You are on page 1of 23

DRAFT Aug 98

GUIDELINES FOR MODELING POWER ELECTRONICS


IN ELECTRIC POWER ENGINEERING APPLICATIONS
Report Prepared by the Power Electronics Task Force
of the IEEE Modeling and Analysis of System Transients Working Group
Contributing Authors: K. K. Sen (Co-Chair), L. Tang (Co-Chair), H. W. Dommel, K. G. Fehrle, A. M. Gole,
E.W. Gunther, I. Hassan, R. Iravani, A. J. F. Keri R. Lasseter, J. R. Marti, J. A. Martinez, M. F. McGranaghan,
O. B. Nayak, C. Nwankpa, P. F. Ribeiro
many power system engineers. This guideline provides general procedures to help these engineers to make their own
simulation cases as needed. The theories of power electronics are not in discussion. Attention is focused on the simulation of the interaction between the power electronic
subsystem and the connected power system. Thus, a model
for a power electronic switching device can be greatly simplified. Detailed descriptions about the device representation are presented in the later section of this paper. In the
last section of the paper, the references related to various
power electronics simulations are listed.

ABSTRACT -This paper presents a summary of


guidelines for modeling power electronics in various power
engineering applications. This document is designed for use by
power engineers who need to simulate power electronic devices
and subsystems with digital computer programs. The guideline
emphasizes the basic issues that are critical for successful modeling of power electronics devices and the interface between
power electronics and the utility or industrial system. The modeling considerations addressed in this guideline are generic for
all power electronics modeling independent of the computational tool. However, for the purposes of illustration, the simulation examples presented are based on the EMTP or EMTP
type of programs. The procedures used to implement power
electronics models in these examples are valuable for using
other digital simulation tools.

2. GENERAL CONSIDERATIONS OF POWER


ELECTRONICS MODELING
2.1. Types of Problems
Power electronics modeling can be divided into
two basic categories, depending on study objectives. The
first category covers all steady state evaluations. The focus
is from the power system response to the harmonics injected
from a power electronics subsystem. Examples of this type
include a study of the steadystate harmonic propagation in a
transmission and distribution system, harmonic frequency
resonance, system voltage and current distortion, filtering
design calculation and performance evaluation, telephone
interference analysis and system losses associated with harmonics. In this type of study, the harmonic current injection
can often be assumed independent of the voltage variations
at a point of common coupling (PCC) which is an electrically dividing point between the utility system and the customer circuit. Therefore, the power electronics subsystem
can greatly be reduced to a shunt circuit equivalent.

1. INTRODUCTION
As a consequence of the advances in power electronics technologies over the last two decades, power electronics applications have quickly spread to all voltage levels,
from EHV transmission to low voltage circuits in end user
facilities. Commonly observed power electronics applications include HVDC terminals, various static var compensation (SVC) systems, high power AC to DC converter for DC
arc furnaces, static phase shifter, isolation switch, load transfer switch, converter/inverter based drive technologies,
active line conditioning, energy storage and instantaneous
backup power systems, renewable energy integration, and
numerous others covered under subjects of Flexible AC
Transmission Systems (FACTS) and Custom Power Systems
(CPS). The need for power electronics modeling and simulation is driven by both existing and new applications.

The second type of power electronics modeling


covers a much more extensive and complex range of practical problems. In many applications, operation of a power
electronics subsystem depends closely upon the operation
state of the connected system. To evaluate the dynamic and
transient performance of a system with power electronics
interfaces, the monitoring and control loops of the system,
including detailed signal processing and power electronics
device firing need to be modeled. Examples of this type of
applications can be a SVC system, a Superconducting Magnetic Energy Storage (SMES), active power conditioning
system and various adjustable speed drives applications.

Objectives of the simulation include:


verification of an application design
prediction the performance of a system
identification of potential problems
evaluation of possible problem solutions.
The simulation is specially important for a concept validation
and design iteration during a new product development.
Power electronics applications are relatively new to

2-1

work needs to be included in the simulation or when the main


interest of the study is overall system dynamics and transients. In such cases, the EMTP type programs are usually
more suitable tools. The currently available versions of the
EMTP type programs may be less efficient for detailed
power electronics modeling. However, these programs are
very attractive for an application oriented power electronics
simulation because these programs offer tremendous capability and flexibility in characterizing various types of power
system components, including power electronics switches
with reasonably simplified characteristics. As graphic interfacing features are gradually incorporated in these programs,
a level of difficulty in using these tools will decrease.

When modeling these applications, variations of system


parameters need to be used to derive the power electronic
controls so the output of the power electronics subsystem
changes accordingly as demanded. Since a power electronics
subsystem directly affects overall system operation, a comprehensive treatment of the supply system and the power
electronics subsystem should be performed.
2.2. Frequency Domain and Time Domain Simulation
The solution methods are considered from two basic approaches:
1.
2.

frequency domain solution


time domain solution.

3. MODELING GUIDELINES

Digital computers can only simulate circuit phenomena at discrete frequencies or at discrete intervals of time
(step size (f or (t). This leads to truncation errors in all digital
simulations. Compared with the time domain calculation, a
frequency domain simulation is more robust because a circuit
solution is found at each individual frequency and truncation
errors are not accumulated. The programs using this solution
method often treat the nonlinearity of a system as known current sources. For a harmonic evaluation, the frequency
domain solution usually requires less computation time compared with a time domain solution. However, most available
frequency domain solution programs have difficulties in handling the system dynamics, control interfaces and fast transients. The time domain solution is based on the integration
over a discrete time interval. The numerical methods applied
in different programs can use either iterative techniques or
direct solution methods. The solution stability and accuracy
achieved are closely related to the time step size selection.
Because truncation errors can accumulate from step to step,
the solution m ay diverge from the true solution if an
improper time step is selected. The time domain simulation
has great advantages over a frequency domain simulation in
handling the system dynamics, power electronic interface
and transients.

