconnect
systems operated at different frequencies. However, the most important property of these
devices is the ability to continuously and independently control the active and reactive
power flow through them. The latter property can also be achieved by using
AC
link
interties based on the vector switching converter, as suggested in Figure 3-1.
Figure 3-1 Schematic of a VeSC−based AC link intertie.
In this section, the details on the implementation of the vector switching converter
to function as an intertie are provided, along with computer simulation results that
validate the approach.
3.2.1
Topology
The schematic of the configuration proposed is presented in Figure 3-2. The main
56
goal is to control the power flowing across the two three−phase
AC
systems, namely the
sending System S and the receiving System R. As indicated in the figure, the intertie is
configured by including a phase−shifting transformer (SPT), filter capacitors (TFC), a
quadruple throw−single pole three−phase VeSC, and boost transformer (BIT).
Figure 3-2 Topology of a VeSC −based AC link intertie.
The transformer
SPT
has the double function of lowering the voltage to the level where
the switches can safely operate and of simultaneously phase−shifting the input voltage in
0, 90, 180 and 270° to generate four sets of three−phase voltages of equal magnitude. The
capacitor bank
TFC
is used to absorb the high frequency currents introduced due to
switching process from permeating into the system across the transformer. The controlled
voltage is synthesized at the poles by modulating the duty ratios d1, d2, d3, d4. Finally, the
transformer BIT located at the receiving side raises the voltage to the corresponding level
57
of the receiving end.
3.2.2
Equivalent Circuit
By following the procedure explained in Chapter 2, the single−phase averaged
equivalent circuit shown in Figure 2-3 can readily be obtained. As can be seen, the
controlled source approach has been chosen to represent the converter. In addition, the
turn ratios of transformer
STP
have been made complex numbers in order to reflect the
corresponding phase−shift. Notice as well that the Systems S and R have been replaced
by their Thévenin equivalent circuits. Although resistive elements are not shown in the
figure, they may be easily added to represent the power losses throughout the various
stages.
Figure 3-3 Single−Phase averaged equivalent circuit.
This equivalent circuit may be conveniently utilized to derive and study the
58
application of VeSC−based interties for
AC
power flow control by incorporating them in
power system load flow and dynamic analysis programs.
3.2.3
Principle of Operation
In order to understand the principle of operation, consider a purely inductive link
between the sending end system S and the receiving end system R, as indicated in Figure
3-4(a). Also, assume that the voltages VS and VR have equal magnitude, say V, and they
are displaced by a transmission angle of, say θ°.
(a) Systems S and R connected via a purely
inductive link
(b) System S and R connected via an inductive link plus
the intertie.
Figure 3-4 The ideal AC link intertie.
Under these assumptions, the power flow equations are
V2
sin θ
P=
X
Q S = −Q R =
(3-1)
V2
(1 − cos θ)
X
(3-2)
59
Now consider a fully ideal
AC
link intertie i.e., both transformer
SPT
and
BIT
are
ideal and have unity turn ratio and also the reactive power injected by the capacitor bank
FTC
is negligible so it may be removed. Under these assumptions, the equivalent circuit
of the intertie sketched in Figure 3-2 may be further reduced to the equivalent circuit
presented in Figure 3-4(b). The turn ratio of the transformer is given by
D = (d1 − d 3 ) + j (d 2 − d 4 ) = d13 + j d 24
and reflects both the phase−shift introduced by
(3-3)
SPT
transformer as well as the action of
the converter. Now the power flow equations take the form of,
V2
[d13 sin θ + d 24 cos θ]
X
V2 2
QS =
(d13 + d 224 − d13 cos θ + d 24 sin θ)
X
P=
QR =
V2
(d13 cos θ − d 24 sin θ − 1)
X
(3-4)
(3-5)
(3-6)
Equations (3-4) through (3-6) along with restriction imposed by the continuity of the
converter i. e.,
d1 + d 2 + d 3 + d 4 = 1
(3-7)
fully described the ideal AC link intertie. Equations (3-4) through (3-6) also show that the
60
duty ratio differences d13 and d24 give two degrees of freedom that can be conveniently
chosen to control P and Q at one of the ends.
In addition, as can be seen from Figure 3-4, the equivalent circuit has two legs in
shunt that resemble the Greek letter pi, Π, so the
AC
link intertie can be termed
Π−controller.
Although it is strictly valid only for the ideal case, it is convenient for
visualization purposes to represent the feasible range of values for the various duty ratios
in d13─d24 plane, as shown in Figure 3-5. Values inside the “unity” rhombus only are
valid for the duty ratio differences, given the restriction of non−negativity for each duty
ratio (di≥0, for i=1..4) as well as the continuity equation. In the real case, the response of
the converter also depends on d i2 factors, but it is largely dominated by the duty ratio
differences, that is, a given set point {P,Q} is determined mostly by a duple {d13, d24}.
Figure 3-5 The “unity” rhombus. Permissible duty ratio values.
61
3.2.4
Control Capabilities
Although the model presented in the previous section demonstrates the control
capabilities, a more realistic model would account for the effect of the transformers as
well as the capacitor bank.
A 110kV/100MVA/60Hz case−study system was developed to explore the
performance of the intertie further. In reference to Figure 3-2, consider the system S
(modeled as a stiff voltage source of VS = 1∠0 pu) is connected to the system R (modeled
as a stiff voltage source of VR = 1∠ − 20 pu) through a resistive−inductive link of
Z L = 0.34∠86 pu. Thus, without the Π−controller embedded, one would measure
{Ps=1pu Qs=0.1pu} and {PR=0.99pu QR=─0.2pu} at the sending and receiving end,
respectively. This will be considered as the base case for the following discussion.
Now consider the insertion of the Π−controller, say at the midpoint of the link,
splitting it into two equal impedances ( ZS = Z R = 0.17∠86 pu). The turn ratio of
transformers
SPT
and
BIT
were selected to be ap=10:1 and as=1:20, respectively. Typical
values were selected for the parasitic elements of the transformers. Finally, the capacitor
bank was selected to be XC=3pu. Refer to the Design Considerations Section for further
details on the selection of these values and to the Appendix 1 for the complete list of
parameters.
Next, based on the operating point described above, various control strategies are
discussed. The attention is focused on the sending end control variables, namely, PS and
62
QS.
3.2.4.1 Active Power Control
Figure 3-6(a) and (b) show the values of d13 and d24 that must be set in order to
transmit active power in the range {-2pu, 2pu} shown on the horizontal axis, with zero
reactive power maintained at the sending end. Figure 3-6(c) shows the corresponding
reactive power needed at the receiving end.
d1 - d3
1
0
-1
d2 - d4
1
0
-1
QR(pu)
1.5
1
0.5
0
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
Figure 3-6 Variation of sets of duty ratios, (a), (b) and reactive power at receiving end, (c) as a function of
active power control while Qs=0pu and θ=─20°.
It can be seen that control points around PS=0pu are achieved at a high
requirement of reactive power at the other end, therefore becoming impractical.
However, that can be solved by local reactive power injection at the bus, as it is done in
some
HVDC
ties [13]. Although the reactive power at the sending end was set at QS=0pu
in this example, it can be set at any other reasonable value that may be desired.
63
Since the intertie can transfer power in both directions, it can be named as a
back−to−back system and the terms sending and receiving ends become interchangeable.
3.2.4.2 Reactive Power Control
Figure 3-7(a) and (b) show the values of d13 and d24 that must be set in order to
regulate the reactive power at the sending end within the range {-1pu, 1pu} shown on the
horizontal axis, with the active power transfer fixed at PS=1.5pu. Figure 3-7(c) shows the
corresponding reactive power needed at the receiving end.
d1 - d3
1
0
-1
d2 - d4
1
0
-1
QR(pu)
1
0
-1
-1
-0.8
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
0.8
1
Figure 3-7 Variation of sets of duty ratios, (a), (b) and reactive power at receiving end, (c) as a function of
reactive power control while PS =1.5pu and θ=─20°.
As can be seen from Figure 3-7(c), the intertie can provide voltage support (QS<0
and QR>0) at both ends while keeping a fixed level of active power transfer.
64
3.2.4.3 Angle Control
Figure 3-8(a) and (b) show the values of d13 and d24 that must be set in order to
keep a given transfer level while the transmission angle θ is varied within the range {90°, 90°} shown on the horizontal axis. Figure 3-8(c) shows the corresponding reactive
power needed at the receiving end. The excellent performance of the intertie as a
phase−shifting controller is evident from Figure 3-8(c). It can keep PS, QS and QR at a
fixed level for a wide transmission angle range.
1
0
-1
1
0
-1
1
0.5
0
-80
-60
-40
-20
0
20
40
60
80
Figure 3-8 Variation of sets of duty ratios, (a), (b) and reactive power at receiving end, (c) as a function of
phase−shifting control while PS =1.5pu QS=0pu.
65
3.2.5
Design Considerations
3.2.5.1 Parameter Calculations
As a general rule, power converters need filter capacitors across the throw
terminals and filter inductors in series with the pole terminals [33]. In the Π−controller,
the pole terminals appear in series with the boost transformer
SIT
and the transmission
line. These elements provide enough series inductance naturally, eliminating the need for
separate filter reactors. On the other hand, at the throw terminals, filter capacitors are
needed to isolate the high frequency currents introduced due to the switching process
from propagation into the transformer
SPT,
and further into the system. However, as the
capacitors bank appears directly shunt−connected into the systems, it ends up playing an
undesired major role on the reactive power balance of the entire system.
Furthermore, the selection of the turn ratios of both phase shifting and boost
transformers also plays an important role in system operation. The turn ratio of the
transformer at the sending end also must be compatible with voltage levels where the
switches can safely operate. Hence, the capacitors bank and both transformer turn ratios
must be selected to simultaneously fulfill the various system requirements.
Detailed time−domain simulations showed that the capacitive reactance needs to
be less than 5pu in order to realize satisfactory system performance. With that bound, and
recognizing the operation at zero reactive power at both ends to be a desirable target
under a wide range of power flow conditions, the process of finding an acceptable set of
design parameters can be formulated as an optimization problem. By running
66
optimizations at various points within the range of active power control (-2pu to 2pu in
this case), it is possible to converge to a set of values that satisfactorily meet the system
requirements. For the system data specified in Appendix 1, XC=3pu and turn ratios
aP=0.1 and aS=20 give acceptable performance.
3.2.5.2
Feasible operating range
The full feasible operating range for the Π−controller can be obtained by mapping
the unity rhombus from Figure 3-5 into the PS−QS plane, which is sketched in Figure
3-9(a). However, this operating nomogram is only theoretical, due to the enormous
values of reactive power needed to realize some of the operating conditions. Thus, only a
subset of these operating conditions is practical. By discarding all the impractical points
(accepting a maximum of 1pu of reactive power at both ends), the operating nomogram
becomes restricted as illustrated in Figure 3-9(b). This corresponds to a subset of the
region from Figure 3-9(a). The axis of the figure has been changed to emphasize the
practical feasible region.
It can be seen that the resulting restricted operating region still provides a wide
range of controllability, with a limited variation in duty ratios, especially with low levels
of reactive power requirements, validating the proposed design process.
67
6
4
2
0
-2
2
1
0
-1
-2
-3
-2
-1
0
1
2
3
Figure 3-9 Complex Power Nomogram at the sending end. (a) Entire operating range, (b) restricted and
more practical operating range.
3.2.6
Simulation Results
A detailed computer simulation model of the system shown in Figure 3-2 was
developed to verify the operation of the Π−controller at various operating conditions.
The switching frequency to operate the converter was selected to be FS=2.4kHz. The base
values and per unit values of the parameters used in the computer simulation example are
shown in Appendix 1.
As an illustration, consider the base case described in the Control Capabilities
Section. It was stated that, without the Π−controller embedded, one would measure
{Ps=1pu Qs=0.1pu} at the sending end. Now, by inserting the Π−controller, the operating
point described by {Ps=2pu Qs=0pu} is selected as target, that is, the active power
transfer will be doubled while reducing to zero the requirements of reactive power. From
68
either the power nomogram or the equivalent circuit, the values for the duty ratios for this
condition are easily obtainable: d1=66%, d2=17.4%, d3=16.6%, d4=0. With these duty
ratio values, the simulation is executed, whose results are described next.
Figure 3-10 shows the terminal variables at the sending end over a 60Hz cycle.
As can be seen, the voltage and current appear approximately in−phase so the reactive
power is zero. In addition, these waveforms have been perunitized with respect to their
maximum base values, therefore it can be implied that the active power is effectively
PS=2pu, as predicted early.
Figure 3-10 Phase−A voltage and current waveforms at the sending end.
Figure 3-11 shows the same waveforms but at the receiving end. Although the
fundamental components agree with the predictions made by the equivalent circuit, the
69
main drawback of this approach becomes evident. The rich harmonics content makes the
feasibility of the Π−controller dependent upon big−sized output filters, which may be
impractical at the power level considered.
Figure 3-11 Phase−A voltage and current waveforms at the receiving end.
3.2.7
Summary
This section has introduced the Π−controller: an application of the vector
switching converter acting as a back−to−back intertie. A simplified illustration of the
system was used to demonstrate the principles of power flow control using duty ratio
modulation. Reactive power control, active power flow control and voltage phase shift
were variables identified as controllable. The capabilities of the system under different
modes of operation were demonstrated using more detailed transformer equivalent circuit
70
model. The variation of duty ratio and power flow quantities under three different
operating modes was demonstrated. The results from a computer simulation of the
system including the dynamics of the switching process were presented to validate the
approach. A brief discussion of design considerations indicates the trade−offs in the
design of converter parameters.
Although the capabilities of operation of the converter have been illustrated by
analytical methods and their feasibility has been demonstrated using computer
simulations, the power rating of the converter system and harmonics content are seen as
major limitations of the approach. Therefore, further augmentation of the approach by
embedding the converter in a shunt−series configuration such as the
UPFC
is seen to be
advantageous in reducing the power converter throughput in relation to the controlled
power flow. The operation of such a configuration is the subject of Section 3.4.
In addition, since with present technology this
FACTS
device seems impractical, it
will not be discussed further in this thesis.
3.3
Ξ−Controller: An AC Link Series Compensator
The concept of transmission lines compensation by means of series capacitor has
its roots in elementary phasor ideas. Transmission lines have inductive reactance that can
be compensated by adding series capacitive reactance. This very simple concept has
prompted many different devices throughout the years. First, mechanically switched
fixed capacitors were installed onto transmission lines to provide a static amount of
71
reactive compensation. Then, with the development of power semiconductors, devices
such us the Thyristor Controlled Series Capacitor (TCSC) and the Static Synchronous
Series Compensator (SSSC) have been developed. The former has the ability of
continuously controlling the equivalent impedance of the line whereas the latter can
inject a variable amount of quadrature voltage into the line. Both devices though seek the
same ultimate goal of controlling the active power through the line.
In this section, yet another approach to seek the same goal is described. Using the
vector switching converter as a controller, it is possible to modulate the amount of
capacitance injected into a transmission line, as suggested in Figure 3-12.
Figure 3-12 Schematic of a VeSC−based AC link series compensator.
This device has already been introduced in the literature [34] in a slightly
different context. In this thesis, it is adapted onto the formalisms of
AC
link vector
AC
link series
switching power flow controllers.
3.3.1
Topology
Figure 3-13 shows the detailed three−phase realization of the
72
compensator. As can be seen, the double throw−single pole three−phase VeSC commutes
between a fixed capacitor bank and a short circuit node. As a result of the switching, a
controlled amount of capacitance is generated which is injected into the line via the
SIT
transformer.
Figure 3-13 Topology of a VeSC−based AC link series compensator.
Notice as well that the requirements of stiff voltage at the throw side and stiff
current at the pole side are naturally given by the compensation capacitors and the
leakage inductance of the SIT transformer, respectively.
3.3.2
Equivalent Circuit
By following the procedure explained in Chapter 2, the single−phase averaged
73
equivalent circuit shown in Figure 3-14 can readily be obtained. As can be seen, the
controlled source approach has been chosen to represent the converter. Notice as well
that the systems S and R, the transmission line impedances, and the secondary−side
impedance of the transformer (the one in series with the transmission line), have all been
aggregated into Thévenin equivalent circuits.
Figure 3-14 Single−Phase Averaged Equivalent Circuit.
This equivalent circuit may be conveniently utilized to derive and study the
application of
VeSC−based AC
link series capacitors for
AC
power flow control by
incorporating them in power system load flow and dynamic analysis programs.
74
3.3.3
Principle of Operation
In order to understand the principle of operation, consider the same situation as in
Section 3.2.3, repeated here for convenience. A purely inductive transmission line is
connecting the sending end system S and the receiving end system R, as indicated in
Figure 3-15(a). Also, assume that the voltages VS and VR have equal magnitude, say V,
and they are displaced by a transmission angle of, say θ°.
(a) Systems S and R connected via a purely
inductive transmission line
(b) System S and R connected via a transmission line plus
an AC link compensation capacitor.
Figure 3-15 The ideal AC link series compensator.
Under these assumptions, the active power flow is given by
V2
P=
sin θ
X
(3-8)
Now consider the
AC
link series compensator is connected into transmission line
by an ideal transformer. Under this assumption, the equivalent circuit of the converter
and the
SIT
transformer can be combined into one as suggested in Figure 3-15(b). By
75
reflecting the capacitor into the transmission line−side of the transformer, the active
power flow equation takes the form of,
P=
V2
sin θ
X − d2XC
(3-9)
where X C = 1 / ω p C is the 60Hz impedance of the capacitor. It becomes evident that by
modulating the duty ratio of the converter, it is possible to control the active power flow
through the transmission line.
In addition, as can be seen from 3-15(b), the equivalent circuit has two legs in
series that resemble the Greek letter xi, Ξ, so the
AC
link series compensator can be
termed Ξ−controller.
3.3.4
Control Capabilities
A 110kV/100MVA/60Hz case study system was developed to explore the
performance of the Ξ−controller further. In reference to Figure 3-13, consider the system
S (modeled as a stiff voltage source of VS = 1∠0 pu) connected to the system R (modeled
as a stiff voltage source of VR = 1∠ − 20 pu) through a transmission line of
Z L = 0.24∠85 pu. Thus, without the Ξ−controller embedded, the active power at the
receiving end would be PR=1.4pu. This will be considered as the base case for the
following discussion.
76
Now consider the insertion of the Ξ−controller splitting the transmission line into
impedances ( Z S = 0.17∠86 pu and Z R = 0.07∠82 pu). The turn ratio of the
SIT
transformer was selected to be 1:1. Typical values were selected for the parasitic
elements of the transformer. Refer to the Appendix 2 for the complete list of parameters.
The reactance of the compensation capacitor was selected to be XC=0.05pu so it can
provide a maximum compensation of around 20% of the transmission line impedance
base value.
Figure 3-16(a) show the variation of the equivalent impedance as a function of the
duty ratio, and Figure 3-16(b) show the corresponding active power at the receiving end.
Figure 3-16 Equivalent impedance (a) and power at the receiving end (b) as a function of the duty ratio.
It can be seen that the control only becomes effective for values greater than
d=30% This is due to the quadratic nature of control action, as can be observed in
77
Equation (3-9).
3.3.5
Design Considerations
As may be evident from Equation 3-9, the maximum degree of compensation is
achieved when d=100%. Therefore, this operating point determines the rating of the
various devices.
The design process starts with deciding the maximum degree of compensation for
the transmission line. This allows computation of the capacitance as,
C = 1 / (ωP ⋅ %Compensation ⋅ XL)
(3-10)
where XL is the line impedance. Since for maximum compensation the capacitor appears
directly in series with the transmission line, it has to handle the transmission line
maximum current. Therefore, the voltage and kVAs can be respectively determined as,
VC=IL⋅ XC
(3-11)
SC=VC⋅IL
(3-12)
where IL is the maximum line current.
Finally, since the converter as well as the
SIT
transformer is cascaded with the
capacitor, they have to handle the same kVAs as the capacitor.
In addition, it is well known that capacitive series compensation introduces a
78
resonance between inductive reactance of the transmission line and the compensation
capacitor. The resonant frequency can be determined as
FR = d ⋅ FP ⋅
XC
XL
(3-13)
where FP is the power frequency of the transmission system. As may be evident from the
last equation, the resonance frequency varies between zero and a maximum values as the
duty ratio is varied between zero and unity. Thus, if there exist system modes at
resonance frequencies of sustained operation, the dynamic effect should be assessed to
prevent Subsynchronous Resonances (SSR).
3.3.6
Simulation Results
A detailed computer simulation model of the system shown in Figure 3-13 was
developed to verify the operation of the Ξ−controller at various operating conditions. The
switching frequency to operate the converter was selected to be FS=2.4kHz. The base
values and per unit values of the parameters used in the computer simulation example are
shown in Appendix 2.
Figure 3-17 shows, over a 60Hz period, the perunitized waveforms of the voltage
injected into the transmission line as well as the line current for a typical operating point.
As expected, the current leads the fundamental component of the voltage by exactly 90°
so the Ξ−controller is providing reactive power only.
79
Figure 3-17 Phase−A Current and Injected Voltage shown over a 60Hz period.
Although the voltage waveform appears to be rich in harmonics, it only represents
a small percentage of the transmission line voltage, therefore the actual phase−to−ground
as well as the phase−to−phase voltages are mostly free of harmonics.
As an illustration of the open loop dynamic performance of the Ξ−controller,
consider the transition from operating at d=20% to d=80%. Figure 3-18 and Figure 3-19
show the response of the line current and the response of the measured power at the
receiving end, when the duty ratio is stepped up or ramped up, respectively.
80
100
50
0
2
0
-2
1.2
1.1
1
0.9
0.9
0.95
1
1.05
1.1
1.15
1.2
1.25
1.3
Figure 3-18 Simulation results for a step−change of the duty ratio (a) from 20 to 80%. (b) and (c) show the
response of the line current and the power at the receiving end, respectively.
Figure 3-19 Simulation results for a ramp−change of the duty ratio (a) from 20 to 80%. (b) and (c) show the
response of the line current and the power at the receiving end, respectively.
81
It can be seen that in both cases the system is able to stabilize although in the step
case a significant overshoot is observed, suggesting that for open loop operation it would
more appropriate to slowly tune the duty ratio. Closed loop operation will be studied in
Chapter 5.
3.3.7
Summary
This section has introduced the Ξ−controller: an application of the vector
switching converter acting as a series compensator. A simplified illustration of the
system was used to demonstrate the principles of power flow control using duty ratio
modulation.
It
was
demonstrated that,
as
other semiconductor−based series
compensators, the Ξ−controller can continuously adjust the transmission line impedance
and therefore realize active power control. The results from a computer simulation of the
system including the dynamics of the switching process were presented to validate the
approach. A brief discussion of design considerations indicates the trade−offs in the
design of converter parameters.
A quantitative comparison of Ξ−controller against its DC link counterpart (SSSC) is
studied in Chapter 6.
Γ−
−Controller: An AC Link Unified Power Flow
3.4
Controller
The
DC
link Unified Power Flow Controller (UPFC) has been presented as the
82
ultimate solution for power flow control in transmission network. This is due to its
unique capabilities of simultaneously controlling the bus voltage and both the active and
reactive power down the line [17]. However, it has been recognized, and also realized in
practice, that the main operating mode is power flow control [17]. Bus voltage can
usually be locally controlled at the substation. It has also been agued that by controlling
the three parameters at the same time the system becomes unreliable upon the possible
loss of the UPFC. Thus, controlling two parameters at a time seems to be the most suitable
mode of operation.
The AC link counterpart of UPFC is introduced in this section. The connection into
the system, shown in Figure 3-20, is similar to the DC link UPFC but the topology as well
as the control function is realized in a different manner.
Figure 3-20 Schematic of a VeSC−based AC link Unified Power Flow Controller.
This device can also simultaneously control the active and reactive power through
a transmission line, therefore it can be thought of as a UPFC.
3.4.1
Topology
Figure 3-21 shows the detailed three−phase realization of the
AC
link
UPFC.
The
83
central feature of the approach is to control the power flow along the transmission line by
injecting an
AC
voltage with controllable magnitude and phase angle in series with the
line. The system is configured by including a shunt phase−shifting transformer (SPT),
filter capacitors (TFC), a quadruple−throw single−pole three−phase
VeSC,
and a series
injection transformer (SIT). This system may be located at any point throughout the line.
Figure 3-21 Topology of a VeSC−based AC link Unified Power Flow Controller.
The
SPT
transformer (with four secondary winding sets) has the double function
of lowering the impressed voltage to match the capability of the semiconductor switches
and simultaneously phase−shifting the impressed voltage by 0o, 90o, 180o and 270°
respectively, to obtain four sets of three−phase voltage vectors. This clearly resembles
84
the situation sketched in Figure 2-3(d), with (VT1−VT4) representing the throws. The
SPT
may be implemented as a five winding per leg three−phase transformer as indicated in
Figure 3-21, where all the set of windings that are drawn parallel are magnetically
coupled together and physically located in the same leg.
The capacitor bank
TFC
is used to absorb the high frequency currents introduced
due to switching from permeating into the system across the transformer.
The VeSC uses a quadruple−throw single−pole three−phase switch to synthesize a
controllable voltage at the pole terminals by modulating the duty ratios d1, d2, d3, d4.
Invoking Figure 2-3(d) again, given the throw voltages generated by the SPT, the range of
values the pole voltage can take is 0 ≤ VP ≤ V̂P for the magnitude, and 0 ≤ ∠VP ≤ 360 for
the phase angle, both which can be independently controlled.
Finally, the
SIT
transformer accommodates the voltage to an appropriate level to
be injected back in series with the transmission line and also provides the necessary
electrical isolation.
It is important to note that the stiff voltage and stiff current requirements assumed
in the previous section are given by the capacitor bank and
SIT
transformer leakage
inductance, respectively.
The
SPT
AC
link
UPFC
system may also be realized by using a four−winding per leg
transformer instead. In this case, the throws should be generated as shown in Figure
2-3(c), which is physically obtained by the transformer illustrated in Figure 2-7. This
realization only needs triple throw−single pole switches and therefore is simpler and
easier to control; however, as is evident from Figure 2-3(c), it needs higher throw
85
voltages to synthesize the same V̂P .
3.4.2
Equivalent Circuit
By following the procedure explained in Chapter 2, the single−phase averaged
equivalent circuit shown in Figure 3-22 can readily be obtained. As can be seen, the
controlled source approach has been chosen to represent the converter. Notice as well
that the systems S and R, the transmission line impedances, and the secondary−side
impedance of the transformer (the one in series with the transmission line) have all been
aggregated into the Thévenin equivalent circuit.
In addition, the turn ratios of transformer
STP
have been made complex numbers
in order to reflect the phase−shift introduced. Although resistive elements are not shown
in the figure, they may be easily added to represent the power losses throughout the
various stages.
This equivalent circuit may be conveniently utilized to derive and study the
application of VeSC−based
AC
link
UPFC
power flow control by incorporating them in
power system load flow and dynamic analysis programs.
86
Figure 3-22 Single−Phase Averaged Equivalent Circuit.
3.4.3
Principle of Operation
In order to understand the principle of operation, consider the same situation as in
Sections 3.2.3 and 3.3.3, repeated here for convenience. A purely inductive transmission
line is connecting the sending end system S and the receiving end system R, as indicated
in Figure 3-23(a). Also, assume that the voltages VS and VR have equal magnitude, say
V, and they are displaced by a transmission angle of, say θ°. Thus, the active and reactive
87
power flows through the transmission line are, respectively
P=
V2
sin θ
X
Q S = −Q R =
(3-14)
V2
(1 − cos θ)
X
(3-15)
Consider now the insertion of the
AC
link
UPFC
at the sending end of the line. In
order to limit the discussion to the principle of operation, assume that transformers
and
SIT
SPT
are ideal and have a unity transformation ratio. Furthermore, assume that the
reactive power injection at the fundamental frequency due to the bank
TFC
is negligible
so it may be removed.
Under these assumptions, the single−phase equivalent circuit for the system as
shown in Figure 3-23(b) can be readily obtained following the standard procedure
described in Chapter 2.
(a) Systems S and R connected via a purely
inductive transmission line
(b) System S and R connected via a transmission line plus
an AC link Unified Power Flow Controller.
Figure 3-23 The ideal AC link Unified Power Flow Controller.
88
The transformation ratio of the SPT transformer becomes, as in the case of the
Π−controller, a complex quantity that may be represented as
D = (d 1 − d 3 ) + j (d 2 − d 4 ) = d 13 + j d 24
reflecting the simultaneous phase−shifting introduced by the
(3-16)
SPT.
Furthermore, in order
to satisfy the continuity of switched variables of the VeSC, the duty ratio components
must satisfy
d1 + d 2 + d 3 + d 4 = 1
(3-17)
The power flow equations now take the form of,
P=
V2
[sin θ − d 13 sin θ − d 24 cos θ]
X
QS =
V2
2
2
[ 1 − cos θ + d13 + d13 cos θ − 2d13 + d 24 − d 24 sin θ]
X
V2
[cos θ − 1 − d 13 cos θ + d 24 sin θ]
QR =
X
(3-18)
(3-19)
(3-20)
Equations (3-18) through (3-20), along with (3-16)−(3-17) dictate the control
capabilities of the ideal
AC
link unified power flow controller. It can be seen that, by
choosing appropriate values for d13 and d24, it is possible to control either (P,QS) or
(P,QR), that is, it can independently control the active power and reactive power at either
end of the transmission line. In addition, for a given (P,Q) operating point, it can keep
89
such a point unchanged as the angle θ varies i.e., it has phase−shifting capabilities, or as
the voltage V varies i.e., it also has tap−changer capabilities. Altogether, these
capabilities are comparable to that of the DC link UPFC.
Since the embedded transformer shown in Figure 3-23(b) has a leading shunt leg
followed by a series leg coupled to it, the schematic has the appearance of the uppercase
Greek letter gamma, Γ, so this particular interconnection may be termed Γ−controller.
3.4.4
Control Capabilities
A 110kV/100MVA/60Hz case study system was developed to explore the
performance of the Γ−controller further. In reference to Figure 3-21, consider the system
S (modeled as a stiff voltage source of VS = 1∠0 pu) connected to the system R (modeled
as a stiff voltage source of VR = 1∠ − 20 pu) through a transmission line of
Z L = 0.24∠85 pu. Thus, without the Γ−controller embedded, the active and reactive
power at the receiving end would be PR=1.4pu and QR=−0.36pu, respectively. This will
be considered as the base case for the following discussion.
Now consider the insertion of the Γ−controller splitting the transmission line into
two impedances ( Z S = 0.07∠82 pu and Z R = 0.17∠86 pu). The turn ratios of the
SPT
and SIT transformers was selected to be 10 : 2 and 1:1, respectively. Typical values were
selected for the parasitic elements of each transformer. Finally, the capacitor bank was
selected to be XC=3pu. Refer to the Design Considerations Section for further details on
90
the selection of these values and to the Appendix 3 for the complete list of parameters.
Next, based on the operating point described above, various control strategies are
discussed. The attention will be focused on the receiving end control variables, namely,
PR and QR. The discussion that follows will be based on the d13−d14 plane, with the
variables of interest implicitly displayed. This is in contrast with the Control Capabilities
Section for the Π−controller, where similar information was generated but with the
controlled variables explicitly displayed.
3.4.4.1 Transmitted Power Control
Figure 3-24 shows the operating plane of d13 and d14. Similar to Figure 3-5, the
unity rhombus represents the operating boundaries in d13 and d24 that are realizable given
the restrictions for each duty ratio. The three lines that are drawn inside the unity
rhombus representing the loci of d13 and d24 that must be set in order to independently
control a particular quantity as labeled. These lines take into account operating
constraints of the design posed by the turn ratios of the transformers, size of the
capacitor, voltage ratings of the different elements, etc.
91
θ = 20 Q max
= 0.17
R
PRmin = 1.01
PRmin = 0.95, Q max
= −0.25
R
Q min
R = −0.81
PRmax = 1.83
PRmax = 1.90, Q min
R = −0.50
Figure 3-24 Receiving−end power control.
Thus, the line labeled as “Active Power” shows that the Γ−controller can vary PR
in the range [1.01, 1.83] while keeping the rest of the parameters from the base case
unchanged. Similarly, the line labeled as “Reactive Power” shows that the Γ−controller
can vary QR in the range [-0.81, 0.17] while keeping the rest of the parameters from the
base case unchanged. Furthermore, the line labeled “Both”, shows that it is possible to
simultaneously control both the active and reactive power in such a way that the power
factor (pf) at the receiving end remains constant.
3.4.4.2 Tap−
−changing/phase−
−shifting control
Often during system operation, it may be desirable to deliver constant active and
92
reactive power at the receiving end, while the system operating conditions (VS, VR or θ)
may vary. Figure 3-25 shows the values of d13 and d24 that must be set in order to deliver
constant active and reactive power while either VS, VR or θ are changing within the
specified range.
VRmax
= 1.12
θ = 20 θ max = 27 VSmax = 1.12
VSmin = 0.88
VRmin = 0.91
θ min = 14
Figure 3-25 Tap−changing, phase−shifting control.
The line labeled as “Sending End Voltage” shows that the Γ−controller can
maintain base operating conditions while VS varies in the range [0.88, 1.12]. Similarly,
the line labeled as “Receiving End Voltage” shows that the Γ−controller can maintain
base operating conditions while VR varies in the range [0.91, 1.12]. Furthermore, the line
labeled “Transmission Angle,” shows that it is possible to simultaneously hold the active
and reactive power received steady as the transmission angle θ varies in the range [14o,
27o].
93
3.4.5
Design Considerations
The major circuit parameters that need to be selected are the transformer
turn−ratios, filter capacitors, and switching frequency.
3.4.5.1 Transformers turn ratio
The main criteria driving the selection for the transformer turn ratios are the
desired maximum voltage to be injected into the line and the voltage level at which the
switches can safely operate. The former can be expressed as V̂P = a P a S VL / 2 , where aP,
aS and VL are the
SPT
and
SIT
transformer turn ratios and the transmission line rated
voltage, respectively. For the latter, the restriction is that aPVL<VB, where VB is the
voltage−withstand capability of the switch. For the particular case studied here,
a P = 2 / 10 and a S = 1 are selected.
3.4.5.2 Filter Capacitors / Switching Frequency
As a general rule, switching power converters need filter capacitors across the
throw terminals and filter inductors in series with the pole terminals. In the Γ−controller,
the pole terminals appear in series with the injection transformer SIT and the transmission
line. These elements provide enough series inductance naturally, eliminating the need for
separate filter reactors. However, at the throw terminals, filter capacitors are necessary to
isolate the high frequency currents introduced due to the switching process from
propagation into the SPT transformer and further into the system.
These capacitors also introduce resonances into the system (three different in this
94
case) so the capacitance needs to be carefully selected in order to obtain satisfactory
performance.
Figure 3-26 shows the variations of these resonance frequencies ( FR1 , FR2 and FR3 )
as a function of the capacitance. For each value considered for the capacitance, the
resonance frequency for all possible values of the various duty ratios is computed. Thus,
each entry in the scatter graph actually consists of a cluster of points. The resonance
frequency FR1 (= 1 / 2 π L PS C ) is independent of the duty ratios. FR2 has some dependence
on the duty ratios but is largely dominated by the term 1 / 2 π L PS C , as may be evident
from the figure. However, FR3 results from the interaction between the capacitor and the
rest of the system and is more sensitive to the duty ratios, as shown in the expanded plot
in Figure 3-26, for a particular capacitance value.
5000
4500
FR1
4000
FR2
FR3
3500
375
370
3000
365
360
2500
355
350
2000
345
340
109
110
1500
FR3
1000
500
0
0
50
100
150
200
250
Figure 3-26 Resonance frequency as a function of the capacitor value.
95
On the other hand, the switching process introduces harmonics into the system
with frequency of the form nFS ± FP (n=1,2,…∞), where FS and FP are the switching and
power frequency, respectively [33, 35]. All of these frequencies must far exceed the
resonances of the system in order to avoid excitation of such frequencies. Clearly the
critical case is when n=1. Thus, since normally FP<<FS, it is sufficient to show that FS
exceed the various resonances.
As can be seen from Figure 3-26, this issue may be readily solved by choosing a
large capacitance or a high switching frequency. However, neither of these solutions
comes at low cost. A high switching frequency implies high switching losses whereas a
large capacitance results in large physical size and reactive power circulation. Thus, the
selection of the capacitor and switching frequency involves a difficult trade−off. For the
particular case studied here, we select C=73µF and FS=2.4kHz, which provides sufficient
operating margin, as indicated in the figure.
3.4.6
Simulation Results
A detailed computer simulation model of the system shown in Figure 3-21 was
developed to verify the operation of the Γ−controller embedded in a transmission system
at various operating conditions. The complete set of parameters used in the computer
simulation is shown in Appendix 3.
The first case−study chosen for the simulation corresponds to moving from the
base case stated in Control Capabilities Section i. e., PR=1.4pu and QR=−0.36pu, to a
96
new operating point given by PR=1.5pu and QR=0. This condition is achieved by setting
the duty ratios to d1=0, d2=18%, d3=57% and d4=25%. Figure 3-27 shows, over a 60Hz
period, the voltage/current waveforms at the receiving−end perunitized with respect to
their maximum base value. It can be seen that the voltage and current waveforms are
in−phase so QR=0 and the maximum values match the set point for 1.5pu active power
(vR=1pu, iR=1.5pu).
2
1.5
1
0.5
0
-0.5
-1
-1.5
-2
0.25
0.255
0.26
0.265
0.27
0.275
0.28
Figure 3-27 Phase−A Voltage and Current at the receiving−end.
Figure 3-28 and Figure 3-29 show the internal workings of the converter. Figure
3-28 shows, over a 60Hz period, the switch SA (refer to Figure 3-21 and Figure 3-22)
throw voltages. As can be seen, they are 90° phase−shifted as predicted early. On the
97
other hand, Figure 3-29 shows the currents through the various terminal of phase−A
transformer. As can be seen,
harmonics,
although the
SPT
SPT
secondary currents (ips(1-4)) are rich in
the actual current absorbed from the transmission line (isht) is fairly
sinusoidal.
1
0
-1
1
0
-1
1
0
-1
1
0
-1
0.25
0.255
0.26
0.265
0.27
0.275
Figure 3-28 Phase−A Capacitor voltages.
0.28
98
1
0
-1
1
0
-1
1
0
-1
1
0
-1
1
0
-1
0.25
0.255
0.26
0.265
0.27
0.275
0.28
Figure 3-29 Phase−A SPT transformer currents.
A second case study was developed to evaluate the open−loop dynamic response
of the Γ−controller upon a change on the commanded duty ratios. The transition studied
is a change on an operating point given by PS=1.2pu and QS=─0.2pu to a different one
given by PS=1.5pu and QS=0pu. Two different simulations are carried out. First, the duty
ratios are directly stepped to achieve the new operating point. Second, the duty ratios are
ramped, within a 100ms interval, to smoothly move to the new operating point.
Results for the step−change case are presented in Figure 3-30 and Figure 3-31.
The former shows the commanded duty ratios whereas the latter shows the response of
the line current, the active power, and the reactive power. As can be seen, the response is
as a second−order system. Power oscillations take place, which are damped out after
approximately seven 60Hz cycles.
99
40
20
0
40
30
20
60
40
20
30
20
10
0
0.9
0.95
1
1.05
1.1
1.15
1.2
1.25
1.3
Figure 3-30 Duty Ratio Commands. Stepped from d1=33%, d2=35%, d3=32% d4=0% to d1=0%, d2=18%,
d3=57% d4=25%.
1
0
-1
1.6
1.4
1.2
1
0.1
0
-0.1
-0.2
0.9
0.95
1
1.05
1.1
1.15
1.2
1.25
1.3
Figure 3-31 Response to a duty ratio step change of the (a) line current, (b) active power and (c) reactive
power, all at the receiving end.
100
Results for the ramp−change case are presented in Figure 3-32 and Figure 3-33.
The former shows the commanded duty ratios whereas the latter shows the response of
the line current, the active power, and the reactive power. As can be seen, the transition is
free of oscillation, which makes the ramping strategy very attractive to be selected to
actually operate the Γ−controller.
40
20
0
40
30
20
60
40
20
30
20
10
0
0.9
0.95
1
1.05
1.1
1.15
1.2
1.25
1.3
Figure 3-32 Duty Ratio Commands. Ramped from d1=33%, d2=35%, d3=32% d4=0% to d1=0%, d2=18%,
d3=57% d4=25%.
101
1
0
-1
1.6
1.4
1.2
1
0.1
0
-0.1
-0.2
0.9
0.95
1
1.05
1.1
1.15
1.2
1.25
1.3
Figure 3-33 Response to a duty ratio ramp change of the (a) line current, (b) active power and (c) reactive
power, all at the receiving end.
3.4.7
Summary
This section has introduced the Γ−controller: an application of the vector
switching converter acting as a unified power flow controller. It is capable of
independently controlling the active and reactive power flow along a transmission line.
The system consists of a multiple quadrature phase−shifting transformer operated in
conjunction with multiple−throw three−phase switches, which are pulse width modulated
to synthesize an adjustable voltage that is injected in series into the line.
The capabilities of the Γ−controller under different modes of operation have been
demonstrated using a detailed transformer equivalent circuit model. The results from a
computer simulation of the system, including the dynamics of the switching process,
have been presented to validate the approach. A brief discussion of design considerations
102
indicates the trade−offs in the design of converter parameters.
The Γ−controller can be considered to be the
AC
link counterpart of the more
classical UPFC system based on DC link voltage source converters (VSC). It is based on the
well−established technology of classical phase−shifting transformers with solid−state
electronic control using pulse width modulation. While the control capabilities of both
approaches are similar as has been illustrated in this chapter, they utilize entirely different
topological structures for the realization of the function. The
utilizes well−established multipulse/multilevel
VSC
DC
link based approach
technologies with advanced high
voltage, high current power semiconductor switching devices such as the gate turn−off
thyristor with modest switching speeds. The integration of the multipulse VSC converters
to realize high quality waveforms requires the use of multiple phase−shifting
transformers. On the other hand, the
AC
link VeSC−based approach utilizes gate turn−off
switch technology as well but is operated under pulse width modulation at reasonably
high switching frequencies. Furthermore, it also utilizes multiple phase−shifting
transformers to realize the topological functionality. While it can be shown that the
throughput ratings of the power converters used in both the approaches will be identical,
given identical range of operation and control capabilities, their cost, size and complexity
involve complex trade−offs that are unclear at first glance. A more detailed comparison
of the converter design aspects and their dynamic capabilities will reveal further
trade−offs between the two approaches, which is proposed as one of the future research
paths in Chapter 9.
103
Chapter 4 Dynamic Modeling of AC
Link Power Flow Controllers
4.1
Introduction
Dynamic modeling of vector switching converters can be performed based on
traditional state space averaging techniques to predict performance at power frequencies.
Using the equivalent circuits developed in Section 2.2.2, the dynamic phasors’ state
space model can readily be obtained [8, 36-39]. While this approach provides results
acceptable for the development of load flow models, it falls short in providing
design−oriented information for the realization of these systems, particularly with respect
to the sensitivity of circuit parameters. The main reason for this is that the switching
frequencies attainable with these converters are only a small multiple of the power
frequency, due to the high levels of power throughput involved in relation to the power
capability of switching semiconductors. Therefore, the dynamic phenomena are often
overwhelmed by interactions between power frequency, switching frequency and the
natural modes inherent in the network structure. Thus, refinements to the classical state
space averaging techniques are necessary to obtain better models for the system.
Generalized averaging theory based on Fourier techniques have been developed to study
switching frequency dynamics for
DC−DC
converters [40-44]. The goal of this chapter is
104
to extend generalized averaging tools for
AC
link power flow controllers to predict
system behavior with fidelity that extends up to the switching frequency [35].
4.2
Generalized Averaging Theory
Generalized averaging theory has its roots [44] in the fact that a periodic signal
x (τ) can be represented to an arbitrary accuracy on the interval τ ∈ [ t − T, t ] using Fourier
series as
x ( τ) =
∞
∑ ⟨ x⟩
k
( t ) e jk ω τ
(4-1)
k = −∞
where ω = 2π / T is the corresponding frequency and
th
⟨ x⟩ k ( t ) is the k Fourier
th
coefficient, also referred to as the k phasor or the index {k} average, given by:
t
1
⟨ x ⟩ k ( t ) = ∫ x (τ) e − j k ω τ dτ
T t −T
(4-2)
Notice that these Fourier coefficients are function of time as the interval [ t − T, t ]
slides over the actual waveform. The aim is to determine an appropriate state space
model in which the coefficients given by Equation (4-2) are the state variables.
105
These coefficients feature the properties listed below, which are key for applying
averaging theory to describe the dynamic behavior of vector switching converters.
4.2.1
Index symmetry
When x ( t ) is a real−valued waveform, it readily follows from Equation (4-2) that
⟨ x ⟩ − k ( t ) = ⟨ x ⟩ *k ( t )
(4-3)
where ⟨⋅⟩ * stands for the complex conjugate of ⟨⋅⟩ .
4.2.2
Reconstruction
It can be shown using Equation (4-1) that the reconstruction formula takes the
form of
∞
Im
x (τ) = ⟨ x ⟩ 0 (τ) + 2∑ [⟨ x ⟩ Re
k ( τ) cos( kωτ) − ⟨ x ⟩ k ( τ) sin( kωτ)]
(4-4)
k =1
where the super indices ⟨⋅⟩ Re and ⟨⋅⟩ Im stand for real and imaginary components of ⟨⋅⟩ ,
respectively.
106
4.2.3
Derivative with respect to time
Differentiating Equation (4-1) yields the following formula to express the
derivative of the kth phasor with respect to time
d
d
⟨ x⟩ k ( t ) = − j ω k ⟨ x⟩ k (t ) + ⟨ x⟩ k ( t )
dt
dt
4.2.4
(4-5)
Average of a product
Using Equation (4-1), the index {k} average of a product can be expressed by
means of the following discrete convolution relationship
⟨ x y⟩ k ( t ) =
∞
∑ ⟨ x⟩
k −i
( t ) ⟨ y⟩ i ( t )
(4-6)
i = −∞
4.3
Dynamics for electric elements
As will become evident in the next section, the index {k} dynamics for the
various elements present in electric power systems are needed to develop the state space
models using generalized averaging theory. Next, the index {k} dynamics for resistors,
inductors and capacitors are developed.
107
4.3.1
Resistor
The equation that governs the dynamic of a resistive element is the well−known
Ohm’s Law,
v( t ) = R ⋅ i( t )
(4-7)
Averaging both sides of Equation (4-7) and since averaging is a linear operator, it
readily follows that the index {k} dynamics for a resistor is,
⟨ v⟩ k ( t ) = R ⋅ ⟨i⟩ k ( t )
4.3.2
(4-8)
Inductor
Voltage and current are related in an inductor through Faraday’s Law as,
v( t ) = L ⋅
di( t )
dt
(4-9)
Averaging the above equation and using Equation (4-5) the index {k} dynamics
for an inductor becomes,
⟨ v⟩ k ( t ) = L
d
⟨i⟩ k ( t ) + j ω kL⟨i⟩ k ( t )
dt
(4-10)
108
4.3.3
Capacitor
The terminal current−voltage relationship for a capacitive element is,
i( t ) = C ⋅
dv( t )
dt
(4-11)
With a similar argument as in the inductor case, the index {k} dynamics for a
capacitor becomes,
⟨ i⟩ k ( t ) = C
d
⟨ v⟩ k ( t ) + j ω kC⟨ v⟩ k ( t )
dt
4.4
(4-12)
Application to State Space Models for Vector
Switching Converters
Vector switching converters realize power flow control synthesizing a desirable
pole voltage obtained by switching among stiff−voltage throws. The requirements of
stiff voltage at the throw side and stiff current at the pole side are obtained by
respectively adding capacitors and inductors. The switch itself is modeled by means of
controlled sources, where the switching function q(t) takes the discrete values of {0,1} at
a switching frequency, say FS, depending on the position of the switch. Furthermore,
109
vector switching power flow controllers are embedded in power systems excited by
sinusoidal voltage sources at a power frequency, say FP. This is schematically described
in Figure 4-1.
Figure 4-1 Schematic of a generic power system with a vector switching converter embedded.
The development of the state space model shows that the differential equation that
governs the system has the form of
d
x ( t ) = A (q ( t )) x ( t ) + Bu ( t )
dt
(4-13)
where the state vector, x(t), corresponds to a collection of capacitor voltages and inductor
currents, and the input vector, u(t), corresponds to a collection of excitation voltages.
Notice that the Linear Time Variant (LTV) system described by Equation (4-13) has no
approximations whatsoever i.e., q(t) represents the true switching function and the
excitation voltages embedded in u(t) are the sinusoidal waveforms. Due to the discrete
nature of q(t), the exact determination of the states from Equation (4-13) in an extremely
difficult problem.
The classical approach has been to, under proper assumptions, directly replace the
110
switching function q(t) by the duty ratio d(t). This makes the system described by
Equation (4-13) continuous and therefore solvable. The key underlying assumption for
this is that the switching frequency should exceed both the power frequency and the
natural modes of the physical system. Under these conditions, the ripple becomes
negligible and the response of the states is largely dominated by the interaction between
the
DC
component of the switching function (i.e., the duty ratio) and the excitation
frequency quantities. Indeed, the response becomes independent of the switching
frequency.
However, in the case of vector switching converters, this assumption may not
always be satisfied. Switching at high frequency may not be possible due to the inherent
switching losses. Thus, the need for a more refined method becomes evident.
Generalized averaging provides the framework in which the dynamic modeling
for vector switching power flow controllers can be achieved at an arbitrary accuracy.
Thus, the analysis, component sizing, and control design can be performed with a greater
degree of accuracy.
To apply the generalized averaging scheme to a system described by Equation (413), it is necessary that the system have a unique identifiable period. In the state equation
presented above, however, there are two different periods: the power period TP associated
with ve(t) and the switching period TS associated with q(t). Therefore, averaging would
not be applicable if TP and TS were arbitrarily chosen. However, that limitation can be
overcome if the analysis is restricted to the special case when the ratio power
period/switching period is an integer number, say N, because then the switching function
111
is also TP−periodic.
Averaging (4-13), the index {k} differential equation takes the form of
⟨
d
x ⟩ k ( t ) = ⟨ A (q ( t )) x ( t ) + Bu ( t )⟩ k
dt
(4-14)
Using the fact that averaging is a linear operator and also using (4-5), the equation
above can be further reduced to
d
⟨ x ⟩ k ( t ) = − j ω k ⟨ x ⟩ k ( t ) + A ' ⟨ q ( t ) x ( t )⟩ k + B⟨ u ( t )⟩ k
dt
(4-15)
where A’ can readily be deduced from A(q(t)). The simplifying task now becomes the
representation of ⟨ q ( t ) x ( t )⟩ k in Equation (4-15) to a summation of various terms at
distinct frequencies using the convolution property (4-6). Therefore, it is first necessary
to develop expressions for q(t) and x(t) in terms of their corresponding phasors.
As stated earlier, q(t) is a TS−periodic function so in terms of the power period
(TP=N⋅TS) it can be expressed as
q(t ) =
∞
∑ ⟨ q⟩
Nk
(t ) e
j Nk ω p t
k = −∞
Expanding this series,
(4-16)
112
q ( t ) = ⟨ q⟩ 0 + ⟨ q ⟩ N e
jN ωp t
+ ⟨ q⟩ − N e
− jN ω P t
+
∞
∑ ⟨ q⟩
Nk
e
j Nk ω p t
(4-17)
k = −∞
k ≠ 0 ,1, −1
Notice that the time dependency for the phasor quantities has been dropped to
avoid notational clutter. Since the interest is to capture phenomena at switching
frequency, the switching function is approximated by its three first terms. Therefore,
Equation (4-17) becomes
q( t ) ≈ ⟨ q⟩ 0 + ⟨ q⟩ N e
j N ωp t
Conceptually
+ ⟨ q⟩ − N e
this
j N ωp t
corresponds
(4-18)
to
augmenting
the
switching
function
representation by including the fundamental component in addition to the traditional
representation of the duty ratio only, as suggested in Figure 4-2, shown for the particular
case of d=50%. This allows capture of the high frequency (ripple) dynamics in the state
space model.
113
Figure 4-2 Waveforms corresponding to the switching function, index {0} approximation (duty
ratio) and index {0, -N, N} approximation. All with d=50%.
On the other hand, it is known that x(t) is a TP−periodic function but its
harmonics content is not known a−priori. Therefore, it will be assumed that the
harmonics are all present, i.e.,
x(t) =
∞
∑< x >
k
e
jk ωp t
(4-19)
k = −∞
Now it is possible to apply Equation (4-6) to obtain the various index averages
that result from the product ⟨ q ( t ) x ( t )⟩ k ,
114
⟨q x ⟩1 = ⟨q⟩ 0 ⟨ x ⟩1 + ⟨ q⟩ − N ⟨ x ⟩ ( N +1) + ⟨q⟩ N ⟨ x ⟩ − ( N −1)
(4-20)
⟨ qx ⟩ N −1 = ⟨ q⟩ 0 ⟨ x ⟩ N −1 + ⟨ q⟩ N ⟨ x ⟩ −1
(4-21)
⟨ q x ⟩ N +1 = ⟨ q ⟩ 0 ⟨ x ⟩ N +1 + ⟨ q ⟩ N ⟨ x ⟩ 1
(4-22)
⟨ q x ⟩ −1 = ⟨ q⟩ 0 ⟨ x ⟩ −1 + ⟨ q⟩ − N ⟨ x ⟩ N −1 + ⟨ q⟩ N ⟨ x ⟩ − ( N +1)
(4-23)
⟨q x ⟩ − ( N −1) = ⟨ q⟩ 0 ⟨ x ⟩ −( N −1) + ⟨ q⟩ − N ⟨ x ⟩ 1
(4-24)
⟨ qx ⟩ − ( N +1) = ⟨ q⟩ 0 ⟨ x ⟩ − ( N +1) + ⟨ q⟩ − N ⟨ x ⟩ −1
(4-25)
⟨ qx ⟩ g = ⟨ q⟩ 0 ⟨ x ⟩ g
(4-26)
g ≠ ±1, ±(N-1), ± (N+1)
In order to exploit certain properties, the various indices presented above are
grouped into (a) positive group: index {1, (N-1), (N+1)}; (b) negative group: index {-1, (N-1), -(N+1)}; and (c) zero group: index {g}.
The above equations can be notably simplified by examining the nature of the
excitation functions. Although the excitation frequency is always ωP (usually 2π⋅60Hz),
the actual excitation index can take two different forms, as discussed next.
4.4.1
Real Excitation
When the three−phase system is balanced and no phase shifts are introduced by
transformers (resulting from a ∆−Y connection, for example), the equivalent circuit may
be represented by only one of the three−phases (usually phase A). The other two phases
carry no extra information as they are just a phase−shifted repetition of phase A. Hence,
115
for a three−phase system given by
VA ( t ) = 2V sin(ω P t + θ)
(4-28)
VB ( t ) = 2V sin(ω P t + θ − 120)
(4-29)
VC ( t ) = 2V sin(ω P t + θ + 120)
(4-30)
the excitation in the single−phase equivalent circuit takes the form of,
u ( t ) = 2 V sin(ω P t + θ)
(4-31)
In terms of the average index, u(t) can be expressed as
∞
u (t ) =
∑ ⟨ u⟩
k
(t ) e
jk ωp t
= ⟨ u⟩ 1 e
j ωp t
+ ⟨ u ⟩ −1 e
− j ωp t
(4-32)
k = −∞
where,
⟨ u⟩ 1 =
⟨ u ⟩ −1 =
V
2
(sin θ − j ⋅ cos θ)
V
2
(sin θ + j ⋅ cos θ)
(4-33)
(4-34)
116
Notice that since u(t) is a real−valued waveform, ⟨ u ⟩ −1 = ⟨ u ⟩ 1* . Thus, considering
the expression for the index products and the series expansions for the excitation
waveforms, the averaged dynamics given by Equation (4-15) becomes
d
⟨ x ⟩ 1 = − j ω P ⟨ x ⟩ 1 + A '{⟨ q⟩ 0 ⟨ x ⟩ 1 + ⟨ q⟩ − N ⟨ x ⟩ ( N +1) + ⟨ q⟩ N ⟨ x ⟩ − ( N −1) } + B⟨ u ⟩ 1
dt
(4-35)
d
⟨ x ⟩ N −1 = − j ω P ( N − 1)⟨ x ⟩ N −1 + A '{⟨ q⟩ 0 ⟨ x ⟩ N −1 + ⟨ q⟩ N ⟨ x ⟩ −1 }
dt
(4-36)
d
⟨ x ⟩ N +1 = − j ω P ( N + 1)⟨ x ⟩ N +1 + A '{⟨ q⟩ 0 ⟨ x ⟩ N +1 + ⟨ q⟩ N ⟨ x ⟩ 1 }
dt
(4-37)
d
⟨ x ⟩ −1 = j ω P ⟨ x ⟩ −1 + A '{⟨q⟩ 0 ⟨ x ⟩ −1 + ⟨ q⟩ − N ⟨ x ⟩ N −1 + ⟨ q⟩ N ⟨ x ⟩ − ( N +1) } + B⟨ u ⟩ −1
dt
(4-38)
d
⟨ x ⟩ −( N −1) = j ω P ( N − 1)⟨ x ⟩ − ( N −1) + A '{⟨ q⟩ 0 ⟨ x ⟩ − ( N −1) + ⟨ q⟩ − N ⟨ x ⟩ 1 }
dt
(4-39)
d
⟨ x ⟩ −( N +1) = j ω P ( N + 1)⟨ x ⟩ −( N +1) + A '{⟨ q⟩ 0 ⟨ x ⟩ −( N +1) + ⟨q⟩ − N ⟨ x ⟩ −1 }
dt
(4-40)
d
⟨ x ⟩ g ( t ) = − j ω P g ⟨ x ⟩ g + A '{⟨ q⟩ 0 ⟨ x ⟩ g }
dt
(4-41)
g ≠ ±1, ±(N-1), ± (N+1)
The examination of Equations (4-35) through (4-41) brings to the discussion two
important conclusions:
(i)
Since the excitation is at index {-1, 1} only, the state variables from the zero
group have no forcing function and therefore do not exist whatsoever, assuming
no initial condition at those indices. Therefore, Equation (4-41) can be eliminated,
117
as it bears no information.
(ii)
Since the input signals are real−valued waveforms, positive−index variables are
the complex conjugate of the corresponding negative−index variable, therefore it
is possible to solve the positive group or the negative group independently of each
other. Moreover, each of them gives a closed−form solution for the state variables
and therefore can be used to predict the dynamic behavior of the entire system.
Therefore, the initially infinite number of averaged differential equations
described by (4-15) has been reduced to three differential equations, namely Equations
(4-35) through (4-37). Notice the final number of variables is (two) x (three) x (number
of
original
states).
Each
original
state
is
augmented
by
three
( x ( t ) → ⟨ x ⟩ 1 ( t ), ⟨ x ⟩ N −1 ( t ), ⟨ x ⟩ N +1 ( t ) ) and by two after separating in real and imaginary
parts each complex index variable. These equations, as mention before, described the
dynamics of a phase−shift free balanced power systems with vector switching power
flow controllers embedded.
4.4.2
Complex Excitation
In the case of a power system with phase−shifting elements, it is necessary to
generalize the notion of the single−phase equivalent circuit introduced in the previous
section. For example, as discussed in Section 2.3, for the case of multiple secondary
transformers, the desired 90° phase−shift on the secondary side voltage VA2 was obtained
118
by designating as phase A the corresponding primary side voltage VBC. The notion of
complex state vectors [38] fits perfectly in this framework as complex vectors carry the
information concerning the three phases simultaneously. Complex vectors are defined as
x V (t) =
2
( x A ( t ) + γx B ( t ) + γ ∗ x C ( t ))
3
where γ = e
j
2π
3
(4-42)
. The real and imaginary parts of the space vector are equal to the α and β
components when the three−phase quantities are transformed to the Clarke’s αβ
stationary reference frame [45]. For balanced systems, there is a simple inverse
transformation,
x A ( t ) = Re( x V ( t ))
x B ( t ) = Re( γ * x V ( t ))
x C ( t ) = Re( γx V ( t ))
(4-43)
Applying this transformation to the set of balanced voltages described by Equations (428)
through (4-30), the state vector for the excitation voltages becomes
u ( t ) = 2V (sin(ω P t + θ) − j ⋅ cos(ω P t + θ))
In terms of the average index, u(t) can be expressed as
(4-44)
119
u (t ) =
∞
∑ ⟨ u⟩
k
(t ) e
jk ωp t
= ⟨ u⟩ 1 e
j ωp t
(4-45)
k = −∞
where,
⟨ u ⟩ 1 = 2V (sin θ − j ⋅ cos θ)
(4-46)
Notice that in contrast with the previous case, now the only excitation index is {1}. Thus,
considering the expression for the index products and the series expansions for the
excitation waveforms, the averaged dynamics given by Equation (4-15) becomes
d
⟨ x ⟩ 1 = − j ω P ⟨ x ⟩ 1 + A '{⟨ q⟩ 0 ⟨ x ⟩ 1 + ⟨ q⟩ − N ⟨ x ⟩ ( N +1) + ⟨ q⟩ N ⟨ x ⟩ − ( N −1) } + B⟨ u ⟩ 1
dt
(4-47)
d
⟨ x ⟩ N −1 = − j ω P ( N − 1)⟨ x ⟩ N −1 + A '{⟨ q⟩ 0 ⟨ x ⟩ N −1 + ⟨ q⟩ N ⟨ x ⟩ −1 }
dt
(4-48)
d
⟨ x ⟩ N +1 = − j ω P ( N + 1)⟨ x ⟩ N +1 + A '{⟨ q⟩ 0 ⟨ x ⟩ N +1 + ⟨ q⟩ N ⟨ x ⟩ 1 }
dt
(4-49)
d
⟨ x ⟩ −1 = j ω P ⟨ x ⟩ −1 + A '{⟨q⟩ 0 ⟨ x ⟩ −1 + ⟨ q⟩ − N ⟨ x ⟩ N −1 + ⟨ q⟩ N ⟨ x ⟩ − ( N +1) }
dt
(4-50)
d
⟨ x ⟩ −( N −1) = j ω P ( N − 1)⟨ x ⟩ − ( N −1) + A '{⟨ q⟩ 0 ⟨ x ⟩ − ( N −1) + ⟨ q⟩ − N ⟨ x ⟩ 1 }
dt
(4-51)
d
⟨ x ⟩ −( N +1) = j ω P ( N + 1)⟨ x ⟩ −( N +1) + A '{⟨ q⟩ 0 ⟨ x ⟩ −( N +1) + ⟨q⟩ − N ⟨ x ⟩ −1 }
dt
(4-52)
d
⟨ x ⟩ g ( t ) = − j ω P g ⟨ x ⟩ g + A '{⟨ q⟩ 0 ⟨ x ⟩ g }
dt
(4-53)
g ≠ ±1, ±(N-1), ± (N+1)
120
As in the case with real excitation, the examination of Equations (4-47) through
(4-53) brings to the discussion two important conclusions:
(i)
Since the excitation is at index {1} only, the state variables from the zero group
have no forcing function and therefore do not exist whatsoever, assuming no
initial condition at those indices. Therefore, Equation (4-53) can be eliminated, as
it bears no information.
(ii)
Since the only equation that exhibits excitation is (4-47), the indices present on
that equation {1, N+1, -(N-1)} represent the corresponding state variables that are
different than zero. Thus, Equations (4-48), (4-50) and (4-52) may also be
eliminated, as they have no forcing function on the RHS.
Therefore, the initially infinite number of averaged differential equations
described by (4-15) has been reduced to three differential equations, namely Equations
(4-47), (4-49) and (4-51). As in the case of the real excitation, the final count for
variables is (two) x (three) x (Number of original states). These equations described the
dynamics of a balanced power systems with vector switching power flow controllers
embedded.
Since this is a generalization of the case with real excitation, it should be noted
that the vector space approach may also be used in the case of a power system fully
Y−connected. The final time−domain waveforms should be identical.
The augmented schemes developed above have the advantage of incorporating the
121
ripple into the model and therefore could be used for design purposes in the sense that the
predicted
RMS
quantities give a better approximation of the actual stress that the various
devices will have to tolerate.
Also, closed−loop regulators that are switching frequency dependent can be
designed and therefore capture phenomena that are out of reach with the traditional
approach of approximating the switching function by the duty ratio only.
Finally, note that when approximating the switching function by the duty ratio
only, the traditional state space in terms of dynamic phasors is recovered. This will
become evident in the following examples. The number of variables in this case is (two)
x (number of original states).
4.5
Case Studies
In this section the techniques of generalized averaging are applied to the modeling
of the Ξ−controller and the Γ−controller. In the Ξ−controller case, the advantages
towards the design process are investigated, whereas in the Γ−controller case the
attention is focused on the dynamic performance.
4.5.1
Generalized Averaging Applied to the Ξ−controller
4.5.1.1 The Model
In order to develop the generalized averaged model, consider the equivalent
122
circuit for Ξ−controller presented in Figure 3-14, but now the converter is modeled
directly using the switching function rather that the duty ratio. As usual, the switching
function q(t) takes values {0,1} according to a switching frequency FS (or ωS) and a duty
ratio d. This is shown in Figure 4-3.
Figure 4-3 Exact model for the Ξ−controller.
Let the sending end voltage and receiving end voltage be represented as
v S ( t ) = 2V sin(ω P t ) and v R ( t ) = 2V sin(ω P t + θ) , respectively. Considering i(t) and
v(t) as state variables, the instantaneous state equation takes the form of
 R
d  i( t )   − L
=
dt  v( t )  q ( t )
 C
−
q(t ) 
1
L   i( t )  +   v ( t )

 L SR
0   v( t )  0 

(4-54)
123
where R = R S + R R , L = L S + L R + L T and v SR ( t ) = v S ( t ) − v R ( t ) .
Since this is a phase−shifting free system, the rules developed for real excitation
apply. Hence, the generalized averaged state space model becomes
⟨ q⟩ 0
⟨q⟩ *N
⟨ q⟩ N
d
R
1
⟨ i⟩ 1 = ( − j ω p − ) ⟨ i⟩ 1 −
⟨ v⟩ 1 −
⟨ v⟩ N +1 −
⟨ v⟩ *N −1 + ⟨ v SR ⟩ 1
dt
L
L
L
L
L
(4-55)
⟨q⟩ N
⟨ q⟩ 0
d
R
⟨i⟩ N −1 = (− j ω p ( N − 1) − )⟨i⟩ N −1 −
⟨ v⟩ N −1 −
⟨ v⟩ 1*
dt
L
L
L
(4-56)
⟨ q⟩ N
⟨ q⟩ 0
d
R
⟨i⟩ N +1 = (− j ω p ( N + 1) − )⟨i⟩ N +1 −
⟨ v⟩ N +1 −
⟨ v⟩ 1
dt
L
L
L
(4-57)
⟨ q⟩ 0
⟨ q⟩ N
⟨ q⟩ *N
d
⟨ v⟩ 1 = − j ω p ⟨ v⟩ 1 +
⟨ i⟩ 1 +
⟨i⟩ N +1 +
⟨i⟩ *N −1
dt
C
C
C
(4-58)
⟨ q⟩ N
⟨ q⟩ 0
d
⟨ v⟩ N −1 = − j ω p ( N − 1)⟨ v⟩ N −1 +
⟨i⟩ N −1 +
⟨i⟩ 1*
dt
C
C
(4-59)
⟨ q⟩ N
⟨ q⟩ 0
d
⟨ i⟩ 1
⟨ v⟩ N +1 = − j ω p ( N + 1) ⟨ v⟩ N +1 +
⟨i⟩ N +1 +
dt
C
C
(4-60)
where < v SR >1 = −(V / 2 )(sin(θ) + j (cos(θ) − 1)) , ⟨ q⟩ 0 = d , and
⟨ q⟩ N = (1 / π)(sin(dπ) cos(dπ) − j sin 2 (dπ))
As can be seen, the state space model described by Equations (4-55) through (459) has six complex differential equations and captures the interaction between the
fundamental and high frequency quantities, as there is an explicit dependency on the
124
switching frequency (ωS=NωP).
On the other hand, approximating q(t) by the duty ratio only ( q ( t ) ≈ ⟨ q⟩ 0 ( t ) ), the
traditional state space model in term of the fundamental quantities is recovered,
R

d  ⟨ i⟩ 1   − j ω p − L
=
dt ⟨ v⟩ 1   ⟨ q⟩ 0
C

⟨ q⟩ 0 
1
L   ⟨ i⟩ 1  +   ⟨ v ⟩

 L SR 1
− j ω p  ⟨ v⟩ 1   0 

−
(4-61)
The above state space has two complex differential equations and there is no
switching frequency dependency. Notice that this index {1} model differs from the exact
model only in two aspects: (i) the switching function has been replaced by the duty ratio
and (ii) the diagonal elements of the state matrix have been augmented by the term –jωP.
4.5.1.2 Model Accuracy
The generalized averaged model can bring insights to the design process. For
example, say the interest is in calculating the steady state capacitor stress, specifically the
RMS
current. In order to investigate the accuracy of each model, a comparison of the
prediction made by the traditional model and the generalized averaged scheme against
the true switched waveforms can be made. Figure 4-4 shows the results obtained. The
switched waveform is obtained from a detailed computer simulation. On the other hand,
the index {1} and {1, N-1, N+1} approximations are obtained by first setting to zero the
125
left hand side (LHS) of their corresponding state space representation. This gives the
capacitor voltages so by using Equation (4-12) the corresponding phasor current can
readily be obtained. Finally, the reconstruction formula (4-4) allows determination of the
time−domain waveforms.
1.5
Switched
1
Index {1, N-1, N+1}
Index {1}
0.5
0
-0.5
-1
-1.5
1.7
1.705
1.71
1.715
1.72
1.725
1.73
Time (s)
Figure 4-4 Waveform of the switched capacitor current along with the traditional and augmented
approximations. FP=60(Hz), FS=1020(Hz), d=50%.
As can be seen, the prediction made by the augmented model is much better than
the traditional model. The RMS values for this particular case are ISW=722(A), I1=510(A)
and IN=685(A) for the switched, index {1} and index {1, N-1, N+1} waveforms,
respectively. Although the augmented model still falls short, it gives a more realistic
insight about the current stress the capacitor has to manage.
126
4.5.1.3 Eigenvalues Pattern
It is well known that capacitive series compensation introduces a resonance
between the inductive reactance of the line and the compensation capacitor. However, the
frequency at which this resonance occurs is usually on the order of a few tens of hertz.
On the other hand, the switching frequency is at a few thousands of hertz, therefore a
dynamic interaction between the switching frequency and the modes of the system should
not be expected.
The pattern followed by the eigenvalues is of special interest as it directly relates
to the physical resonance of the system as well as to the power and switching frequency.
The system exhibits a physical resonance of approximately FR=30(Hz) which, neglecting
the various resistances, can be computed as FR = (1 / 2 π)(d / LC ) . When the system is
modeled by the index {1} approximation, two extra oscillatory modes are added into the
state matrix whose imaginary parts are given by ±FR±FP (±30, ±90(Hz)). When the
system is augmented to include the fundamental component of the switching frequency
eight extra modes are incorporating, yielding to a total of 12 modes whose imaginary
parts are ±FR±FP±FS (±30, ±90, ±930, ±990, ±1050, ±1110(Hz)). This is summarized in
Figure 4-5.
127
Figure 4-5 Eigenvalues distribution for the different models. FR=30(Hz), FP=60(Hz), FS=1020(Hz).
This observation illustrates a simple litmus test that may be employed to identify
the need for modeling higher index dynamics. By examining the system dynamic
resonant frequencies, power frequency and switching frequency, it is possible to
determine the degree of mathematical complexity needed to obtain a faithful
representation.
4.5.2
Generalized Averaging Applied to the Γ−controller
4.5.2.1 The Model
Similarly to the Ξ−controller case, the exact dynamic model for the Γ−controller
128
can be obtained from its corresponding equivalent circuit (Figure 3-22) using q(t) instead
of d(t). This is shown in Figure 4-6.
Figure 4-6 Exact model for the Γ−controller.
In this case, vS and vR represent the corresponding space vectors,
v S ( t ) = 2V(sin(ω P t ) − j ⋅ cos(ω P t ))
(4-62)
v R ( t ) = 2V (sin(ω P t + θ) − j ⋅ cos(ω P t + θ))
(4-63)
129
Considering as state variables the voltages and currents shown in the equivalent
circuit above, the exact state space model can be constructed as follows,
d
x ( t ) = A (q ( t )) ⋅ x ( t ) + B ⋅ u ( t )
dt
(4-64)
where the matrices A and B, as well as the state and input vectors are described below.
 z4
 -z
z5
1

 - jz 1

 z1
A (q ( t )) =  jz 1

 q 1a s / C
q a / C
 2 s
q 3 a s / C
q a / C
 4 s
z3
- R ps /L ps
- jz 3
- z3
jz 3
- (q 1 z 6 - z 8 )
- (q 2 z 6 + jz 8 )
- (q 3 z 6 + z 8 )
- jz 5
- z5
jz 5
q 1 z 7 + z 2 - 1/L ps
q 2 z 7 - jz 2
(q 3 z 7 - z 2 )
jz 5
z 5 - R ps /L ps
- jz 5
- z5
j(q 1 z 7 + z 2 )
j(q 2 z 7 − jz 2 ) - 1/L ps
j(q 3 z 7 - z 2 )
- z5
jz 5
z 5 - R ps /L ps
- jz 5
- (q 1 z 7 + z 2 )
- (q 2 z 7 − jz 2 )
- (q 3 z 7 − z 2 ) - 1/L ps
- jz 5
1/ C
- z5
0
jz 5
0
z 5 - R ps /L ps
0
- j(q 1 z 7 + z 2 )
0
- j(q 2 z 7 − jz 2 )
0
- j(q 3 z 7 - z 2 )
0
0
1/ C
0
0
0
0
0
0
0
1/ C
0
0
0
0
0
0
0
1/ C
0
0
0
- (q 4 z 6 - jz 8 )




j(q 4 z 7 + jz 2 )

- (q 4 z 7 + jz 2 )

- j(q 4 z 7 + jz 2 ) − 1 / L ps 

0


0

0


0

q 4 z 7 + jz 2
(4-65)
 z9
 z
 11
 jz 11

 − z11
B = − jz 11

 0
 0

 0
 0

z 10 
z 8 
jz 8 

− z8 
− jz 8 

0 
0 

0 
0 
(4-66)
 i R (t ) 
i (t) 
 PS1 
i PS 2 ( t ) 


i PS3 ( t ) 
x ( t ) = i PS 4 ( t ) 


 v C1 ( t ) 
 v ( t )
 C2 
 v C3 (t ) 
 v ( t )
 C4 
(4-67)
 v (t ) 
u (t ) =  S 
 v R ( t )
(4-68)
The constants zi (i=1,2,..,11) are functions of the parameters of the equivalent
circuit. See Appendix 4 for a complete description of them. Also, notice that the explicit
130
time dependence of qi (i=1,2,…,4) in the state matrix has been dropped to avoid
notational clutter.
Since in this case the excitation is complex, the rules developed in Section 4.4.2
have to be applied in order to develop the generalized averaged model. However, the
resulting model is a (two) x (three) x (nine) = 54 state space system. Therefore, it will not
be shown here. It can easily be calculated via symbolic computation using standard
software such as MatlabTM, MapleTM or MathematicaTM.
As in the case of the Ξ−controller, the traditional index {1} state space model can
be computed by adding -jωP onto the diagonal elements of the state matrix (4-65) and
replacing the switching functions by the corresponding duty ratios. The resulting model
in this case would be a (two) x (nine) = 18 state space system.
4.5.2.2 Model Accuracy
The Γ−controller exhibits a resonance between the phase−shifter transformer
secondary−side leakage inductor and the filter capacitor ( FR = 1 / 2π L PS C ), which might
be in the order of the switching frequency and therefore lead to unstable operation.
Although in the Design Considerations Section for the Γ−controller (Section 3.2.5) the
way in which to approach this issue qualitatively was discussed, it will now be formally
discussed in terms of the generalized averaging theory.
It has been shown that the switching process introduces harmonics into the system
with frequencies of form nFS ± FP (n=1,2,…∞). These frequencies need to far exceed the
131
resonances of the system in order to avoid their excitation. Clearly, the critical case is
when n=1 ( FS ± FP ), since (FS ± FP ) > FR ⇒ (nFS ± FP ) > FR . The case of n=1 is fully
captured by the index {1, N-1, N+1} model and therefore it can describe the actual
system at a much higher degree of accuracy as discussed in the example below.
As an illustration, the parameters that define the resonance frequency are selected
as Lps= 0.3mH and C=73µF so the resonance occurs at 1862(Hz). These values are the
same ones used to show steady state operation in the Control Capabilities Section
(Section 3.4.4), so they may be considered reasonable.
Detailed switched simulations were conducted considering FS=2400(Hz) and
FS=1800(Hz), whose results are shown in Figure 4-7 and Figure 4-8, respectively.
1
0
-1
(a)
Switched,
Index {1}
and
Index {1, N+1, -(N-1)}
1
0
-1
(b)
1
0
-1
(c)
1
0
-1
(d)
0.25
0.255
0.26
0.265
0.27
0.275
0.28
Time (sec)
Figure 4-7 Capacitor voltage waveforms with FR=1862(Hz) and FS=2400(Hz).
132
1
0
-1
1
0
-1
1
0
-1
1
0
-1
0.25
0.255
0.26
0.265
0.27
0.275
0.28
Figure 4-8 Capacitor voltage waveforms with FR=1862(Hz) and FS=1800(Hz).
The state variables plotted are the capacitor voltages vCi (i=1,2,…4) from the
equivalent circuit shown in Figure 4-6. The simulated switched waveforms are plotted
against the index {1} and index {1, N+1, -(N-1)} approximations, which are obtained
from their corresponding state space models.
As can be seen in Figure 4-7, when the resonance frequency and switching
frequency are sufficiently different, the index {1} model describes the system accurately
and the augmented scheme gives no extra information. However, when the resonance
frequency becomes closer to the switching frequency, Figure 4-8 shows that the duty
ratio based model fails to predict the oscillatory behavior of the state variables. The
augmented scheme does capture such behavior and hence its potential becomes evident.
Since these devices have been developed to control power flows along
transmission lines, it is likely that a closed−loop regulator would be used to maintain
133
appropriate set−points. In such cases, the ripple phenomena would not be captured
whatsoever by measuring net P and/or Q at 60Hz. Therefore, the duty ratio−based
approximation would be blind to the potential instabilities that may be incipient in the
system.
4.6
Summary
This Chapter has introduced the dynamic modeling for
AC
link power flow
controllers under the generalized averaging framework. The presence of excitation at
power frequencies that interact with the switching frequency requires a systematic
integration of the techniques developed for multifrequency averaging for
converters and phasor dynamic modeling common for study of
AC
DC−DC
converters. The
mathematical development of the integration is presented under the condition that the
switching frequency be an integer multiple of the power frequency.
The application of the technique to predict system behavior was illustrated using
the Ξ−controller and the Γ−controller. The results indicate that the augmented
multifrequency averaged model is superior to the traditional
DC
averaged model in
predicting high frequency interactions. The nature of the augmentation of the eigenspace
of the system under
DC
averaged and multifrequency averaged models was also
illustrated. In addition, it was briefly discussed how the augmented state matrix can lead
to identifying potential system resonances, especially when the system exhibits
oscillatory modes in the vicinity of the switching frequency.
134
Chapter 5
5.1
Since
Feedback Control for AC
Link Power Flow Controllers
Introduction
AC
link vector switching power flow controllers have been developed to
control power flows along transmission lines, it is likely that a closed−loop regulator
would be needed to maintain appropriate set−points.
Figure 5-1 shows, in the Laplace domain [46, 47], the basic structure for a
feedback control system. The plant, g(s), represents the power circuit whose outputs, y,
need to be maintained at a certain level dictated by yREF. The approach is to measure the
outputs via a transducer (β) and compare such measurements against the references to
generate an error signal (ε). This error is then processed by an appropriate compensator
h(s) to generate the manipulated inputs uM, which drive the controllable devices of the
plant. In addition, the plant usually has other inputs represented by u.
u
yREF
Controller
h(s)
uM
Plant
g(s)
Figure 5-1 Configuration for classical control design.
y
135
The control design problem is to determine h(s) in order to meet the functional
requirements of the overall system. These requirements include, among others, stability,
maintenance of the desired operating condition in case of disturbances, and acceptable
dynamic performance (overshoot, settling time).
As will become evident throughout this chapter, the Ξ−controller, although it
corresponds to a nonlinear system, is a single−input, single−output (SISO) system so the
control design problem for a nominal operating condition can be assessed by classical
techniques such as Bode plots [46, 47]. On the other hand, the Γ−controller, in addition
to being nonlinear, is a multiple−input, multiple−output (MIMO) system; therefore,
techniques that are more refined are required for the analysis [48].
5.2
Feedback Control for the Ξ−controller
The control design problem starts with the identification of the dynamic model for
the power circuit. The model chosen for this purpose is based on the single−phase index
{1} averaged equivalent circuit shown in Figure 3-14. The mathematical description was
deduced in Section 4.5.1.1. It was found that the corresponding state space is described
by Equation (4-61), repeated here for convenience,
136
R

d  ⟨ I⟩ 1  − j ω p − L
=
⋅
d
dt ⟨ V⟩ 1  
C

d 
1
L  ⋅  ⟨ I⟩ 1  +   ⋅ [⟨ V ⟩ ]
 
SR 1
 L
− j ω p  ⟨ V⟩ 1   0 

−
(5-1)
As the interest is in controlling the active power at the receiving end, the output of
the system can be defined as,
PR = {⟨ VR ⟩ 1 ⟨ I⟩ 1* }Re
(5-2)
Before proceeding further, notice the following: (i) Equations (5-1) and (5-2)
presented above are both in terms of complex quantities, so separation into real and
imaginary parts is necessary; (ii) Since there is a single (manipulated) input, d, and a
single output, PR, Equations (5-1) and (5-2) describe a
SISO
system; (iii) Due to the
inherent nonlinearities introduced by the products between the manipulated input, d, and
the state variables, ⟨ I⟩ 1 and ⟨ V⟩ 1 , linearization is needed in order to apply classical
control techniques.
5.2.1
Proposed Controller
Traditionally, the first attempt in controlling a process has been to linearize the
plant around a nominal operating condition and try a combination of a proportional (P),
integral (I), and/or derivative (D) controller.
PID
controllers are widely used in industry
137
and have become the standard for control system design. Only when the
PID
approach
fails are more sophisticated techniques considered.
In the case of processes involving switching power converters, differentiation is
particularly undesirable. Switching power converters are rich in high frequency noise
generated by the switching process. Derivative controllers have a gain that increases with
frequency and hence tends to amplify high frequency noise. Thus, as in most processes
involving switching power converters, a
PI
controller is proposed here. A formal
justification will be given in Section 5.2.1.2.
The overall closed loop system is shown in Figure 5-2. The
PI
controller acts
upon the error to generate the duty ratio as,
t
d ( t ) = K P ε ( t ) + K I ∫ ε ( τ ) dτ
(5-3)
0
vS(t)
PREF
PI
Controller
d
PWM
Pulse
Generator
Pulses
vR(t)
Power System
with a
-Controller
PR
Carrier
Figure 5-2 Block diagram of the control system.
Since the model for the power circuit is the index {1} average, the
PWM
pulse generator
138
may be considered to have a unity transfer function so the duty ratio, d, acts directly upon
the plant. In addition, since the power circuit is nonlinear, the plant may be modeled by
its small signal transfer function.
5.2.1.1 Small Signal Model
The small signal model for the plant can be obtained by linearizing Equations (51) and (5-2) around a nominal operating point. For that purpose, consider the general
form of Equations (5-1) and (5-2),
xɺ ( t ) = f ( x ( t ), u ( t ), u M ( t ))
(5-4)
y( t ) = g( x ( t ), u ( t ), u M ( t ))
(5-5)
Since the interest is in the transfer function with respect to the manipulated inputs,
the linearization of f(x(t),u(t),uM(t)) and g(x(t),u(t),uM(t)) around x 0 becomes
f ( x ( t ), u ( t ), u M ( t )) = f ( x 0 , u 0 , u M 0 ) +
∂f
∂f
(x(t) − x 0 ) +
(u M ( t ) − u M 0 ) + h.o.t. (5-6)
∂x 0
∂u M 0
g ( x ( t ), u ( t ), u M ( t )) = g ( x 0 , u 0 , u M 0 ) +
∂g
∂g
(x(t) − x 0 ) +
(u M ( t ) − u M 0 ) + h.o.t. (5-7)
∂x 0
∂u M 0
where h.o.t. stands for higher order terms. The operating point x0 can readily be obtained
139
by assigning some value to the duty ratio, say d0, and then solving Equation (5-4) for
steady state i. e., making the LHS equal to zero.
Applying Laplace’s transform to Equations (5-4) through (5-7), the small signal
model becomes
~
~
s ⋅ ∆x = A ⋅ ∆x + B ⋅ ∆u M
(5-8)
~
~
∆y = C ⋅ ∆x + D ⋅ ∆u M
(5-9)
where,
~ ∂f
A=
∂x
0
∂f
~
, B=
∂u M
0
~ ∂g
, C=
∂x
0
∂g
~
, D=
∂u M
, and ∆x = ( x − x 0 ) , ∆y = ( y − y 0 ) ,
0
∆u M = (u M − u M 0 ) .
For the particular case of the Ξ−controller, each term of Equations (5-8) and (5-9)
becomes,
[
x = ⟨ I⟩ 1Re
⟨ I⟩ 1Im
[
Im
⟨ I⟩ 10
y = PR
(5-12)
Re
x 0 = ⟨ I⟩ 10
⟨ V⟩ 1Re
Re
⟨ V⟩ 10
⟨ V⟩ 1Im
]
Im
⟨ V⟩ 10
y 0 = PR 0
T
]
(5-10)
T
(5-13)
ωp
− d0 / L
0 
− R L
 −ω
−R L
0
− d 0 / L 
~
p
A=
d0 / C
0
0
ωp 


d0 / C
− ωp
0 
 0
[
~
C = ⟨ VR ⟩ 1Re
⟨ VR ⟩ 1Im
0 0
]
(5-11)
uM = d
(5-14)
u M0 = d 0
(5-16)
− ⟨ VSR ⟩ 1Re

− ⟨ VSR ⟩ 1Im
~
B=
Re
 ⟨ I⟩ 10
C

Im
 ⟨ I⟩ 10 C
(5-18)
~
D=0
L

L



(5-15)
(5-17)
(5-19)
140
Notice that the separation in real and imaginary parts for the states variables has
taken place. The plant’s input/output transfer function, g(s), readily follows from
Equations (5-8) and (5-9),
g (s) =
∆PR
~ ~ ~
= (sI − A ) −1 B + C
∆d
On the other hand, the
(5-20)
PI
controller’s input/output transfer function, h(s), can be
obtained applying Laplace’s transform to Equation (5-3),
h (s) =
K
∆d
= (K P + I )
s
∆ε
(5-21)
The overall small signal model developed in this section is shown in Figure 5-3.
The control design problem can now be visualized as calculating appropriate values for
the gains KP and KI in order to satisfy system performance.
PREF
KP +
KI
s
d
~ ~ ~
(sI − A) −1 B + C
Figure 5-3 Block diagram for the small signal model.
Notice that the global input/output transfer function is
PR
141
∆PR
1 T (s)
= ⋅
∆PREF β 1 + T (s)
(5-22)
where T(s) is the loop gain defined as
T (s) = β ⋅ h (s) ⋅ g (s)
(5-23)
5.2.1.2 Controller Design via Bode Plots
The controller design is usually done by “shaping” the frequency response of the loop
gain, T(s), as it carries most of the information of interest [46, 47]. The Bode Plots
approach focuses on three performance indices:
(i)
Stability margins: As may be evident from Equation (5-22), the situation
T( jω) = 1∠180 o should be avoided in order to maintain stability in the
bounded−input, bounded−output (BIBO) sense. In practical situations, control
designers impose security margins to keep the system away from such a
critical value. The quantitative definition of these security margins begins by
introducing the gain crossover frequency (GCF) and the phase inversion
frequency (PIF). The former refers to the frequency at which the magnitude of
T(jω) equals one ( | T ( jω) |= 1 ) while the latter refers to the frequency at which
the angle of T(jω) becomes 180° ( ∠T ( jω) = 180 o ). Thus, at the
GCF,
the
142
corresponding angle should be away from 180° by an acceptable phase
margin (φM). In addition, at the
PIF,
the corresponding magnitude should be
away from 1 by an acceptable gain margin (GM). Furthermore, it should be
evident that in order to have a stable system, the
PIF (GCF
GCF
has to occur before the
< PIF). This provides a positive φM at GCF and a positive GM at PIF. A
GM of 2 and a φM of 60° are usually considered acceptable, and are used as
targets in the control design process.
(ii)
Bandwidth: Equation (5-22) shows that the feedback is effective only for
large values of T(s). The GCF is also called the bandwidth of the regulator as it
defines the value at which the feedback becomes ineffective. The bandwidth
also features the property of being proportional to the speed of response of the
regulator. For both of these reasons, it is important to keep the bandwidth as
high as possible.
(iii)
Steady state error: The steady state error can be directly evaluated from T(0).
It should be evident from Equation (5-22) that it T(0)=∞ drives the steady
state error to zero.
These indices sometimes conflict with each other and the designer has to evaluate the
tradeoffs involved in the design process.
Based on the data shown in Appendix 2, the regulator for the Ξ−controller is designed
as follows. The transducer β is considered as a pure gain free of noise. The value of
β=10−6 is selected as it measures power in the order of MW. The frequency response of
the partial loop gain βg(s) is shown in Figure 5-4. It can be seen that PIF<GFC which is a
143
sufficient condition for instability, so the first and most important task is to stabilize the
system with an appropriate regulator.
50
0
-50
-100
0
-100
-200
-300
-400
-2
10
-1
10
0
10
1
10
2
10
3
10
4
10
Figure 5-4 Bode Plot for the partial loop gain β g(j ω).
Notice that as a result of modeling the system with the index {1} average, two
resonant peaks are observed in the frequency response. These peaks occur at FR±FP, as
predicted early by analyzing the eigenvalues’ pattern.
Figure 5-5 shows the frequency response for the proposed
PI
regulator, which is
obtained by selecting the gains as KP=10-4 and KI=0.5. Cascading this regulator as shown
in Figure 5-3, the frequency response of the final loop gain, T(s)=βh(s)g(s), becomes as
shown in Figure 5-6.
144
50
0
-50
-100
0
-100
-200
-300
-400
-2
10
-1
10
0
10
1
10
2
10
3
10
4
10
Figure 5-5 Bode Plot for the PI controller h(j ω).
50
0
-50
-100
0
-100
-200
-300
-400
-2
10
-1
10
0
10
1
10
2
10
3
10
4
10
Figure 5-6 Gain and Phase margins from the loop gain T(j ω)= β h(j ω)g(j ω).
It can be seen that now
GFC<PIF
and the Nyquist’s criterion for stability [46]
would show that the system is indeed stable. It can also be seen that the stability security
margins are within the specified range (GM=7.5dB=2.4, φM=89°) and the integral part of
145
the regulator guarantees to have T(0)=∞. The bandwidth, however, seems low but
adequate for the application, as will become evident in the next section.
5.2.2
Simulation Results
Once the regulator has been designed, computer simulations are carried out in
order to verify the regulator performance. The typical tests evaluated are the response of
the system upon a step change in the reference and upon small disturbances on the
non−manipulated inputs.
5.2.2.1 Response to a step change in the reference
This test is carried out to verify the
BIBO
stability of the system as well as to
evaluate the dynamic performance of the regulator. Figure 5-7 shows the results for a
15% step change in the reference power. As can be seen, a first order stable response is
obtained. The regulator is also able to track the reference and the speed of response is
about 0.5s, which indicates that the bandwidth is adequate.
146
1.2
1.1
1
0.9
2
0
-2
1.2
1.1
1
0.9
0.9
1
1.1
1.2
1.3
1.4
1.5
Figure 5-7 Response to a 15% step change in the reference power. (a) shows the step change applied to the
reference power while (b) and (c) show the response on the line current and the measured power at the
receiving end, respectively.
In addition, Figure 5-7 may be directly compared to the open loop response
shown in Figure 3-18, as the step change in the reference power was made coincident to
the duty ratio step change (notice that the steady state measured power in both figures is
the same). It can be seen that the closed loop regulator dramatically improves both the
overshoot and the oscillatory behavior observed in the open loop case.
5.2.2.2 Response to disturbances
This test is carried out to evaluate the ability of the system to track the reference
upon small disturbances on the non−manipulated inputs. Figure 5-8 shows the response
of the system for a step change of 5% in the magnitude of the sending end voltage. Figure
5-9 shows the response of the system to a step change of 3° in the transmission angle.
147
1.05
1
0.95
0.9
2
0
-2
1.1
1
0.9
0.8
0.7
0.9
1
1.1
1.2
1.3
1.4
1.5
Figure 5-8 Response to a 5% step change in the sending end voltage. (a) shows the step change applied to
the magnitude of the sending end voltage while (b) and (c) show the response on the line current and the
measured power at the receiving end, respectively.
24
22
(a)
20
2
(b)
0
-2
1.1
1
(c)
0.9
0.8
0.7
0.9
1
1.1
1.2
1.3
1.4
1.5
Time(s)
Figure 5-9 Response to a 3° step change in the transmission angle. (a) shows the step change applied to the
transmission angle while (b) and (c) show the response on the line current and the measured power at the
receiving end, respectively.
It can be seen that the regulator is able in both cases to track the reference,
148
although the oscillatory response of the power might not be acceptable. Suitable changes
in the regulator design, including a different control approach, may be needed in order to
improve these responses. This is discussed in the Future Work Chapter.
5.3
Feedback Control for the Γ−
−controller
The overall control system for the Γ−controller is shown in Figure 5-10. Since
this device was realized to simultaneously control the active and reactive power through
a transmission line, it naturally has two outputs. References to track the corresponding
outputs are likely to be used as shown in the figure. It becomes evident that, under these
circumstances, this is the case of a
MIMO
system. Classical techniques such as the
Nyquist’s stability criterion or feedback loop design using Bode Plots are not easily
extendable.
vS(t)
PREF
QREF
P
Q
?
Controller
d1
d2
d3
d4
PWM
Pulse
Generator
Pulses
vR(t)
Power System
with a
-Controller
Carrier
P
Q
Figure 5-10 Control system for the Γ−controller.
PR
QR
149
Therefore, as suggested in the figure, at first glance it is not clear what regulator
to use in order to satisfy the various performance indices introduced in the previous
section. In addition, notice that the regulator needs to generate three outputs (one of the
duty ratios can be expressed as 1−summation of the other three) from two input errors. It
has also been observed that there is cross−coupling between the outputs i. e., a change in
say εP would also affect the output value of QR.
On the other hand, as discussed in Section 4.5.2.2, it is not clear whether the
index {1} average model is precise enough to describe the behavior of the Γ−controller,
as it exhibits a resonance that might be in the same order the switching frequency. The
augmentation to the indices {1, N+1, -(N-1)} average might be needed for satisfactory
closed loop performance.
All of these issues make the Γ−controller regulator design a challenging problem
that is under investigation is proposed as future research in Chapter 9.
5.4
Summary
This chapter has introduced the notion of closed loop regulation for AC link power
flow controllers. In the case of the Ξ−controller, the dynamic modeling for the power
circuit was made by using the index {1} average model introduced in Chapter 4.
Classical control techniques indicated that a PI regulator would perform satisfactorily. On
the other hand, in the case of the Γ−controller, a brief introduction to the control design
problem was made, indicating that the subject is left as an open question for further
investigations.
150
Chapter 6
A comparative
evaluation of Power Flow
Controllers using AC Link and DC
Link Converters
6.1
Introduction
When introducing a new technology it is customary to contrast its features against
the state−of−the−art. As pointed out in Chapter 1, power flow controllers that use a
DC
link voltage source inverter are considered to be superior to previous realizations in terms
of harmonic performance, dynamic response and ease of operation.
Although the functional equivalency between the
AC
link and
DC
link approaches
has been illustrated in the literature [14, 27], a detailed comparison between their features
was not attempted. It is the goal of this chapter to present a comparative evaluation of the
two approaches.
A comparative evaluation of particular solutions for a given application may be
made on the basis of several performance features. Salient features include: harmonics of
waveforms, operating losses, ratings of power converters, reactive component
requirements, transformer kVA requirements, complexity of control, dynamic response in
terms of improving voltage stability, small signal and large signal stability, damping of
151
sub−synchronous resonance, etc. Given the degree of variability based on the application
of these devices in the form of shunt, series, shunt−series and interline connections, a
definitive evaluation appears to be a formidable task. Therefore, in order to maintain a
focus in the evaluation, a particular application, namely series compensation is
considered in this chapter. Furthermore, the evaluation is limited to solutions that feature
superior waveform quality arising from high frequency or high pulse number switching
with sinusoidal line current waveforms. This precludes the consideration of
and
GCSC
TCSC, TSSC
[14] solutions from evaluation in this study. It is felt that due to their
fundamentally different harmonic characteristics due to phase controlled switching, the
benchmarking would not be consistent, and unreasonably extend the scope of the chapter.
In this chapter, a focused analytical modeling and design study of a candidate
application using the
DC
link and the
AC
link power converters is performed in order to
evaluate their performance. The comparison criteria used for the evaluation include
voltage, current and power throughput ratings of the main power circuit components
(including transformers, capacitors, and semiconductors), quality of terminal voltage and
current waveforms in terms of harmonics, losses in power semiconductors and transient
performance of the solutions in tracking commands and rejecting disturbances. No
attempt is made herein to estimate the comparative cost of these solutions, as such costs
are a complex function of market trends, economic factors and engineering development,
and would change considerably with respect to time and location. However, the design
ratings for transformers, capacitors and power semiconductors together may be
considered to be a proxy for the capital costs of equipment.
152
6.2
Candidate Compensators
6.2.1
DC
The
link series compensator, termed
manners.
DC
DC
Link Series Compensator
capacitors or
DC
SSSC,
can be realized in a variety of
inductors could be used to realize the
DC
link, leading to
the use of a voltage source inverter or a current sourced inverter, respectively. Moreover,
multi−pulse or
PWM
can be selected as a modulation strategy to synthesize the output
voltage. The multi−pulse (48−pulse) voltage source inverter option is chosen here as it is
considered the state−of−the−art approach and it is commercially available with ratings up
to 100(MVA) [19].
Figure 6-1(a) shows a simplified schematic of the main stages involved in the
realization of the 48−pulse system [49]. As can be seen, the DC link is realized using two
series connected
DC
capacitors with access to their mid−point
N
[50-52]. This allows
three voltage levels (VC1, VC2, 0). Four independent diode−clamped three−phase
three−level converters are connected to the
DC
bus. Figure 6-2(a) shows the detailed
switch realizations for each converter. The voltage synthesis for phase A is shown in Fig.
3(a). Phase−shifted waveforms (±120o) are synthesized for phases B and C. As suggested
in Figure 6-1(a), each three−level converter is then connected to a magnetic structure
based on Y−zigzag ∆−zigzag phase−shifting transformers that allows creation of a
24−pulse waveform. The final 48−pulse waveform is obtained by making the conduction
angle of each converter σ=172.5o. Thus, the lowest characteristic harmonic numbers
153
present in the output waveforms are 47 and 49.
It may be concluded from Figure 6-1(a) and Figure 6-2(a) that stable operating
points are achieved only when the output voltage is in quadrature (±90o) with the output
current, because otherwise there would be active power exchange between the
capacitors and the
AC
DC
system, leading to a permanent charge or discharge of the
capacitors and therefore unstable operation. The same argument is used to control the
magnitude of the AC output voltage. By allowing a small real power exchange for a short
duration between the
DC
link and the
AC
system, it is possible to drive the capacitors’
voltage to a desirable value and therefore adjust the output voltage. Finally, it should be
noted that in order to generate symmetric
AC
output voltage waveforms, a suitable
algorithm for equalizing the average value of VDC1 and VDC2 needs to be implemented
[53]. Section 6.5.2 describes the implementation of such algorithm.
Figure 6-1 Schematic of candidate compensators. (a) SSSC (b) Ξ−controller.
6.2.2
AC
Link Series Compensator
The Ξ−controller was extensively studied in Section 3.3. As the goal of this
chapter is to contrast the features of this device against the
SSSC,
some schematics and
154
concepts will be repeated here for convenience.
Figure 6-1(b) shows a realization for the Ξ−controller that naturally follows from
the
SSSC
topology introduced in the previous section. As can be seen, four banks of
three−phase
AC
capacitor make up the
link. Each bank is switched using a
AC
three−phase vector−switching converter whose principle of operation is depicted in
Figure 6-2(b) and Figure 6-3(b). The former shows the switch realization of a
double−throw single−pole three−phase switch and the latter the voltage synthesis process
for phase A. Phase−shifted waveforms (±120o) are synthesized for phases B and C.
VCA
IDC1
+
T1A
D1A
T1B
D1B
DC1A
VDC1
T2A
D2A
T1C
T2B
D2B
VCB
D1C
DC1B
DC1C
T2C
IA
D2C
T3A
D3A
T3B
D3B
DC2A
VDC2
T4A
IDC2
D4A
T3C
T4B
D4B
IA
VA
T1B
IB
VB
T1C
IC
VC
T2A
IC
DC2C
T4C
D1C
VB
D3C
DC2B
D1A
VCC D1B
VA
IB
N
T1A
VC
D4C
D2A
T2B
D2B
T2C
N
(a)
-
(b)
D2C
Figure 6-2 Converter realizations. (a) SSSC (b) Ξ−controller.
PWM Injected Voltage
Carrier
Multi-pulse Injected Voltage
T2AT3A
T1AT2A
T2AT3A
T3AT4A
T2AT3A
Duty Ratio
VDC1
(a)
T1A
T 2A
(b)
VDC2
0
180
360
t (deg)
t (deg)
Figure 6-3 Waveforms illustrating AC voltage synthesis. (a) SSSC. (b) Ξ−controller.
As can be seen from Figure 6-3(b), the
AC
voltage generated by each capacitor
155
bank is rich in harmonics and hence unless output filters are used, the resulting waveform
would be unacceptable for utility applications. However, by using the magnetic structure
suggested in Figure 6-1(b) and phase−shifting the PWM carriers of each bank by 90o, it is
possible to operate with no output filter at a modest frequency of FS=720(Hz). This, as in
the SSSC case, also limits the lowest characteristic harmonic numbers to be 47 and 49. A
more detailed contextual comparison of the waveforms injected by each compensator is
discussed later in Section 6.4.2.4.3.
6.3
Test system description
Figure 6-4 shows the case study system that will be used to perform the
comparative evaluation. As can be seen, it corresponds to a simple transmission line with
a two bus power system on which a series compensator has been embedded. The
compensator can be either the SSSC or the Ξ−controller.
Figure 6-4 One line diagram of the candidate series compensation system.
In order to make the comparison meaningful, both the
SSSC
and the Ξ−controller
are assumed to provide the same amount of compensation. Table 6-1 shows the specific
values for the transmission system considered in this case study. Due to the lumped
156
resistance of the transmission line, the line power slightly varies depending on the
location where the measurement is taken. The location downstream of the converter (at
V2 in Figure 6-4) is selected to measure the power. It readily follows that, without
compensation, the nominal line power is PL{Nom}=140(MW). A 40(MW) increase is
selected for full compensation, which represents a ~30% increase above the nominal
value. Thus, at full compensation, one would measure PL{Comp}=180(MW).
Table 6-1 System Data
Base Values: 110kV/100MVA/60Hz
Bus Voltage (pu)
Impedance (pu)
VS = 1∠0
ZS=0.01+j 0.17
VR = 1∠ − 20
6.4
6.4.1
ZR=0.01+j 0.07
Steady state performance evaluation
Operating Modes
As suggested in Figure 6-4, each compensator can operate in different control
modes. Although the ultimate objective is the control of the active power through the
line, this can be achieved through different regulation strategies. The most direct
approach involves commanding a certain amount of desired power PL and synthesizes a
voltage waveform to achieve such an amount. This approach has the advantage of
maintaining the active power through the line at a pre−established value but, under some
contingences, this may not be desirable or possible [14]. Alternatively, one can command
a certain amount of injected voltage VINJ or a certain amount of injected impedance
157
ZINJ=VINJ/IL. These three control methods, namely power mode, voltage mode and
impedance mode, can be used in each approach to realize power flow control.
The remarks made in Section 6.2 readily lead to the steady state equivalent
circuits presented in Figure 6-5. As can be seen, the natural steady state representation of
the
SSSC
is a controlled voltage source while for the Ξ−controller is a controlled
capacitor. For the
SSSC,
the manipulated input is the magnitude of the
while for the Ξ−controller it is the duty ratio (d). The operation of the
the synthesis of
AC
synchronized to the
DC
voltage (VDC)
SSSC
depends on
voltage waveforms using the constituent inverters appropriately
AC
line using a phase−locked−loop (PLL). On the other hand, the
Ξ−controller is regulated through a simple duty ratio control that varies the amount of
injected
AC
capacitance with the line in the open loop mode. However, through
appropriate feedback control, both of these devices can be regulated to operate in any o
the operating modes described above.
Figure 6-5 Steady sate equivalent circuits. (a) SSSC (b) Ξ−controller.
In order to explicitly show the ability of each compensator to operate under the
various operating modes, formulas that relate the line power (PL), the injected voltage
158
(VINJ) and the injected impedance (ZINJ), with respect to the manipulated inputs are
presented in Table 6-2 [29]. These formulas readily follow from the equivalent circuits
from Fig. 5. For the sake of simplicity, it is assumed that the transmission line is purely
inductive and |VS|=|VR|=V. In addition, it is assumed that the injection transformer is
ideal (XT=0) and has unity turn ratio. Finally, it is also supposed that the line current is
instantaneously available for control purposes. In the case of the Ξ−controller the
assumption of XT=0 is removed to explicitly show that this device can provide a certain
degree of inductive compensation (when XT>d2XC). However, it has been pointed out in
the literature [16] that series compensators seldom provide steady state inductive
compensation due to the increase in reactive power circulation associated with it, so the
focus herein is on capacitive compensation only.
Table 6-2 Operating Modes
Operating Mode
Voltage
VINJ = mVDC
Impedance
ZINJ =
Power
PL =
6.4.2
Ξ−Controller
SSSC
1
XL
1 ± (2V sin θ / 2) / mVDC
V2
V
θ
sin θ ±
mVDC cos
XL
XL
2
VIJN =
2V
θ
sin
2
2
1 + X L /(X T − d X C )
Z INJ = X T − d 2 X C
PL =
V2
sin θ
X L + (X T − d 2 X C )
Design Considerations
As may be evident from Figure 6-4, the main components to be rated are the
injection transformer, the compensation capacitors and the power converter. The overall
design process begins by assuming that the bus voltages at locations VS and VR are
159
known so that the nominal operation of the transmission line without compensation can
be found. Next, defining the percentage (say λ) by which the line power is going to be
increased above its nominal value, the line power at full compensation can be computed
as PL{Comp} = (1 + λ)PL{Nom} . Finally, in reference to Figure 6-4, by solving Equations (6-1)
and (6-2), the electrical variables at the injection transformer terminals, VINJ and IINJ(=IL)
can be found.
VS − ZS ⋅ IL − | VINJ | ∠(∢IL − 90°) − ZR ⋅ IL − VR = 0
(6-1)
ℜ{3 ⋅ (VR + ZR ⋅ IL ) ⋅ I*L } = PL{Comp}
(6-2)
It should be noted that Equations (6-1) and (6-2) provide three scalar equations
that allow the unknown variables |IL|, ∢IL and |VINJ| to be solved.
The values of VINJ and IINJ are independent of the type of compensator used and
provide the basis for the design process. For the case study, λ=30% so the solution of
Equations (6-1) and (6-2) gives VINJ=6.1(kV) and IINJ=965(A). It follows that the injected
power is QINJ=3VINJIINJ=17.6(MVAr).
6.4.2.1 Choice of power semiconductors
Power semiconductor switch ratings and properties remain the limiting factor that
determines the highest possible power throughput in any application. Selection and
application of switching devices involves a careful trade−off between the various
families of power semiconductors matched against the ratings required for the particular
power converter. An examination of the topology of the converters indicates that single
quadrant force commutated semiconductors and diodes form the building block for both
160
realizations. The device families that fit these needs at voltage levels in the order of a few
kV are
GTOs, IGCTs
and
IGBTs.
family of devices while the
Among these,
IGBT
GTOs
and
IGCTs
belong to the thyristor
belongs to the transistor family of devices. While both
families of devices are capable of turning off without the need for any snubbers, due to
the regenerative mechanism that dominates the turn on process in thyristors,
IGCTs
GTOs
and
require di/dt limiting reactors to be included in their realizations. The introduction
of di/dt limiting reactor further necessitates a ‘pole clamp’ that captures the trapped
stored energy in the di/dt limiting reactors and diverts it to the
DC
bus during each
turn−off event. A detailed discussion of realization and performance of
devices for
DC
link converters along with a comparison with
GTO
IGCT
switching
devices may be found
in the literature [54]. Furthermore, series connected IGCT devices for continued converter
operation during device failures in such applications to enhance system reliability have
also been demonstrated. However, such series operation requires
RC
networks in parallel
with the devices to ensure steady state and dynamic voltage sharing. On the basis of these
evaluations,
IGCT
may be concluded to be the preferable switching device for the
SSSC
system.
In comparison to the
DC
link converter used in realizing the
SSSC
which involves
single quadrant switching during commutations, the switching mechanism in the
three−phase vector switching converters to realize the Ξ−controller involves
four−quadrant switching during commutations. For each throw, the commutation process
involves leaving a forward or reverse conducting state and entering a forward or reverse
voltage blocking state during turn−off and vice−versa during turn−on. The actual
161
quadrants of conduction and commutation sequence are in fact determined by the phase
of the three−phase excitation waveforms and also the operating duty ratio. Due to this
situation, the viability of including a di/dt limiting reactor for each throw and devising a
suitable mechanism for absorbing the trapped energy in a lossless and effective manner is
challenging. Therefore, the use of IGCTs in the realization of Ξ−controller is considered to
be demanding.
On the other hand, high power
IGBTs
are capable of operation without snubbers
for turn−on and turn−off, and are capable of operating in parallel with ease [55].
Furthermore, various laboratory prototypes of three−phase vector switching converters
have been realized using IGBTs [27, 56]. The application of a four−step switching strategy
during commutation for effective and efficient clamp−free four quadrant switching using
IGBT
realizations have been demonstrated in matrix converters, and can be readily
extended to the three−phase vector switching converters [26]. Therefore, the
IGBT
is
chosen to be the preferred power semiconductor for realizing the Ξ−controller.
The detailed engineering design of the switch realization including appropriate
snubbers to manage the commutation, thermal management and protection features for
both the
SSSC
and Ξ−controller are beyond the scope of this evaluation and is not
considered in this thesis. However, the ratings and selection of particular devices and the
losses in them will be considered further in the evaluation.
6.4.2.2 Injection Transformer
The total transformer fundamental kVA ratings are the same for both types of
162
compensators as they are meant to deliver the same amount of compensation for the line.
However, the number of transformers and turn ratio of the transformers are selected on
the basis of harmonics and voltage compatibility with the power converters.
The selection of the number of transformers (T) is related to the harmonic pulse
number that can be realized across the entire series compensator, given a switching
strategy or frequency. Recommendations of acceptable harmonics injected in power
systems from power converters are specified up to the 40th harmonic [57]. As discussed
earlier, the realizations illustrated in Figure 6-1 with T=4 transformers in series, limits the
lowest harmonic numbers to be 47 and 49, thereby meeting the recommended waveform
quality levels.
The turn ratio of each transformer (n) is selected on the basis of the maximum
voltage blocking capability of the semiconductor switches, namely IGCTs for the SSSC and
IGBTs
for the Ξ−controller. A more detailed discussion of this including the provision for
(n−1) reliability for each of the realization is presented in a subsequent sub−section
(Section 6.4.2.4.1). The chosen values for the turns ratio and different transformer
parameters are summarized in Table 6-3. It should also be noted that, in the case of the
SSCC, minor arrangements ought to be made in order to properly rate each winding
considering the zigzag and ∆ connections [58]. Although the current through each of the
transformers are substantially sinusoidal, as will be illustrated further, the excitation
voltage waveforms for both type of compensators contain a significant amount of
harmonics. Thus, the transformer design features will have to be modified to
accommodate this. First, the core area of the transformers will have to be derated to
163
account for the increased flux arising from the non−sinusoidal voltage waveforms in
order to prevent saturation. A core area derating factor DFφ for each of the compensators
may be defined as the ratio of the area of one half cycle of the excitation voltage
waveforms shown shaded in Figure 6-3 to the area of one half cycle of a sinusoidal
waveform with a corresponding amplitude of the fundamental component. Secondly, the
non−sinusoidal nature of the waveforms leads to variations in core losses in the
transformer [59, 60]. Various loss penalty factors corresponding to classical eddy current
losses (PFCE), excess eddy current losses (PFEE) and hysteresis losses (PFH) for each of
the compensator that accounts for any increase in core losses have been determined by
studying the excitation waveforms in each case. These derating factors and penalty
factors for the both the compensators are shown in Table 6-3 [60].
Table 6-3 Injection Transformer Rating
Rated Power: ST =3·(VINJ/T)·IINJ , Turn Ratio: n= VLS/VCS
Transmission Line Side (LS): VLS = √3 (VINJ/T), ILS = IINJ
Converter Side (CS): VCS = √3 (1/n) (VINJ/T), ICS = n IINJ
Ξ−controller
SSSC
ST = 4.4(MVA), n= ¼
ST = 4.4(MVA), n= 1
VLS = 2.6(kV), ILS = 965(A)
VLS = 2.6(kV), ILS = 965(A)
VCS = 10.6(kV), ICS = 241(A)
VCS = 2.6(kV), ICS = 965(A)
DFφ = (8sin σ / 2) /(σπ) = 0.86
DFφ = (2d max ) / ξ = 0.99
PFCE = 8 /(σπ) = 0.84
PFCE = (4d max ) /(ξ 2 ) = 1.1
PFEE = 16 /(7 2σ ) = 1.44
PFCE = (12 2) /(7ξ 2 ) = 1.0
PFH = 1
PFH = 1
3
The parameter σ, dmax and ξ are properties of the converters operating conditions
164
and therefore are properly defined in the next subsections. It may be observed from these
derating and penalty factors, that although the
SSSC
would require a marginally larger
core size, and larger excess eddy current losses in comparison to the Ξ−controller, it
would feature correspondingly smaller classical eddy current losses. It may be concluded
that on balance, the transformer features would be comparable for both cases.
6.4.2.3 Capacitor
As a general rule, capacitors in power converters must be selected so that the
ripple is minimized and the resonance frequency resulting from the interaction of the
capacitor and the
AC
side inductors is far away from the harmonics associated with the
switching process.
6.4.2.3.1
SSSC
Capacitor Selection
The “Unit Capacitance Constant” (UCC) was introduced in the literature [53] as a
rule of thumb for the preliminary selection of the total capacitance value of the DC link. It
is defined as
2
UCC = (1/ 2 ⋅ C ⋅ VDC
) / SCONV (Joule/MVA)
where C, VC and SCONV represent the
DC
(6-3)
link capacitance value, the
DC
voltage, and the
converter’s throughput power (MVA), respectively. It is analogous to the unit inertia
constant in synchronous rotary condensers. As in Equation (6-3) each variable represents
equivalent total values, it follows that
165
VDC = (1/ m) ⋅ (1/ n) ⋅ (VINJ / T) = 13.6(kV)
(6-4)
SCONV = 3 ⋅ VINJ ⋅ I INJ = 17.6(MVA)
(6-5)
In (6-4), m = (√2/π)·sin(σ/2) is the converter gain that relates the total
DC
voltage to the
fundamental RMS output voltage. For ideal shunt connected reactive power compensators,
the
UCC
value can be quite small – in the range of 1000 to 1500 (Joule/MVA) [53, 61].
This is because the voltage source converter does not exchange any real power with the
utility and
DC
bus capacitance requirement for only reactive power compensation is
rather small [61]. However, for voltage source converters used in an
SSSC,
any line
disturbance will transiently cause real power transfer between the converter and system.
Therefore, UCC values ranging from 3000 to 6000 (Joule/MVA) are considered typical
for voltage source converters used for series reactive power compensators [16, 62, 63].
Using Equations (6-4) and (6-5) and choosing
UCC=5250,
C=1000(µF). However, as shown if Figure 6-2(a), the
DC
the solution of (6-4) gives
link in the three level neutral
point clamped converter is actually realized using two capacitors connected in series.
Therefore, the ratings for each capacitor are,
C1 = C2 = 2C = 2000(µF)
(6-6)
VDC1 = VDC2 = VDC / 2 = 6.8(kV)
(6-7)
The
RMS
current through the
DC
capacitors can be computed analytically in terms of IINJ
considering the converter’s switching function [51] and the phase−shift introduced by the
166
various transformers,
IRMS
DC =
(6-8)
1
3 6
3 5 2 1
3 6 3 3 3 2 3
⋅ [(
+
+
+ )⋅π + (
+
+
+ ) ⋅ cos σ] ⋅ I INJ
2π
16
4
16
2
4
4
4
2
At full compensation it turns out to be I RMS
DC = 130(A) .
6.4.2.3.2
Ξ−Controller Capacitor Selection
In this case, as all variables are
AC,
the selection of the capacitance value is
directly determined by the amount of compensation required by the transmission line. For
security reasons, the operation of the
PWM
switches is usually restricted to dmin<d<dmax,
so the targeted increase in the line power should be achieved at d=dmax. Typical values
for dmin and dmax are 10% and 90%, respectively. Considering the realization shown in
Figure 6-1(b), the capacitance of each individual capacitor bank can be computed as
C=
T ⋅ (d max )2 ⋅ n 2
= 1360(µF)
(2πFP )VINJ / IINJ
(6-9)
where FP=60(Hz) is the line frequency. As the capacitor is by nature a voltage−stiff
variable, the voltage across it can be assumed to be harmonic−free and therefore the RMS
value can be computed as,
(6-10)
VC = (1/ d max ) ⋅ (1/ n) ⋅ (VINJ / T) = 1.7(kV)
On the other hand, the capacitor current is a switched variable so the
RMS
value has to
167
account for both the fundamental as well as the harmonics generated from the switching
process. From the converter’s switching functions, the
RMS
capacitor current can be
computed a
RMS
IC
= d max ⋅ n ⋅ I INJ = 915(A)
(6-11)
On the other hand, the fundamental component alone, as computed from Equations (6-9)
and (6-10) is
(6-12)
IC = VC ⋅ (2πFP ) ⋅ C = d max ⋅ n ⋅ IINJ = 869(A)
The amount of stored energy is a rough measure of the physical size of the capacitors and
may be determined from the voltage and capacitance ratings. Table 6-4 summarizes the
capacitor requirements for each compensator.
Table 6-4 Capacitor Requirements Summary
Capacitance (µF)
Voltage (kV)
Rms Current (A)
Number of Capacitor Banks
Total stored energy capacity (kJ)
Computer simulations have shown that in the
SSSC
Ξ−Controller
2000
6.8 (DC)
130
2
92.48
1360
1.7 (RMS AC)
915
12
47.16
SSSC
case and the Ξ−controller case, these
choices for the capacitance leads to acceptable voltage ripple with no resonant
interactions.
168
6.4.2.4 Converter
6.4.2.4.1
Semiconductors Ratings
High power semiconductor datasheets usually specify the maximum voltage a
device can block, the maximum average on−state current and the maximum controlled
interruptible current. Analytical formulas to express these requirements in terms of the
compensation level have been developed and are presented in Table 6-5, along with the
numerical values for the case study system.
In addition, any realization of switch using discrete power semiconductors needs
to account for continued operation during possible failures. This generally requires a
parallel and/or series connection of power semiconductors, typically with ‘n−1’
reliability. Parallel connected devices are able to ride−through an open circuit failure,
while series connected devices are able to withstand a short circuit failure, if the
remaining devices in the realization have adequate margins of safety. This leads to the
choice of series connected string of IGCTs for the SSSC and a parallel connected IGBTs for
the Ξ−controller. Since IGCTs have been readily available as press−packs and IGBTs have
been readily available as planar modules, the choice naturally fits the ride−through
criteria. Such realizations have been demonstrated and definitively evaluated in previous
investigations [54, 55].
A simplified schematic of the switch realization is shown in Figure 6-6. The
blocking voltage / conducting current capability of each of these arrangements under
normal operation as well as under ‘n−1’ contingency are summarized in Table 6-6. These
169
may be compared against the converter operating requirements listed in Table 6-5. Table
6-7 shows the device count for each compensator. As can be seen, although the device
count for the SSSC largely exceeds the Ξ−controller, total semiconductor MVA required is
only 60% higher.
Figure 6-6 Switch realization in terms of real semiconductors. For the SSSC, 4 series IGCT with antiparallel
diode (ABB 5SHX 08F4510) are used. The clamping diode is also realized using 4 diodes in series (ABB
5SDF 03D4502). For the Ξ−controller, 3 parallel IGBT with antiparallel diode (EUPEC FZ600R65KF1) are
used.
Table 6-5 Devices Stress
sssc
Ξ−Controller
Blocking Voltage VDC/2 = 6.8(kV)
√3·√2·VC=4.2(kV)
2
⋅ n ⋅ I INJ ⋅ ξ =390(A)
2π
Max. Ave.
2
σ
sin(2d max − 1)π / q
⋅ n ⋅ IINJ (cos + 1) =58(A) Notes:
ξ = 1+
On−State Current 2π
2
sin π / q
q= FSW / FP =12
Max. Switching
Current
2 ⋅ n ⋅ I INJ sin
σ
=339(A)
2
2 ⋅ n ⋅ IINJ =1360(A)
170
Table 6-6 Devices Capability
Ξ−Controller
3 Parallel IGBT−Diode
SSSC
4 Series IGCT−Diode
Normal Operation
Blocking Voltage
4·2.8 = 11.2(kV)
6.5 (kV)
Max. Ave. On−State Current
250 (A)
3·600 =1800(A)
Max. Switching Current
630 (A)
3·1200=3600(A)
n−1 Contingency
Blocking Voltage
3·2.8 = 8.4(kV)
6.5 (kV)
Max. Ave. On−State Current
250 (A)
2·600 =1200(A)
Max. Switching Current
630 (A)
2·1200 =2400(A)
Table 6-7 Device Count
Device
6.4.2.4.2
SSSC
Ξ−Controller
IGBT
−
72
IGCT
192
−
Diode
288
72
Total semiconductor MVA
561.6
336
Device Losses
Typically, semiconductor losses are categorized as conduction loss and switching
loss. The former occur during the semiconductor’s on−state while the latter take place at
the transition from on−state to off−state (or vice−versa).
The conduction loss for diodes as well as for IGBT and IGCT is usually modeled by
the product of the voltage drop across the device and the current through it. If the voltage
drop is approximated by a linear dependence on the current
171
(6-13)
VON (ION ) = V0 + rON ION
then, it readily follows that the average power loss can be determined as
2
PON = V0 ION{ave} + rON ION{rms}
(6-14)
The on−state voltage V0 and rON are usually available on the device datasheet. On the
other hand, the switching loss depends on the energy dissipated at every switching event.
The switching energy is usually assumed to be proportional to the blocking voltage and
the conducting current at the switching event
Ei =
ER
V R IR
(6-15)
Vi Ii
where, ER, VR, and IR represent the reference values for the energy loss, the blocking
voltage and the conducting current. These values are usually available on the
manufacturer’s datasheets. The subindex i represents the switching event. The power loss
over a power period is computed as
P = FP
∑ Ei
(6-16)
i
where, the subindex i sweeps all switching events that occur during the
corresponding period. In the case of
IGBT
and
IGCT
there is loss at both the turn−on and
turn−off events. While for diodes switching loss occurs only during the reserve recovery
172
that takes place at the turn−off.
Given the symmetries encountered in the voltage and current waveforms,
analytical formulas to express the losses in terms of the compensation level have been
developed and are presented in Table 6-8.
Table 6-8 Converter Losses
Type
SSSC
PC = s ⋅ 24 ⋅ n ⋅ I INJ ⋅
Conduction
Switching
 2

n ⋅ I INJ
(VCE0 + VF0 ) +
(rCE + rF ) 

π
2


3πVINJ I INJ
PS = FP
⋅
2
 E ON
E OFF
E RR 
4 R R + 4 R R + 6 R R 
VOFF IOFF
VRR I RR 
 VON ION
Ξ−Controller
PC = p ⋅12 ⋅ (n ⋅ I INJ / p) ⋅
 2

(n ⋅ I INJ / p)
(VCE0 + VF0 ) +
(rCE + rF ) 

2
 π

2π
2π 

cos d
cos(1 − d)
VINJ ⋅ IINJ 
q
q 
⋅
⋅ 3q +9
PS = FP
2π
d


s
in


q


 E ON

E
E
 R R + R OFFR + R RRR 
 VON ION VOFF IOFF VRR IRR 
Notes:
1) Subindex CE refers to IGBTs and IGCTs and subindex F refers to diodes
2) s is the number of IGCTs connected in series
3) p is the number of IGBTs connected in parallel
In order to illustrate the development of these formulas, the situation presented in
Figure 6-7 may be considered. The top plot shows the Phase−A pole current along with
the pole voltage. Depending on the throw connected to the pole is, the pole current flows
through different semiconductors. The waveforms of current carried by the different
semiconductors for the positive half cycle of the current is shown in the various plots.
The labels are consistent with Figure 6-2 and Figure 6-3. Table 6-9 summarizes the
semiconductors’ active states for an entire cycle of the current.
173
IA
VA
IA
VA
0
360
180
(a)
0
360
180
t (deg)
(b)
t (deg)
Figure 6-7 Pole current breakup. (a) SSSC, (b) Ξ−controller.
Table 6-9 Active state for semiconductors
Ξ−controller
SSSC
VDC1
VDC2
N
IA>0 T1A T2A D3A D4A DC1A T2A
IA<0 D1A D2A T3A T4A DC2A T3A
VCA
N
T1A
D1A
T2A
D2A
IA>0
IA<0
As established by Equations (6-13) through (6-16), the calculation of losses
involves computing an average or
RMS
quantity, or the identification of a switching
instant, for waveforms such as the ones presented in Figure 6-7. Once these waveforms
are described analytically, then the formulas for the losses are easily computed using the
definitions of Equations (6-13) through (6-16). As an example, Table 6-10shows the
analytical description for the current through T1A.
Table 6-10 Analytical description for IT1A
Ξ−Controller
SSSC
IT1A = ˆIA sin(ωt)
(π − σ) / 2 < ωt < (π + σ) / 2
IT1A = ˆIA sin(ωt)
i ⋅ 2π / q < ωt < (i + d) ⋅ 2π / q
i=0..q/2
174
These formulas can further be integrated and/or summed in order to obtain the
final formulae presented in Table 6-8. Table 6-11 shows the numerical values of loss
distribution for a typical operating point given by PL=160(MW). In order to achieve such
operating point, the compensators need to inject QINJ=7.7(MVAr). It can be seen that the
total converter losses are very similar for each approach, and they represent about 1% of
the converter’s throughput power. In addition, the power loss in the transmission line is
in the order of few megawatts, so the losses added by the converter represent a very small
fraction of the losses of the entire system.
Table 6-11 Converter Losses for PL=160(MW), QINJ=7.7(MVAr)
SSSC
Ξ−Controller (kW)
Conduction Diode
36.0
14.1
Conduction IGCT/IGBT
21.0
18.9
Turn−on
0.4
23.5
Turn−off
4.8
13.9
Reverse Recovery
3.7
6.3
65.9
76.7
Total
6.4.2.4.3
(kW)
Waveform Quality
Figure 6-8 shows typical injected waveforms for each compensator. The IEEE Std.
519−1992 [57] establishes that for the power level involved in this case study, the total
harmonic distortion (THD) for the bus voltage should be <2.5% and <10% for the line
current. As can be seen in Table 6-12, the injected voltage’s
THD
does not meet the
values established by the standard, but this is a differential voltage. The actual
175
line−to−ground voltages (V1 and V2 in Figure 6-4) as well as the line current are well
below the value established by the standard.
Table 6-12 Total Harmonic Distortion (THD)
Variable
SSSC (%)
Ξ−Controller (%)
VINJ
3.60
17.50
IL
0.01
0.04
V1
0.30
0.60
V2
0.10
0.20
Figure 6-8 Voltage/Current injected into the transmission line. (a) SSSC, (b) Ξ−controller.
6.5
6.5.1
Dynamic performance evaluation
Dynamic Modeling
The dynamic equivalent circuits presented in Figure 6-9 can be used to develop
the state space model for each compensator. These circuits are a generalization of the
steady state equivalent circuit presented in Figure 6-5. It should be noticed that only the
interaction between fundamental quantities is considered, so the resulting state space is a
176
dynamic phasors−based model [35, 51].
IL
+
VINJ
IL
-
IDC
+
VDC
IC
m VDC
-m IL
+
+
VINJ
-
d VC
+
VC
(a)
-
-d IL
+
(b)
-
Figure 6-9 Dynamic equivalent circuits. (a) SSSC, (b) Ξ−controller.
In the SSSC case, the converter gain is a phasor quantity defined as
m = m∠(∢I L ± 90 + α)
(6-17)
where, α is the control angle that allows charging/discharging the capacitors in order to
achieve the desired compensation. In the Ξ−controller case, the duty ratio, d, provides the
means to vary the compensation level.
6.5.2
Control Structure
Section 6.4.1 has shown that both the
SSSC
and the Ξ−controller can operate in
voltage, impedance or power control mode. The operation under power control mode has
been chosen in order to illustrate the regulator design process.
Figure 6-10 shows the control structure for each compensator under power
control mode. As can be seen, the main loop that allows the line power to be set to a
reference value is the same in both cases. In order to compute the instantaneous value of
the line power, the voltage/current at location V2 are measured, transformed into the dq
coordinates and processed through a first order filter. This allows computation of the line
power as PL=VdId+VqIq. The value of PL is then passed through a feedback gain and
177
compared to the reference power P*. The error signal generated is processed through a
regulator to finally obtain the control variable of each compensator (α, d).
Figure 6-10 Feedback control loop. TS=5 (ms), β =10-6. (a) SSSC, (b) Ξ−controller.
In the case of the
SSSC
two additional inner loops are necessary to operate the
converter. As shown in Figure 6-10(a), a phase−lock−loop is needed to assure that the
injected voltage is in quadrature with the line current. In addition, a voltage equalizing
algorithm is needed to ensure that the DC capacitors share the same voltage. Details of the
implementation of such an algorithm are presented in Figure 6-11 and discussed in the
literature [53].
Figure 6-11 Dc voltage equalizing algorithm.
178
6.5.3
Regulator Design
The procedure for the design of the feedback regulator was extensively discussed
in Chapter 5. The same approach is used here for both the SSSC and the Ξ−controller. The
transfer functions g(s)=PL/α for the
SSSC
and g(s)=PL/d for the Ξ−controller are both
nonlinear functions so small−signal transfer functions linearized around a typical
operating point are actually considered in the design process. Figure 6-12 illustrate the
design process: T(s) is first deployed without compensation (T(s)=β·g(s)). Next, adequate
regulators h(s) are introduced in order to achieve acceptable gain margin (GM), phase
margin (PM) and bandwidth (BW). In the SSSC case, only proportional action is needed as
T(s) has a natural infinite
gain [64]. In contrast, the Ξ−controller needs both
DC
proportional and integral action in order to achieve adequate performance. Table 6-13
summarizes the particulars of each design.
SSSC Bode Diagram
-Controller Bode Diagram
200
200
150
150
g(s)|
100
GM=20(dB)
50
50
0
0
-50
g(s)|
100
BW=1.8(Hz)
-50
h(s)g(s)|
-100
270
GM=16(dB)
h(s)g(s)|
-100
360
PM=84o
180
180
90
0
g(s)
PM=86o
g(s)
h(s)g(s)
h(s)g(s)
0
10
BW=1.7(Hz)
-180
-1
10
0
10
1
Freq (Hz)
2
10
3
10
(a)
10
-1
10
0
10
1
2
10
Freq (Hz)
Figure 6-12 Small signal frequency response of loop gain of (a) SSSC, (b) Ξ−controller.
3
10
(b)
179
Table 6-13 Regulator Design Results
Variable
6.5.4
Ξ−Controller
SSSC
-3
3.0 10-7
KP
3.7 10
KI
0
0.16
GM (dB)
20
16
PM (deg)
84
86
BW (Hz)
1.8
1.7
Command Response
Figure 6-13 shows the system response upon changes in the commanded power.
The top trace shows the changes applied to the reference power along with the response
of the line power. The line current, capacitor voltage and the control variables are also
shown in the figure. The comparison of Figure 6-13(a) and (b) shows that the
SSSC
and
the Ξ−controller respond in a very similar manner, as predicted by similar values of GM,
PM and BW.
Instantaneous Line Power
170
Instantaneous Line Power
170
PMEAS
PREF
PMEAS
PREF
160
160
Instantaneous Line Current
Instantaneous Line Current
1.2
1.2
1.1
1.1
DC Link Voltage
0.8
AC Link Voltage
1
0.6
0.4
Control Angle
3
Duty Ratio
80
0
-3
1
2
3
Time (sec)
4
5
(a)
60
1
2
3
4
5
(b)
Time (sec)
Figure 6-13 Computer simulation waveforms illustrating response to command changes. (a) SSSC, (b)
Ξ−controller.
180
6.5.5
Disturbance Response
In order to test the ability of the
SSSC
and the Ξ−controller to maintain control
upon system’s changes, two types of disturbances are considered. Figure 6-14 shows the
system’s response when a voltage sag/swell at the sending end bus occur and when
disturbances on the transmission angle take place. These are four independent events that
have been plotted together for convenience. The two top plots show the disturbances
while the two bottom ones show the response of the line power and the line current,
respectively. The response of the Ξ−controller and the
SSSC
response are similar to each
other.
Voltage Disturbance
Voltage Disturbance
1
1
Sag
Sag
Swell
0.95
Swell
0.95
Transmission Angle Disturbance
Transmission Angle Disturbance
20
20
Increase
Decrease
Increase
Decrease
17
17
Instantaneous Line Power
PREF
170
Instantaneous Line Power
PREF
170
160
160
PMEAS
150
PMEAS
150
Instantaneous Line Current
Instantaneous Line Current
1.2
1.2
1.1
1.1
1
1
1
2
3
Time (sec)
4
5
(a)
1
2
3
Time (sec)
4
5
(b)
Figure 6-14 Computer simulation waveforms illustrating response to bus voltage and angle disturbances.
(a) SSSC, (b) Ξ−controller.
6.6
Summary
This chapter has presented an analytical comparative evaluation of a modern
DC
link based power flow control device, namely the Static Synchronous Series
Compensator and one of the devices introduced in this thesis, namely the Ξ−controller.
181
Although these approaches represent radically different principles of operation to realize
reactive series compensation for power flow control, they feature similar functional
capabilities.
Extensive analytical formulations that determine the various design elements have
been developed in this chapter. They have been used to realize a benchmark case study to
provide numerical values for various design features. The measures of comparison that
have been presented in various tables in the chapter include: transformer ratings,
capacitor ratings, semiconductor ratings, semiconductor losses, waveform quality indices
and dynamic control properties. The analytical solutions and operation of both the
designs have been verified using detailed computer simulations. Selected computer
simulation results that compare the performance have been presented in this chapter.
Based on the comparison of the results from the case study, it may be concluded
that suitably designed
DC
link and
AC
link approaches are both competitive against each
other in realizing series power flow control. The DC link approach requires about twice as
much capacitive energy storage and about 66% additional power semiconductor
MVA
rating, while the AC link approach has about 15% more losses in power semiconductors,
using state of the art semiconductor devices. All other summative measures indicate
similar levels of steady state and dynamic performance.
182
Chapter 7
Experimental
Verification of the Γ−Controller
7.1
Introduction
This chapter describes efforts aimed at experimentally proving the concept of
using a vector switching converter for realizing power flow control. Out of the Π, Ξ and
Γ controllers introduced in Chapter 3, the Γ−controller was selected for experimental
verification as it can be considered the most general case. The reason for this is twofold:
(i) the circuital interconnection involves a shunt−series arrangement and (ii) the converter
involves the realization of a multiple−throw switch.
Figure 7-1 shows a schematic of the system implemented in the laboratory.
VS
S
Bus S
VR
-Controller
SPT
AC
link
VESC
SIT
Transmission
Line
R
Bus R
Figure 7-1 Schematic of the Γ−controller system implemented in the laboratory.
Notice that the above figure strongly resembles Figure 3-20, which was use to introduced
the Γ−controller in Section 3.4. The following sections describe the experiment layout,
183
along with experimental waveforms contrasted against computer simulation results.
7.2
Experiment Layout
As explained in Chapter 3, the Γ−controller operates embedded into a
transmission line. Therefore, a laboratory prototype of a transmission line is discussed
first followed by a laboratory prototype of the Γ−controller itself. Also, given the
availability of resources at the laboratory, the prototype described below is conceived to
operate in a three−phase 230V/60Hz system.
7.2.1
Laboratory Prototype of a Transmission Line
A complete model of a transmission line usually includes four distributed
parameters: series resistance, series inductance, shunt conductance and shunt capacitance.
However, it is customary in power flow controller studies to consider only the effect of
the series elements. Moreover, the distributed nature of these parameters is also neglected
which leads to the representation of the transmission line as shown in Figure 1-1 i.e., as a
series combination of constant resistance and constant inductance [4]. This is
implemented with an adjustable inductor available in the laboratory. The inductor is set at
L=5.6mH and the measured series resistance is R=0.27Ω. This leads to a ratio XL/R of
7.8 which is within the typical values encountered in transmission lines. The inductor is
rated at Imax=40A.
184
Figure 7-2 Laboratory prototype of a transmission line.
As suggested in Figure 7-1, the transmission line connects busbars whose
voltages (magnitude and angle) are stiff quantities. This reflects the operation of actual
power systems where at every busbar the voltage magnitude is kept very close to its
nominal value and the power flows result mostly from angle differences between the
busbars. Figure 7-3 shows a laboratory prototype of a transmission line connecting two
busbars that properly takes into account these effects [65]. As can be seen in the figure, a
phase−shifted transformer connected in the so−called “Marserau configuration” [66] has
been introduced in order to create a transmission angle. This configuration allows
introduction of a transmission angle between the busbars without changing the magnitude
of the voltage. This is shown in the phasor−diagram accompanying the figure. The actual
value of the transmission angle θ is determined by the windings turn ratio. Each
transformer is rated at 230V at the primary side and 36V at the secondary side. This
allows creation of a transmission angle θ=15o.
185
sin
n=
θ Vax / 2
=
2
Va
nH
3
=
n L 2 ⋅ tg θ
2
Figure 7-3 Laboratory prototype of a transmission line connecting two stiff voltage busbars.
7.2.2
Laboratory Prototype of a Γ−Controller
Figure 7-4 shows the schematic of the Γ−controller laboratory prototype. The
blocks labeled as SPT,
AC
link and 3T−1P 3φ VeSC were built for a different application
and in this thesis they have been adapted to operate as a power flow controller. As can be
seen, a realization with one primary and three secondary transformers is being
considered. This leads to the implementation of a triple throw single pole three−phase
vector switching converter for synthesizing the pole voltage. This topology with
Y−connected transformers was studied in detail in Chapter 2. The use of ∆−connected
transformers does not change any of the claims made in Chapter 2 whatsoever. On the
contrary, it adds to the versatility of VeSC−based topologies.
186
Sa
Va1
a1
SIT
Va2
c1
Vai
b1
Va3
Sb
Vb1
b2
A
Vb2
C
a2
B
Vbi
c2
Vb3
SPT
Sc
Vc1
c3
Vb2
b3
Vci
a3
Vc3
AC LINK
3T-1P 3
VESC
Figure 7-4 Laboratory prototype of a Γ−controller.
The ratings of the main components are presented in Table 7-1.
Table 7-1 Γ−controller component rating
Item
SPT Transformer
AC
Capacitor
IGBT−Diode
Ratings
240V/240V 6kVA
30µF 300V 4A
Modules
600V 15A
Fuji 1MB15D−060
SIT Transformer
The
IGBTs
are driven by a standard
230V/36V 5.4kVA
DSP/FPGA/PC
system developed at the University of
Wisconsin [67]. The switching frequency was randomly picked to be FSW=5kHz.
Computer simulations showed that no resonance frequencies are excited at 5kHz. The
converter is embedded in the system as shown in Figure 7-1. For completeness, the entire
system is shown again in Figure 7-5 along with a picture of the hardware built in the
laboratory in Figure 7-6.
Figure 7-5 Detailed laboratory prototype of a Γ−controller system.
187
188
Figure 7-6 Picture of hardware built at the laboratory.
7.3
Experimental Waveforms
Although the system was designed to operate at 230V, experimental waveforms
were captured with the variac set at 100V. This was for the safety of the semiconductors
because no protective devices were installed.
In reference to Figure 7-3, Equation (1-1) allows calculation of the complex
power without the Γ−controller embedded in the transmission line, which is found to be
S=1246 +j 11 (VA). This is the base power which will be subject to alterations by the
insertion of the Γ−controller.
A detailed computer simulation of the Γ−controller system shown in Figure 7-5
was implemented in order to contrast experimental waveforms against theoretical
189
predictions. Two sets of measurements are considered. A steady state case study is used
to show the internal workings of the Γ−controller operation and a dynamic case study is
also implemented in order to show the system level operation of the Γ−controller.
7.3.1
Steady Sate Operation
As mentioned in Section 7.3, the complex power flow through the transmission
line without the Γ−controller embedded is given by S=1246 +j 11 (VA). The insertion of
the Γ−controller into the transmission line allows complex flow control over a region
determined by the voltage synthesized by the converter. Figure 7-7(a) shows that this
voltage can be anywhere inside the triangle formed by the throw voltages of the
converter. However, it was explained in Chapter 3 that the interest is in symmetrical
realizations for the pole voltage and therefore the operation of the converter is limited to
the circle inscribed in the triangle.
1
400
V1
300
200
0.5
Q (VAr)
Im
100
0
Base case
0
Target
-100
-200
-0.5
V2
V3
-300
-1
-1
-0.5
0
Re
0.5
1
(a)
-400
800
900
1000
1100
1200
1300
P (W)
1400
1500
1600
1700
(b)
Figure 7-7 Attainable pole voltage (a) and the corresponding power diagram (b).
Figure 7-7(b) shows the power diagram that results from mapping the pole
190
voltage circle into the P−Q plane. The complex power to generate this diagram is
measured right before the RL branch, at location V2 in Figure 7-5. The power diagram
indicates all reachable values for the complex power that can be obtained by modulating
the duty ratios of the converter. In order to show the operation of the converter, a random
operating point specified in Table 7-2 was selected.
Table 7-2 Operating Point
Item
Value
Duty ratios
d1=33% d2=50% d3=17%
Voltage @ V2 (peak)
V2= 81(V)
Line Current (peak)
IL= 8.4 (A)
Injected Voltage (peak)
Vinj= 3.8 (V)
Active Power
P= 1023 (W)
Reactive Power
Q= −40 (VAr)
Figure 7-8 shows the gate voltage applied to the IGBTs. The top, middle and bottom trace
shows the gate voltage applied to the
IGBTs
corresponding to the throws 1, 2 and 3,
respectively. It should be noted that the average value of those voltages corresponds to
the duty ratios specified in Table 7-2.
191
Figure 7-8 Experimental waveforms showing the gate voltage applied to the IGBTs.
Figure 7-9 and Figure 7-10 show the voltage at location V2 and the line current,
respectively. In both figures, part (a) shows the experimental waveforms while part (b)
shows waveforms obtained from simulations.
125
100
75
50
25
0
-25
-50
-75
-100
-125
Figure 7-9 Waveforms showing the voltage at location V2. (a) Experimental (b) Simulated.
192
Ia
Ib
10
Ic
Ia
Ib
Ic
8
6
4
2
0
-2
-4
-6
-8
-10
Time
(a)
(b)
Figure 7-10 Waveforms showing the line current. (a) Experimental (b) Simulated.
In order to help the visualization, the grid from the simulated waveforms has been
made to match the division from the oscilloscope screen. The match between
experimental and simulated waveforms is evident. In the experimental case, some
unbalance between the phases is observed. This is because the inductors used to emulate
the transmission line were not exactly equal.
Figure 7-11 through Figure 7-13 illustrate the pole voltage synthesized by the
converter. Again, part (a) shows experimental results and part (b) shows the
corresponding simulated waveforms and the grid of the simulated waveforms is made to
match the divisions of the oscilloscope screen. Figure 7-11 shows the actual three−phase
switched voltages, but due to the high frequency switching process taking place the
pictures in both cases are blurry.
193
25
20
15
10
5
0
-5
-10
-15
-20
-25
Time
(a)
(b)
Figure 7-11 Waveforms showing the injected voltage. (a) Experimental (b) Simulated.
Thus, in order to be able to contrast these waveforms, two extra sets of plots are
presented. Figure 7-12 shows the fundamental component of the switched waveforms
depicted above and Figure 7-13 shows an experimental versus simulated comparison of
one of the phases at the switching frequency timescale. Once again, the match between
experiments and predictions made by simulations is evident.
25
20
15
10
5
0
-5
-10
-15
-20
-25
Figure 7-12 Waveforms showing the fundamental component of the injected voltage. (a) Experimental (b)
Simulated.
194
25
20
15
10
5
0
-5
-10
-15
-20
-25
Time
(a)
(b)
Figure 7-13 Waveforms showing the injected voltage at the switching frequency timescale. (a)
Experimental (b) Simulated.
7.3.2
Dynamic Operation
The previous section has shown an experimental verification of the Γ−controller
from the converter’s point of view. The intent of this section is to illustrate the operation
of the Γ−controller from the system’s point of view i.e., to show the transition between
different operating points. These ideas were discussed in Section 3.4 where the
Γ−controller was first introduced. It was shown through computer simulations that the
transitions between operating points are smooth and free of oscillation if the converter’s
duty ratios are slowly ramped from one value into another. Figure 7-14 shows an
example of this. The top trace shows the duty ratios applied to the converter. It can be
seen that the transition takes place in “one division” which corresponds to two seconds.
The next traces show, from top to bottom, the voltage at location V2, the line current and
the injected voltage, respectively.
195
Figure 7-14 Experimental waveforms illustrating the transition from one operating point to another by
means of duty ratio ramping.
As can be seen from all three traces, the transition is smooth and free of
oscillation which confirms the theoretical claim that duty ratio ramping is an appropriate
strategy for the operation of VeSC−based power flow controllers.
7.4
Summary
This chapter has been devoted to discussion of efforts to experimentally prove the
concept of using VeSC−based devices for realizing power flow control in transmission
networks. Out of the three topologies proposed in this thesis, the Γ−controller was
selected for experimental verification as it is seen by the author as the most general case.
196
A Γ−controller prototype was built and tested at the
WEMPEC
laboratory.
Experimental waveforms were contrasted against predictions made by computer
simulations. Several illustrative experimental waveforms were presented which very
closely matched the corresponding simulated waveforms.
Also, the idea of duty ratio ramping to change operating points was implemented
in the hardware. Once again, the results from experiments confirmed theoretical
predictions.
As mentioned above, this experiment did not include the actual measurement of
the complex power. The reason is that the laboratory does not have proper equipment to
measure instantaneous complex power, especially in the presence of harmonics and
unbalanced currents. This is indeed an ongoing research topic that is beyond the scope of
this thesis.
197
Chapter 8
Optimal Power Flow
with AC Link Power Flow
Controllers Embedded
8.1
Introduction
In the previous chapters,
AC
link power flow controllers were analyzed in great
detail, but the focus was on the device itself rather than its function. This chapter is
intended to give a system point of view for the problem of load flow control in electric
power systems with AC link power flow controllers embedded.
Electric power systems can be operated in a variety of ways, depending on the
specific goals to be attained. Typically, such goals may include the minimization of the
overall energy production cost of the system (economic dispatch), the minimization of
the emission of particles (environmental dispatch), the minimization of system losses, or
a combination of these goals [5]. In addition, modern power systems consist of various
control areas, so the optimization of the power exchange between them is an additional
goal that one may want to achieve. This particular issue is of special interest within the
deregulated power markets framework. This is closely related to what is referred to as
loop flow problem, as explained next [68].
Consider a power system consisting of three control areas, as suggested in Figure
198
8-1. The control areas are connected via interfaces, which correspond to a number of
transmission lines that interconnect the corresponding control areas [69]. Assume that,
initially, area A is exchanging power with areas B and C by an amount of PAB and PAC,
respectively.
Figure 8-1. A 3−area electric power system.
In addition, consider that areas A and B want to increase their power exchange by
an amount of ∆PAB, but the power exchange between areas A and C needs to be held
constant, because, for example, the interface is at its limit. In order to perform the
transaction, Area A increases its net power by ∆PAB and Area B decreases its net
production by the same amount. However, the actual flows would be as shown in the
figure. A portion of ∆PAB would effectively flow through the corresponding interface, but
some of it, a loop flow, would go through the Area C, violating the restriction of holding
the AC interface flow constant. This is because power flows are not ruled by contractual
paths but by Kirchhoff’s laws.
Loop flows, as well as the optimal dispatch problem, can be greatly improved by
embedding
FACTS
devices into the transmission system, as has been shown in numerous
199
publications [1, 3, 7, 69-78]. The methodology consists of developing a steady state
model for the corresponding
FACTS
device and incorporating this model into the
so−called Optimal Power Flow problem (OPF).
Next, the standard
OPF
problem is formulated followed by its adaptation when
Ξ−controllers are present. A case study is also discussed.
8.2
Standard OPF
The standard optimal power flow [5] is formulated as follows:
Objective Function
Optimize f (Vb , θ b )
b∈B
B: All system Buses
(8-1)
g∈G
G: Generation Buses
(8-2)
Equality Constraints
Pg = P(Vb , θ b ) + PgL
Q g = Q(Vb , θ b ) + Q gL
PdL = P(Vb , θ b )
(8-3)
d∈D
D: Load Buses
Q dL = Q(Vb , θ b )
(8-4)
(8-5)
Inequality Constraints
Vbmin ≤ Vb ≤ Vbmax
max
θ min
bS ≤ θ bS ≤ θ bS
(8-6)
b S ∈ BS
BS: Same as B excluding slack bus
Pgmin ≤ Pg ≤ Pgmax
(8-7)
(8-8)
Q gmin ≤ Q g ≤ Q gmax
g∈G
| S l (Vb , θ b ) |≤ S lmax
l∈L
(8-9)
L: Transmission Lines
(8-10)
200
The state of a power system is fully defined by the magnitude and angle of the
bus voltages, Vb’s and θb’s, so they are used as explicit variables in the problem
formulation.
Equation (8-1) represents the objective function to be optimized. In the economic
dispatch case, it corresponds to the minimization of the energy production costs
summation,
Min
∑C
g
(Pg (Vb , θ b ))
g∈G
(8-11)
g
where Cg represents the cost function of the various generators. If the goal is to maximize
the power exchange through an interface, the objective function becomes,
Max
∑ P (V , θ
i
b
b
)
i∈I
(8-12)
i
where I represents the set of transmission lines belonging to the interface. Other objective
function can be readily formulated; and via the Pareto’s optimality criterion [76],
multiple objective functions can be pursued.
The equality constraints described by Equations (8-2) through (8-5) represent the
power flow equations that dictate the physical restrictions of the problem. They force a
balance between the power injected into the network by the generators and the power
demanded by the loads plus the transmission losses.
201
Finally, the inequality constraints described by Equations (8-6) through (8-10)
represent the operational limits of the various components. Equations (8-6) and (8-7) say
that every bus in the system has to operate within acceptable voltage ranges. Equations
(8-8) and (8-9) dictate the capacity limits for generators. Lastly, Equation (8-10) specify
the most important restriction within the
FACTS
context, saying that transmission lines
have a limited capacity to carry power. This transmission line limit can be expressed in
terms of either current, active power, or apparent power. The last approach is chosen
here.
8.3
8.3.1
OPF with Ξ−controllers embedded
The model
The steady state model for the Ξ−controller readily follows from its equivalent
circuit. Aggregating the distributed parameters of Figure 3-14, the following simplified
model can be obtained,
Figure 8-2 Simplified Equivalent Circuit for the Ξ−controller.
where X L = ω P (L S + L R + L T ) , R L = R S + R R , and X C = 1 / ω P C . Thus, the overall
transmission line impedance can now be thought of as,
202
Z = R L + j (X L − d 2 X C )
(8-13)
The additional degree of freedom introduced by the controllable variable d can be
interpreted as having a tunable line reactance within range X L − X C ≤ X ≤ X L . Once the
value for X has been determined, the corresponding duty ratio can be calculated as,
d=
XL − X
XC
8.3.2
(8-14)
OPF
Formulation
The OPF problem for a power system with Ξ−controllers embedded can be
formulated as follows,
Objective Function
Optimize f (Vb , θ b , X ξ )
b∈B
B: All system Buses
ξ∈Ξ
Ξ: Lines with a Ξ−controller
g∈G
G: Generation Buses
(8-15)
Equality Constraints
Pg = P(Vb , θ b , X ξ ) + PgL
Q g = Q(Vb , θ b , X ξ ) + Q gL
PdL = P(Vb , θ b , X ξ )
Q dL = Q(Vb , θ b , X ξ )
(8-16)
(8-17)
d∈D
D: Load Buses
(8-18)
(8-19)
203
Inequality Constraints
Vbmin ≤ Vb ≤ Vbmax
max
θ min
bS ≤ θ bS ≤ θ bS
(8-20)
b S ∈ BS
BS: Same as B excluding slack bus
(8-21)
Pgmin ≤ Pg ≤ Pgmax
(8-22)
Q gmin ≤ Q g ≤ Q gmax
(8-23)
| S l (Vb , θ b , X ξ ) |≤ S lmax
L: Transmission Lines
l∈L
X Lξ − X Cξ ≤ X ξ ≤ X Lξ
(8-24)
(8-25)
Equations (8-15) through (8-24) have the same interpretation as in the standard
OPF formulation, but now they feature an extra degree of freedom introduced by the
explicit variables Xξ. Equation (8-25) forces the corresponding duty ratios for Xξ to be
within admissible range ( 0 ≤ d ξ ≤ 1 ).
When control areas are present, additional constraints in the form of inequalities
or equalities can be added on the corresponding interfaces in order to satisfy operating
restrictions. For example, to avoid the loop flow between areas A and B in the
introductory discussion, the following equality constraints can be added,
Additional Equality Constraints
∑ P (V , θ
b
) = PAC
i ∈ I AC
IAC: AC Interface
(8-26)
∑ P (V , θ
b
) = PBC
i ∈ I BC
IBC: BC Interface
(8-27)
i
b
i
i
i
b
where PAC and PBC represent pre−established values for the total power exchange
204
between the areas.
The OPF problem described above was implemented in the MATLAB environment.
It clearly corresponds to a non−linear optimization problem. The optimality conditions,
known as Karush−Kuhn−Tucker (KKT) conditions, are obtained by successively
approximating the objective function by a quadratic function and the constraints by linear
functions. This problem, known as a quadratic programming (QP), is solved by standard
methods described in the literature [79, 80].
As will become evident in the case study discussed below, the presence of the
Ξ−controllers allows for more flexible operation of the overall system, yielding better
results in the objective function pursued.
8.3.3
Amount and Location
In the OPF formulation discussed above, nothing was said about the nature of the
set Ξ (the set of transmission lines containing Ξ−controllers), in terms of how many
devices should be used and where to locate them. The number of controllers as well as
the allocation within the system has been the subject of many publications throughout the
last few years [69, 74, 78, 81, 82]. Results vary depending on the objective function to be
pursued and the nature of the controller considered. This topic however, is beyond the
scope of this thesis. In addition, since the Ξ−controller is functionally similar to the SSSC
and TCSC, previous studies for these devices can be utilized for this purpose.
205
8.3.4
Case Study
In order to appreciate the flexibility resulting from the presence of the
Ξ−controller, consider the
IEEE
39−bus system sketched in Figure 8-3. This system
represents a simplification of the New England interconnection, in the Northeastern U.S.
The 39−bus system has 10 generators, 19 loads, 35 transmission lines and 12
transformers. In addition, the system is organized into three areas. The complete set of
data is presented in Appendix 5.
32
9
31
8
Buses
Loads
Lines
Generators
10
34
33
11
7
36
13
6
Area C
5
20
12
23
14
19
39
4
24
IAC
15
22
IBC
21
3
17
2
Area A
27
18
IAB
Area B
30
35
16
1
38
26
28
25
29
37
Figure 8-3 The IEEE 39−bus system.
The study considers the economic dispatch
OPF
and the interface power
maximization OPF separately. In each case, the results obtained with the system as is i.e.,
206
without any Ξ−controller, are compared against the results obtained with a fully
controlled system i. e., with a Ξ−controller embedded in every transmission line. The
amount of maximum compensation for each Ξ−controller is assumed to be 50% of the
corresponding transmission line reactance. Thus, the value of the equivalent reactance of
each line can vary within the range,
XL
≤ Xξ ≤ XL
2
8.3.4.1
ξ∈L
L: Transmission Lines
(8-28)
Economic Dispatch
For the economic dispatch
OPF,
Equation (8-11) describes the objective function.
The cost function for each generator is assumed quadratic, as shown in TABLE II,
Appendix 5.
(i)
Results without Ξ−controllers embedded
Objective function: 98,674 ($/h)
Table 8-1 Active Constraints
Bus
Max V
Min V
38
20, 30, 33, 36
Generator
Max P
Max Q
Min Q
30, 31, 32, 36, 31, 32, 39
37, 38, 39
Line
Max MVA
16-17
Since this case will be used as base case for the interface power maximization
discussion, these results are also provided,
207
Table 8-2 Interface Flows
Line
28-26
16-17
29-26
Total
A− B
Rating
P
(MVA) (MW)
500
181
500
499
500
232
1500
912
A− C
P
Rating
(MVA) (MW)
15-14
500
178
Total
500
178
Line
B−C
P
Rating
(MVA) (MW)
1-39
500
363
3-4
500
102
Total 1000
465
Line
In the above table, the total values represent the power exchange between the
corresponding control areas.
(ii)
Results with Ξ−controllers embedded
Objective function: 80,106 ($/h)
Table 8-3 Active Constraints
Bus
Max Min
V
V
10,
19,
22,
25,
29,
36
Generator
Max Max Min
P
Q
Q
30,
31
31,
33,
36,
37,
38,
39
Line
Max
MVA
16-17
For this case, Equation (8-28) is added into the
Equivalent Impedance
Max
Min
3-18, 4-14,
2-3, 2-25,
5-6, 6-7,
3-4, 4-5,
10-13, 13-14, 7-8, 8-9,
17-27, 21-29, 9-39, 14-15,
25-26, 16-17, 15-16, 16-19,
16-21, 16-24,
19-20
17-18, 21-22,
22-23, 28-29,
12-13, 6-31,
23-36
OPF
formulation. As can be seen,
embedding Ξ−controllers into the transmission lines leads to a reduction of almost 20%
in the operation cost of the overall system. An adequate economic evaluation based on
the cost benefit ratio (C/B) would give insights about whether it is worth it to invest in
208
these devices [83, 84].
In addition, the active constraints on the equivalent impedance of the lines give
valuable information about the real need of Ξ−controllers. Thus, as may be implied from
Equation (8-25), the equivalent impedances at their maximum limit are not receiving any
compensation from the Ξ−controllers whatsoever and may therefore be eliminated.
Similarly, equivalent impedances at their minimum limit are receiving 100% of
compensation indicating a potential need for an even larger compensator.
Finally, although the active constraints on the bus voltages and generators have
changed, the interface AB remains congested, limiting Area B from obtaining cheaper
power from Area A. This leads to the next discussion on power exchange between
control areas.
8.3.4.2 Interface Power Maximization
As was seen in the previous case study, the transmission line 16-17 is acting as a
bottleneck binding further power exchange from Area A to B. Say the interest is in
somehow increasing the power transfer through the Interface AB. However, due to
contractual paths restrictions [71], this must be done while keeping the power exchange
between Areas A and C and Areas B and C at their nominal values specified in Table 8-2.
For the situation described above, Equation (8-12) describes the
OPF
objective
function while Equations (8-26) and (8-27) force the total power exchange through
interfaces AC and BC to remain constant.
209
Results without Ξ−controllers embedded
(i)
Objective function: 968 (MW)
Table 8-4 Interface Flows
Line
28-26
16-17
29-26
Total
A− B
Rating
P
(MVA) (MW)
500
209
500
499
500
260
1500
968
A− C
P
Line Rating
(MVA) (MW)
15-14
500
178
Total
500
178
B−C
P
Line Rating
(MVA) (MW)
1-39
500
363
3-4
500
102
Total 1000
465
As can be seen, although it is possible to boost the A−C power transfer without
using controllable devices, the actual increase is 56(MW), only a 6% of the nominal
value.
(ii)
Results with Ξ−controllers embedded
Objective function: 1288 (MW)
Table 8-5 Interface Flows
Line
28-26
16-17
29-26
Total
A− B
Rating
P
(MVA) (MW)
500
368
500
499
500
421
1500
1288
A− C
P
Line Rating
(MVA) (MW)
15-14
500
178
Total
500
178
B−C
Line Rating
P
(MVA) (MW)
1-39
500
454
3-4
500
11
Total 1000
465
210
As in the previous case study, a considerable improvement in the objective
function is now observed. The power exchange has been boosted by 376(MW), a 41%
with respect to its nominal value. Notice that the flows through the Interface
BC
have
readjusted but the total exchange has remained constant.
8.4
Summary
This chapter has given a system point of view for AC link power flow controllers.
It has been demonstrated that, by embedding Ξ−controllers into the transmission
network, dramatic improvements can be achieved in the flexibility of operation for the
overall power system. It was shown that the system model of the Ξ−controller can be
thought as having transmission lines with tunable reactance. Using the
IEEE
39−bus
system as an example, typical operational issues were addressed. Specifically, the
optimal power flow formulation showed that the economic operation as well as the loop
flow problem can be greatly improved by adding the extra degree of freedom provided by
Ξ−controller.
211
Chapter 9 Conclusions and Future
Work
This thesis has introduced a new family of
FACTS
devices: the Π−controller, the
Ξ−controller and the Γ−controller. Based on the vector switching converter (VeSC), these
devices feature similar functional characteristics to those encountered in the voltage
source HVDC interties, the Static Synchronous Series Compensator (SSSC) and the Unified
Power Flow Controller (UPFC), respectively. The control function, however, is realized
by direct
AC↔AC
conversion without frequency change, adopting the strategy of pulse
width modulation (PWM).
A comprehensive analysis of each of these devises has been performed, including
operating principles, equivalent circuits, design considerations, and dynamic modeling
and control. Computer simulations of illustrative examples were executed to validate the
approach. Due to the level of power involved, it was concluded that the Π−controller
may not be practical with the current state−of−the−art gate turn−off semiconductors.
However, the Ξ−controller and Γ−controller would be able to compete with their DC link
counterparts. A case study for series compensation has shown the Ξ−controller is indeed
favorably comparable to the SSSC.
The dynamic modeling for these devices was accomplished using generalized
212
averaging theory. This approach, in contrast with the traditional dynamic phasors model,
explicitly incorporates the switching frequency into the state space model, and therefore
is able to predict oscillatory behavior of the state variables due to switching and thus
preclude potential instabilities in cases when there are system modes in the vicinity of the
switching frequency. It was also shown that it can be used for design purposes, as it can
predict component stress levels as RMS quantities with a higher degree of accuracy than
the traditional model.
Feedback control of power flow using the proposed devices was also investigated.
For the Ξ−controller case, it was shown that closed loop regulation can be approached by
means of classical control design techniques, such us Bode plots and Nyquist’s stability
criterion. For the Γ−controller, however, more sophisticated techniques are needed.
A laboratory prototype of the Γ−controller developed at the
WEMPEC
laboratory
has proved further the real viability of the use of VeSC−based topologies for power flow
control. Laboratory results indicate excellent agreement with the analytical predictions
and computer simulations.
Finally, a system point of view was given, incorporating the Ξ−controller into an
Optimal Power Flow (OPF) program. It was shown that the Ξ−controller adds extra
flexibility to the
OPF
program and dramatically improves the objective function for
different goals pursued.
213
9.1
Contributions
The main contribution of this thesis is the introduction of
AC
link VeSC-based power
flow controllers. Within this context, several specific contributions are enumerated
below.
9.1.1
Modeling and control
• Development of equivalent circuits based on controlled sources as presented here
allows easy representation of VeSC-based topologies for circuit analysis, computer
simulations, steady state and dynamic interaction between the converter and the
rest of system, and sizing the various components (capacitors, semiconductors,
transformers).
• Novel approach for dynamic modeling based on generalized averaging theory
applicable to VeSC-based topologies when the ratio between switching frequency
and power frequency is an integer number leads to capturing the oscillatory
behavior of the state variables when the switching process excites the physical
resonances.
• The development of the small signal loop gain of the power flow control permits
application of classical feedback control design with ease.
9.1.2
Computer simulation
• Switched simulations modeling the switches as switching function−controlled
sources allow capturing the dynamic of the switching process and can be used to
214
estimate the behavior of the converter during commutation events. The main
drawback of switched simulations is the relatively long time they take to run.
• Averaged simulation modeling the switches as duty ratio−controlled sources are
useful when the interest is in the interaction between the fundamental quantities
only. They can be used to test the selection of feedback control gains and for
system level dynamic behavior. They run much faster than switched simulations.
9.1.3
Comparison against the state−of−the−art technology for
series compensation
A detailed comparison between the
SSSC
and the Ξ−controller was performed.
The particulars considered in the comparative evaluation are as follows:
• Detailed design considerations of power circuit components − capacitors and
semiconductors
• Topological approach integrating transformers and appropriate modulation
techniques for meeting benchmark harmonic performance levels.
• Semiconductors loss modeling using analytical techniques to allow fast
computation of thermal performance
• Injection transformer design considerations including core loss components
• Design and verification of closed loop regulator with more than adequate dynamic
performance
• Computer simulations verifying functional operation and controller performance
during command changes and disturbances.
215
9.1.4
Experimental verification
• Experimental waveforms were contrasted against computer simulation predictions.
This validated the concept of using VeSC-based topologies for power flow control.
• Experimental verification of dynamic operation via duty ratio damping.
Experiments confirmed that the best strategy to move from one operating point on
to another was via duty ratio ramping. Duty ratio ramping leads to a smooth
transition of the line current with no oscillations or overshoot.
9.1.5
System level modeling
• Incorporation of
AC
link VeSC-based series power flow controllers into the optimal
power flow problem. It was shown that by embedding
VeSC-based
series power
flow controllers in the transmission lines, a much flexible operation of the system
is achieved.
9.2
Proposed Continuing Research
Several opportunities for future investigation have been identified that would
provide valuable extensions of the research conducted in this thesis.
9.2.1
Feedback Control for Γ−controller
The problem of power flow control using the Γ−controller being a MIMO system
with multiple degrees of freedom was stated in Section 5.3 referring to Figure 5-10 [48].
216
The aim is to find an appropriate regulator that can guarantee stable operation of the
overall system as well as an adequate dynamic performance, in terms of
DC
error, speed
of response, overshot and oscillations. Since the Γ−controller has been conceived as a
complex power flow controller, a proper method to measure instantaneous three−phase
reactive power in the presence of harmonics is seen as a critical part of the
implementation of the feedback regulator. Advanced control techniques such Mode
Predictive Control (MPC) [85] and Passivity Based Control (PBC) [86] are seen by the
author as particularly suitable for developing the control structure. There are several
needs and opportunities for investigations of feedback control for the Γ−controller, as
well as other topologies.
9.2.2
OPF
with embedded Γ−controllers
The development of a suitable model for the Γ−controller to be incorporated into
an OPF program. A research group at CINVESTAV in Mexico has recently initiated work on
this topic [87, 88].
9.2.3
System level dynamics
The performance of
AC
link power flow controllers for system level dynamics
present wide opportunities for research work. System level dynamics topics include
subsynchronous resonance damping, power oscillations damping, transient stability
enhancement, voltage stability, etc. Numerous publications have shown the performance
of
DC
link power flow controllers in this area [1-3, 14]. Similar studies using
AC
link
217
power flow controllers may reveal further tradeoffs between these two technologies. The
same Mexican group has been working on this topic as well [87].
9.2.4
Multilevel realization for the Ξ−Controller
This thesis presented the Ξ−controller as a series device whose principle of
operation was based on commuting between a three−phase capacitor bank and a short
circuit node. A natural expansion of this concept that may lead to better performance is to
use multilevel converters [30] extended to AC link. This would be achieved by using one
(or more) three−phase capacitor bank instead of (or in combination with) the short circuit
node.
9.2.5
Additional topologies
This thesis has studied the use of
VeSC–based AC
link power flow controllers
connected in back−to−back, series and shunt−series configurations. As pointed out in the
introduction, DC link power flow controllers can be configured, in addition of these three
configurations just mentioned, as shunt (STATCOM) and series−series (IPFC) devices.
Topologies based on
AC
link VeSC can also be derived for these configurations, which
could bring forth a treasure−trove of research opportunities. For example, the
VeSC−based
AC
link
shunt configuration was studied for a flicker mitigation application [89].
This topology can readily be extended to operate as an AC link STATCOM.
218
9.2.6
Field demonstration
It is the hope of the author that the ideas discussed in this thesis may become a
reality through a field demonstration in a utility power system. This thesis has established
that this technology is viable and the natural continuation of this work would be the
actual construction of a utility scale AC link VeSC−based power flow controller.
219
Appendix 1
Parameters for the
Π−controller case study
Symbol
VS
RS
XS
VR
RR
XR
FP
TABLE I
SYSTEM PARAMETERS
Actual value
Per unit
110kV 100MVA
value
60Hz
TABLE II
TRANSFORMER STP PARAMETERS
Actual value
Symbol Per unit value 110/11/6.3/11/6.3kV
180MVA 60Hz
1
0.01
0.17
1
0.01
0.17
1
VP
RP
XP
VS1
RS1
XS1
VS2
RS2
XS2
VS3
RS3
XS3
VS4
RS4
XS4
110kV
1.21 Ω
54.5 mH
110kV
1.21 Ω
54.5 mH
60 Hz
TABLE III
FILTER CAPACITOR FTC PARAMETERS
Symbol
Per unit value
XC
3
Actual value
11 kV 100MVA 60Hz
730 µF
1
0.001
0.01
1
0.001
0.01
1
0.001
0.01
1
0.001
0.01
1
0.001
0.01
TABLE IV
TRANSFORMER BIT PARAMETERS
Actual value
Per unit
Symbol
5.5/110 kV 180MVA
value
60 Hz
VP
RP
XP
VS
RS
XS
1
0.0001
0.01
1
0.0001
0.01
TABLE V
CONVERTER PARAMETERS
Actual Value
Symbol Per unit Value
60 Hz
FS
40
110kV
0.2 Ω
5.3 mH
11kV
2 mΩ
53 µH
6.3kV
0.67 mΩ
18 µH
11kV
2 mΩ
53 µH
6.3kV
0.67 mΩ
18 µH
2400 Hz
5.5 kV
16 µΩ
5.3 mH
110kV
6.7 mΩ
53 µH
220
Appendix 2
Parameters for the
Ξ−controller case study
Symbol
VS
RS
XS
VR
RR
XR
FP
TABLE I
SYSTEM PARAMETERS
Actual value
Per
unit
110kV 100MVA
value
60Hz
1
0.01
0.07
1
0.01
0.17
1
110kV
1.21 Ω
22.4 mH
110kV
1.21 Ω
54.5 mH
60 Hz
TABLE III
TRANSFORMER SIT PARAMETERS
Actual value
Per unit
Symbol
15.5/15.5 kV 20MVA
value
60Hz
VSP
RSP
XSP
VSS
RSS
XSS
RPM
XPM
1
0.0001
0.01
1
0.0001
0.01
500
500
15.5 kV
1.2 mΩ
0.3 mH
15.5kV
1.2 mΩ
0.3 mH
6 kΩ
16 H
Symbol
XC
RC
TABLE II
CAPACITOR PARAMETERS
Actual value
Per unit
15.5 kV 20MVA
value
60Hz
1
500
438 µF
6 kΩ
TABLE IV
CONVERTER PARAMETERS
Symbol
Per unit value
Actual value 60Hz
FS
40
2.4kHz
221
Appendix 3
Parameters for the
Γ−controller case study
Symbol
VS
RS
XS
VR
RR
XR
FP
TABLE I
SYSTEM PARAMETERS
Actual value
Per
unit
110kV 100MVA
value
60Hz
TABLE II
TRANSFORMER SPT PARAMETERS
Actual value
Per
unit
Symbol
110/15.5/8.9/15.5/8.9kV
value
20MVA 60Hz
1
0.01
0.07
1
0.01
0.17
1
VPP
RPP
XPS
VPS1
RPS1
XPS1
VPS2
RPS2
XPS2
VPS3
RPS3
XPS3
VPS4
RPS4
XPS4
RPM
XPM
110kV
1.21 Ω
22.4 mH
110kV
1.21 Ω
54.5 mH
60 Hz
1
0.001
0.01
1
0.001
0.01
1
0.001
0.01
1
0.001
0.01
1
0.001
0.01
500
500
110kV
0.6 Ω
16.0 mH
15.5kV
12 mΩ
0.3 mH
8.9kV
4 mΩ
0.1 mH
15.5kV
12 mΩ
0.3 mH
8.9kV
4 mΩ
0.1 mH
302 kΩ
800 H
TABLE III
FILTER CAPACITOR FTC PARAMETERS
Actual value
Per unit
Symbol
15.5 kV 20MVA
value
60Hz
TABLE IV
TRANSFORMER SIT PARAMETERS
Actual value
Per unit
Symbol
15.5/15.5 kV 20MVA
value
60Hz
XC
RC
VSP
RSP
XSP
VSS
RSS
XSS
RPM
XPM
3
500
73 µF
6 kΩ
1
0.0001
0.01
1
0.0001
0.01
500
500
TABLE V
CONVERTER PARAMETERS
Symbol
Per unit value
Actual value 60Hz
FS
40
2400Hz
15.5 kV
1.2 mΩ
0.3 mH
15.5kV
1.2 mΩ
0.3 mH
6 kΩ
16 H
222
Appendix 4 Parameters zi for the
Γ−controller state space description
z0=(4*Lsp*as^2*Ls*ap^2+4*Lsp*as^2*Lpp*ap^2+4*Lr*Ls*ap^2+...
4*Lr*Lpp*ap^2+4*Ls*Lpp*ap^2+4*Lss*Ls*ap^2+4*Lss*Lpp*ap^2+...
Lsp*as^2*Lps+Ls*Lps+Lss*Lps+Lr*Lps);
z1=ap*(Rs*Lsp*as^2+Rs*Lr+Rs*Lss-Rss*Ls-Rr*Ls-Rsp*as^2*Ls)/z0;
z2=1/Lps*ap*(Lsp*as^2*Lpp*ap+Lsp*as^2*Ls*ap+Lss*Lpp*ap+...
Lss*Ls*ap+Ls*Lpp*ap+Lr*Lpp*ap+Lr*Ls*ap)/z0;
z3=((4*Ls*Rpp-4*Rs*Lpp)*ap^3-(-Ls*Rps+Rs*Lps)*ap)/z0;
z4=(-Rsp*as^2*Lps-4*ap^2*Lpp*Rsp*as^2-4*ap^2*Lpp*Rss...
-4*ap^2*Lpp*Rr-4*ap^2*Ls*Rss-4*ap^2*Rs*Lpp-4*...
ap^2*Ls*Rr-Rr*Lps-Rss*Lps-Rs*Lps-4*ap^2*Ls*Rsp*as^2)/z0;
z5=1/Lps*ap^2*(Lpp*Rps*Lr+Lpp*Rps*Lss+Lpp*Rps*Ls+...
Rps*Ls*Lr+Rps*Ls*Lss-Rs*Lps*Lr-Rs*Lps*Lsp*as^2-...
Rs*Lps*Lss+Rps*Ls*Lsp*as^2+Lpp*Rps*Lsp*as^2-Rpp*Lps...
*Lsp*as^2-Rpp*Lps*Ls-Rpp*Lps*Lss-Rpp*Lps*Lr)/z0;
z6=as*(Lps+4*ap^2*Lpp+4*ap^2*Ls)/z0;
z7=1/Lps*ap*as*Ls*Lps/z0;
z8=ap*Ls/z0;
z9=(4*Lpp*ap^2+Lps)/z0;
z10=(-(4*Lpp+4*Ls)*ap^2-Lps)/z0;
z11=ap*(Lr+as^2*Lsp+Lss)/z0;
223
Appendix 5
Number
Type
Pd (MW)
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
2
3
2
2
2
2
2
2
2
2
0
0
322
500
0
0
233.8
522
0
0
0
8.5
0
0
320
329.4
0
158
0
680
274
0
247.5
308.6
224
139
281
206
283.5
0
9.2
0
0
0
0
0
0
0
1104
TABLE I
BUS DATA
Qd (MVA) Vmax (pu)
0
0
2.4
184
0
0
84
176.6
0
0
0
88
0
0
153
32.3
0
30
0
103
115
0
84.6
-92.2
47.2
17
75.5
27.6
26.9
0
4.6
0
0
0
0
0
0
0
250
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
1.1
OPF Data
Vmin (pu)
θmax °
θmin °
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
0.9
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
-90
224
TABLE II
GENERATOR DATA
Number
Pmax (MW)
30
31
32
33
34
35
36
37
38
39
Pmin (MW)
300
500
750
1000
1000
1000
1000
400
1000
500
Qmax (MVA)
0
0
0
0
0
0
0
0
0
0
From
To
R(pu)
X(pu)
Rating (MVA)
1
1
2
2
3
3
4
4
5
5
6
6
7
8
9
10
10
13
14
15
16
16
16
2
39
3
25
4
18
5
14
6
8
7
11
8
9
39
11
13
14
15
16
19
21
24
0.0035
0.001
0.0013
0.007
0.0013
0.0011
0.0008
0.0008
0.0002
0.0008
0.0006
0.0007
0.0004
0.0023
0.001
0.0004
0.0004
0.0009
0.0018
0.0009
0.0016
0.0008
0.0003
0.0411
0.025
0.0151
0.0086
0.0213
0.0133
0.0128
0.0129
0.0026
0.0112
0.0092
0.0082
0.0046
0.0363
0.025
0.0043
0.0043
0.0101
0.0217
0.0094
0.0195
0.0135
0.0059
2500
500
2500
2500
500
2500
2500
2500
2500
2500
2500
2500
2500
2500
2500
2500
2500
2500
500
2500
2500
2500
2500
Qmin (MVA)
300
500
750
1000
1000
1000
1000
400
1000
500
TABLE III
BRANCH DATA
From
Tap
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
17
17
21
21
22
23
25
26
26
16
26
28
12
12
6
10
19
20
22
23
25
2
29
19
-300
-500
-750
-1000
-1000
-1000
-1000
-400
-1000
-500
Cost ($/h)
aP2+bP+c
a
0.06
0.01
0.05
0.01
0.04
0.07
0.01
0.02
0.006
0.006
b
0.1
0.4
0.3
0.3
0.2
0.3
0.5
0.7
0.3
0.3
c
0
0
0
0
0
0
0
0
0
0
To
R(pu)
X(pu)
Rating (MVA)
18
27
22
29
23
24
26
27
28
17
29
29
11
13
31
32
33
34
35
36
37
30
38
20
0.0007
0.0013
0.0008
0.0008
0.0006
0.0022
0.0032
0.0014
0.0043
0.0007
0.0057
0.0014
0.0016
0.0016
0
0
0.0007
0.0009
0
0.0005
0.0006
0
0.0008
0.0007
0.0082
0.0173
0.014
0.014
0.0096
0.035
0.0323
0.0147
0.0474
0.0089
0.0625
0.0151
0.0435
0.0435
0.025
0.02
0.0142
0.018
0.0143
0.0272
0.0232
0.0181
0.0156
0.0138
2500
2500
2500
2500
2500
2500
2500
2500
500
500
500
2500
2500
2500
2500
2500
2500
2500
2500
2500
2500
2500
2500
2500
Tap
0
0
0
0
0
0
0
0
0
0
0
0
1.006
1.006
1.07
1.07
1.07
1.009
1.025
1
1.025
1.025
1.025
1.06
225
References
[1]
N. G. Hingorani, "FACTS technology and opportunities," in proceeding of the
IEE Colloquium on Flexible AC Transmission Systems (FACTS), London, UK,
1994.
[2]
N. G. Hingorani, "Power electronics in electric utilities: role of power electronics
in future power systems," Proceedings of the IEEE, vol. 76, 1988.
[3]
N. G. Hingorani, "Role of FACTS in a deregulated market," in proceeding of the
Power Engineering Society Summer Meeting, 2000.
[4]
A. R. Bergen and V. Vittal, "Power systems analysis," 2nd ed. Upper Saddle
River, NJ: Prentice Hall, 2000.
[5]
A. J. Wood and B. F. Wollenberg, "Power generation, operation, and control,"
2nd ed: J. Wiley & Sons, 1996.
[6]
N. Mohan, T. M. Undeland, and W. P. Robbins, "Power electronics: converters,
applications, and design," 3rd ed: John Wiley & Sons, 2003.
[7]
Y.-H. Song and A. Johns, "Flexible ac transmission systems (FACTS)." London:
Institution of Electrical Engineers, 1999.
[8]
M. D. Ilic and J. Zaborszky, "Dynamics and control of large electric power
systems," Wiley, 2000.
[9]
M. R. Iravani, P. L. Dandeno, K. H. Nguyen, D. Zhu, and D. Maratukulam,
"Applications of static phase shifters in power systems," IEEE Transactions on
Power Delivery, vol. 9, 1994.
226
[10]
M. R. Iravani and D. Maratukulam, "Review of semiconductor-controlled (static)
phase shifters for power systems applications," IEEE Transactions on Power
Systems, vol. 9, 1994.
[11]
B. K. Johnson and G. Venkataramanan, "A hybrid solid state phase shifter using
PWM AC converters," IEEE Transactions on Power Delivery, vol. 13, 1998.
[12]
B. K. Johnson, G. Venkataramanan, and M. Manjrekar, "A hybrid solid state
phase shifter for continuous phase shift control in transmission lines," in
proceeding of the Canadian Conference on Electrical and Computer Engineering,
1996.
[13]
GE Industrial and Power Systems, "HVDC Handbook."
[14]
N. G. Hingorani and L. Gyugyi, "Understanding FACTS: concepts and
technology of flexible AC transmission systems," IEEE Press, 2000.
[15]
L. Gyugyi, "Dynamic compensation of AC transmission lines by solid-state
synchronous voltage sources," IEEE Transactions on Power Delivery, vol. 9,
1994.
[16]
L. Gyugyi, C. D. Schauder, and K. K. Sen, "Static synchronous series
compensator: a solid-state approach to the series compensation of transmission
lines," IEEE Transactions on Power Delivery, vol. 12, 1997.
[17]
L. Gyugyi, C. D. Schauder, S. L. Williams, T. R. Rietman, D. R. Torgerson, and
A. Edris, "The unified power flow controller: a new approach to power
transmission control," IEEE Transactions on Power Delivery, vol. 10, 1995.
[18]
L. Gyugyi, K. K. Sen, and C. D. Schauder, "The interline power flow controller
concept: a new approach to power flow management in transmission systems,"
IEEE Transactions on Power Delivery, vol. 14, 1999.
227
[19]
J. Sun, L. Hopkins, B. Shperling, B. Fardanesh, M. Graham, M. Parisi, S.
MacDonald, S. Bhattacharya, S. Berkowitz, and A. Edris, "Operating
characteristics of the convertible static compensator on the 345 kV network," in
proceeding of the IEEE PES Power Systems Conference & Exposition, New
York, NY, USA, 2004.
[20]
C. Hochgraf, R. Lasseter, D. Divan, and T. A. Lipo, "Comparison of multilevel
inverters for static VAr compensation," in proceeding of the Industry
Applications Conference IAS Annual Meeting, Denver, CO, USA, 1994.
[21]
P. W. Wheeler, J. Rodriguez, J. C. Clare, L. Empringham, and A. Weinstein,
"Matrix converters: a technology review," IEEE Transactions on Industrial
Electronics, vol. 49, 2002.
[22]
B. T. Ooi and M. Kazerani, "Voltage-source matrix converter as a controller in
flexible AC transmission systems," IEEE Transactions on Power Delivery, vol.
13, 1998.
[23]
B.-T. Ooi and M. Kazerani, "Unified power flow controller based on matrix
converter," in proceeding of the IEEE Power Electronics Specialists Conference
(PESC), 1996.
[24]
F. Mancilla-David and G. Venkataramanan, "A Synchronous Back-to-Back
Intertie Using the Vector Switching Converter," in proceeding of the Thirty-Sixth
Annual North American Power Symposium (NAPS), Moscow, Idaho, 2004.
[25]
F. Mancilla-David and G. Venkataramanan, "A pulse width modulated AC link
unified power flow controller," in proceeding of the IEEE Power Engineering
Society General Meeting, San Francisco, CA, 2005.
[26]
G. Venkataramanan, "Three-phase vector switching converters for power flow
control," IEE Proceedings on Electric Power Applications, vol. 151, 2004.
228
[27]
G. Venkataramanan and B. K. Johnson, "Pulse width modulated series
compensator," IEE Proceedings on Generation, Transmission and Distribution,
vol. 149, 2002.
[28]
G. Venkataramanan, B. K. Johnson, and A. Sundaram, "An AC-AC power
converter for custom power applications," IEEE Transactions on Power Delivery,
vol. 11, 1996.
[29]
D. Divan, W. Brumsickle, R. Schneider, B. Kranz, R. Gascoigne, D. Bradshaw,
M. Ingram, and I. Grant, "A distributed static series compensator system for
realizing active power flow control on existing power lines," in proceeding of the
IEEE PES Power Systems Conference & Exposition, New York, NY, USA, 2004.
[30]
G. Holmes and T. A. Lipo, "Pulse width modulation for power converters:
principles and practice." New York: Chichester, 2002.
[31]
J. Holtz, "Pulsewidth modulation-a survey," in proceeding of the IEEE Power
Electronics Specialists Conference, Toledo, Spain, 1992.
[32]
J. G. Kassakian, M. F. Schlecht, and G. C. Verghese, "Principles of Power
Electronics." Reading, Massachusetts: Addison-Wesley Publishing Company,
1991.
[33]
P. Wood, "Switching power converters." New York: Van Nostrand Reinhold Co.,
1981.
[34]
G. Venkataramanan and B. K. Johnson, "Pulse width modulated series
compensator," Generation, Transmission and Distribution, IEE Proceedings-,
vol. 149, 2002.
[35]
F. Mancilla-David and G. Venkataramanan, "Generalized Averaged Dynamic
Modeling of Vector Switching AC Power Flow Controllers," in proceeding of the
36th IEEE Power Electronics Specialists Conference (PESC), Recife, Brazil,
229
2005.
[36]
A. M. Stankovic and T. Aydin, "Analysis of asymmetrical faults in power
systems using dynamic phasors," IEEE Transactions on Power Systems, vol. 15,
2000.
[37]
A. M. Stankovic, P. Mattavelli, V. Caliskan, and G. C. Verghese, "Modeling and
analysis of FACTS devices with dynamic phasors," in proceeding of the IEEE
Power Engineering Society Winter Meeting, Singapore, 2000.
[38]
A. M. Stankovic, S. R. Sanders, and T. Aydin, "Dynamic phasors in modeling and
analysis of unbalanced polyphase AC machines," IEEE Transactions on Energy
Conversion, vol. 17, 2002.
[39]
P. C. Stefanov and A. M. Stankovic, "Modeling of UPFC operation under
unbalanced conditions with dynamic phasors," IEEE Transactions on Power
Systems, vol. 17, 2002.
[40]
V. A. Caliskan, O. C. Verghese, and A. M. Stankovic, "Multifrequency averaging
of DC/DC converters," IEEE Transactions on Power Electronics, vol. 14, 1999.
[41]
B. Lehman and R. M. Bass, "Extensions of averaging theory for power electronic
systems," in proceeding of the IEEE Power Electronics Specialists Conference,
Taipei, Taiwan, 1994.
[42]
B. Lehman and R. M. Bass, "Switching frequency dependent averaged models for
PWM DC-DC converters," IEEE Transactions on Power Electronics, vol. 11,
1996.
[43]
S. R. Sanders, "On limit cycles and the describing function method in periodically
switched circuits," IEEE Transactions on Circuits and Systems I: Fundamental
Theory and Applications, vol. 40, 1993.
230
[44]
S. R. Sanders, J. M. Noworolski, X. Z. Liu, and G. C. Verghese, "Generalized
averaging method for power conversion circuits," IEEE Transactions on Power
Electronics, vol. 6, 1991.
[45]
D. W. Novotny and T. A. Lipo, "Vector control and dynamics of AC drives."
Oxford: Clarendon Press: New York, 1996.
[46]
R. C. Dorf and R. H. Bishop, "Modern control systems," 9th ed. Upper Saddle
River, NJ: Prentice Hall, 2001.
[47]
K. Ogata, "Modern control engineering," 3rd ed. Upper Saddle River, N.J.:
Prentice Hall, 1997.
[48]
J. M. Maciejowski, "Multivariable feedback design." Wokingham, England:
Reading, Mass., 1989.
[49]
M. S. El-Moursi and A. M. Sharaf, "Novel controllers for the 48-pulse VSC
STATCOM and SSSC for voltage regulation and reactive power compensation,"
IEEE Transactions on Power Systems, vol., 2005.
[50]
J. B. Ekanayake and M. Jenkins, "A three-level advanced static VAr
compensator," IEEE Transactions on Power Delivery, vol. 11, 1996.
[51]
J. B. Ekanayake and M. Jenkins, "Mathematical models of a three-level advanced
static VAr compensator," IEE Proceedings on Generation, Transmission and
Distribution, vol. 144, 1997.
[52]
J. B. Ekanayake and M. Jenkins, "Selection of passive elements for a three-level
inverter based static synchronous compensator," IEEE Transactions on Power
Delivery, vol. 14, 1999.
[53]
H. Fujita, S. Tominaga, and H. Akagi, "Analysis and design of a DC voltagecontrolled static VAr compensator using quad-series voltage-source inverters,"
231
IEEE Transactions on Industry Applications, vol. 32, 1996.
[54]
S. Bhattacharya, "Series Connected IGCT based High Power Three-Level Neutral
Point Clamped Voltage Source Inverter Pole for FACTS Application," in
proceeding of the 36th IEEE Power Electronics Specialists Conference (PESC),
Recife, Brazil, 2005.
[55]
J. Nelson, G. Venkataramanan, and B. Beihoff, "Investigation of parallel
operation of IGBTs," in proceeding of the 37th IAS Annual Meeting Industry
Applications Conference, 2002.
[56]
L. A. C. Lopes and G. Joos, "Pulse width modulated capacitor for series
compensation," IEEE Transactions on Power Electronics, vol. 16, 2001.
[57]
IEEE, "Recommended Practices and Requirements for Harmonic Control in
Electrical Power Systems, IEEE Std 519-1992," 1992.
[58]
H. K. Høidalen and R. Sporild, "Using Zigzag Transformers with Phase-shift to
reduce Harmonics in AC-DC Systems," 2005, Avialable online:
http://www.ipst.org/TechPapers/2005/IPST05_Paper044.pdf.
[59]
A. Boglietti, P. Ferraris, M. Lazzari, and F. Profumo, "Iron losses in magnetic
materials with six-step and PWM inverter supply," IEEE Transactions on
Magnetics, vol. 27, 1991.
[60]
W. A. Roshen, "A Practical, Accurate and Very General Core Loss Model for
Nonsinusoidal Waveforms," IEEE Transactions on Power Electronics, vol. 22,
2007.
[61]
H. Akagi, Y. Kanazawa, and A. Nabae, "Instantaneous reactive power
compensators comprising switching devices without energy storage components,"
IEEE Transactions on Industry Applications, vol. 20, 1984.
232
[62]
S. Bhattacharya, B. Fardenesh, and B. Sherpling, "Convertible Static
Compensator: Voltage Soucre Converter Based FACTS Application in the New
York 345 kV Transmission System," in proceeding of the 5th International Power
Electronics Conference, Niigata, Japan, 2005.
[63]
B. Sherpling, J. Sun, and S. Bhattacharya, "Power Flow Control on 345 kV Lines
with the 200 MVA Convertible Static Compensator," in proceeding of the PES
PowerTech 2005 Conference, St. Petersburg, Russia, 2005.
[64]
F. Mancilla-David and G. Venkataramanan, "Modeling and Control of the Static
Synchronous Series Compensator under Different Operating Modes," in
proceeding of the 38th IEEE Power Electronics Specialists Conference (PESC),
Orlando, Florida, 2007.
[65]
H. Fujita, Y. Watanabe, and H. Akagi, "Transient Analysis of a Unified Power
Flow Controller and its Application to Design of the DC-Link Capacitor," IEEE
Transactions on Power Electronics, vol. 16, 2001.
[66]
L. A. C. Lopes, G. Joos, and B. Ooi, "A high-power PWM quadrature booster
phase shifter based on amultimodule AC controller," IEEE Transactions on
Power Electronics, vol. 13, 1998.
[67]
V. Srinivasan, "A Real-Time Digital Controller with Re-Configurable Analog and
Logic Subsystems," M.S. Thesis, University of Wisconsin-Madison, 2003
[68]
D. Hayward, "Operating problems with parallel flows," IEEE Transactions on
Power Systems, vol. 6, 1991.
[69]
L. Naihu, X. Yan, and C. Heng, "FACTS-based power flow control in
interconnected power system," IEEE Transactions on Power Systems, vol. 15,
2000.
[70]
E. Acha, "FACTS: modelling and simulation in power networks," Wiley, 2004.
233
[71]
R. Baldick and E. Kahn, "Contract paths, phase-shifters, and efficient electricity
trade," IEEE Transactions on Power Systems, vol. 12, 1997.
[72]
B. Baran, J. Vallejos, R. Ramos, and U. Fernandez, "Multi-objective reactive
power compensation," in proceeding of the IEEE/PES Transmission and
Distribution Conference and Exposition, Atlanta, GA, USA, 2001.
[73]
D. B. Das and C. Patvardhan, "New multi-objective stochastic search technique
for economic load dispatch," IEE Proceedings-Generation, Transmission and
Distribution, vol. 145, 1998.
[74]
D. J. Gotham and G. T. Heydt, "Power flow control and power flow studies for
systems with FACTS devices," IEEE Transactions on Power Systems, vol. 13,
1998.
[75]
A. Nabavi-Niaki and M. R. Iravani, "Steady-state and dynamic models of unified
power flow controller (UPFC) for power system studies," IEEE Transactions on
Power Systems, vol. 11, 1996.
[76]
A. Schmitt and J. F. Verstege, "A multi-criteria optimization of ancillary services
in a competitive energy market," in proceeding of the IEEE Power Engineering
Society Winter Meeting, Columbus, OH, USA, 2001.
[77]
S. N. Singh and A. K. David, "A new approach for placement of FACTS devices
in open power markets," IEEE Power Engineering Review, vol. 21, 2001.
[78]
X. Ying, Y. H. Song, and Y. Z. Sun, "Power flow control approach to power
systems with embedded FACTS devices," IEEE Transactions on Power Systems,
vol. 17, 2002.
[79]
O. L. Mangasarian, "Nonlinear programming." New York: McGraw-Hill, 1969.
[80]
J. Nocedal and S. J. Wright, "Numerical optimization." New York: Springer,
234
1999.
[81]
F. D. Alvarado, C., "Computational Tools for Planning and Operation Involving
FACTS Devices," in proceeding of the Cigre Brazil - Worshop on FACTS, Rio de
Janeiro, Brazil, 1995.
[82]
M. R. Begovic, B; Lambert, F; Novosel, D, "Optimization of Capacitor
Allocation in Tranmission and Distribution Networks," in proceeding of the Bulk
Power System Dynamics and Control, Cortina d'Ampezzo, Italy, 2004.
[83]
C. Schaffner and G. Andersson, "Value of controllable devices in a liberalized
electricity market," in proceeding of the Seventh International Conference on ACDC Power Transmission, London, UK, 2001.
[84]
C. Schaffner and G. Andersson, "Determining the value of controllable devices in
a liberalized electricity market: a new approach," in proceeding of the
PowerTech, Bologna, Italy, 2003.
[85]
L. Kay-Soon and Z. Hualiang, "Robust model predictive control and observer for
direct drive applications," IEEE Transactions on Power Electronics, vol. 15,
2000.
[86]
R. Ortega, A. Loria, P. Nicklasson, and H. Sira-Ramirez, "Passivity-based
Control of Euler-Lagrange Systems: Mechanical, Electrical and
Electromechanical Applications." Berlin: Springer-Verlag, 1998.
[87]
J. M. Gonzalez and J. M. Ramirez, "AC-AC controller for steady state and
transient stability analysis," in proceeding of the 38th North American Power
Symposium (NAPS), Carbondale, IL, 2006.
[88]
J. M. Ramirez and J. M. Gonzalez, "Incorporating AC/AC Controllers into a
Power Flow Formulation," in proceeding of the Transmission & Distribution
Conference and Exposition: Latin America, Caracas, Venezuela, 2006.
235
[89]
M. Gutierrez, G. Venkataramanan, and A. Sundaram, "Solid State Flicker
Controller using a Pulse Width Modulated AC-AC converter," in proceeding of
the 2000 IEEE Industry Applications Conference, 2000.