3.1. Representation of Power Electronics Switching Devices


For a power level application, the commonly used
switching devices are power diode, Silicon Controlled Rectifier (SCR), Gate Turn-Off thyristor (GTO) and Insulated
Gate Bipolar Transistor (IGBT). The diode is a two-terminal, uncontrollable device and the rest are three terminal controllable devices.
Representing the reverse recovery characteristic,
leakage current, and forward voltage drop of a diode is often
not necessary for an application study. For an application
simulation, some details of a device characteristic can be
reduced. In many cases, a simplified device model is acceptable. For example, instead of trying to represent a power
electronic diode device using its switching characteristic as
shown on the left in Fig. 1, a simplified or an idealized characteristic, shown by solid and dotted lines respectively on the
right, can be used. In commonly used simulation programs,
the simplified diode model may be available as a built-in
device or it can be realized with a voltage controlled switch.
I

Simulated Device
Characteristic

Real Device
Characteristic

2.3. Tools for Power Electronics Application Simulation


V

According to their main functions, commonly used


tools for power electronics related simulations can be classified into three groups:
1. Power electronics simulation tools
2. General transients simulation tools or EMTP type of
programs
3. General harmonics simulation tools or frequency
domain simulation tools.

Cathode

Anode

Fig. 1. Actual and Simplified Switching Characteristic of Power Diode Device

If the idealized power diode model is not available,


its characteristic can be realized using a voltage controlled
unidirectional current flowing switch with a specified conducting voltage threshold, Vs, as shown in Fig. 2.

The tools of the first group may offer the best productivity if a detailed application topology and complicated
operation controls need to be simulated and if the main interest of the study is the power electronics subsystem. These
tools often have difficulties when an extensive utility net2-2

For numerical simulations, if the gating circuit power requirement is excluded from the study, there is very little difference
between modeling a GTO, IGBT or any other three-terminal,
controllable, unidirectional current flowing device. The device can all be represented by an simplified switch with gate
turn-on and turn-off controls. The different switching characteristics can be realized by applying different firing controls.

Vs
Conducting when V>Vs

In an actual power electronics application, in order to provide


a continuous current flowing path for an inductive load, a reversal diode (free wheeling diode) is used in parallel with a
controllable switching device to form the basic power electronics switching unit. However, several alternatives are
available for implementing this basic switching unit in digital
simulations.

Fig. 2. An Idealized Power Diode Realization

Since an existing diode model does not allow to


specify on-state and off-state resistances, the device can be
realized using two above described switches connected in
parallel (Fig. 3) with the opposite polarities and different
conducting voltage thresholds when loss calculation is of
interest. To include losses, the resistors Rf and Rb can be
inserted in series with the forward and backward current
flowing switches respectively, and a leakage resistor Rl can
be connected crossing the terminals of the switch branches.

Considering only the terminal equivalent at the connection


point of the application, the simplest model that can be used
to present one leg of the bridge rectifier or inverter can be constructed with one simplified bi-directional current flowing
switch with gate controls as shown in Fig. 5 (1). This model

V
Anode

Cathode

Vsf

Rf

Rb

Vsb

Gate

(1)

(2)

(3)

Fig. 5. Alternatives for Three-terminal Controllable Switching Device Representation

Rl
uses the least number of devices for a given power
electronics topology. The problem with this bidirectional current flowing switch representation is that it fails to correctly
represent an operation state during the idling period. This
can be better explained with an example of a simplified
inverter scheme used in a UPS as shown in Fig.6.

Fig. 3. Power Diode Characteristic Realization

The switching characteristics of an actual, a simplified and an idealized SCR are shown in Fig. 4. To represent
the simplified SCR device, the turn-on control is added on
the simplified diode model. If the control is applied continuously, this switch simulates the diode which allows unidirectional current flow when the switch is forward biased.
Delaying the gate pulse allows control over the turn-on
instant of the forward biased switch.

'&
%DWWHU\

AC

6XSSO\

Simulated Device
Characteristic

Real Device
Characteristic
Ig1 Ig0

Fig. 6. Simplified Inverter Scheme in a UPS System

Vbo

Cathode

When all the controlled switching devices are


blocked, the DC battery charges through the rectifier bridge
consisting of six reversal diodes. However, if the ideal bidirectional switches are used, while the UPS is idling, the battery will not be charged by the AC supply.

Anode

Fig. 4. An Actual and Simplified Switching Characteristic of SCR


Device

2-3

When the idling mode needs to be more realistically


represented, the simple bidirectional device model can be
improved with a parallel reversal diode as shown in Fig. 5
(2). When the terminal effects as well as the individual
device current carrying conditions are of interest, a further
improvement on the model can be made by adding another
diode as shown in Fig. 5 (3).

DC R and L
Voltage
Source

It should be noted that for some programs, multiple


switch connections on the same circuit node can lead to a singular system admittance matrix. When using a program with
such restriction, a small series resistor or inductor can be
used to create intermediate nodes and avoid the problem.

GTO
Load

To Firing Control
Fig. 8. An Example Circuit to Show GTO Simulation.

The GTO gating signal is illustrated in Fig. 9. When


the gating signal is greater than zero, the GTO is turned on
and when the gating signal is equal to zero, the GTO is
turned off. The GTO is initially open. The current flowing
through the GTO and voltage across the load resistor after
t=0 are illustrated in Figs. 10 and 11 respectively. By controlling the GTO firing, the average output voltage at the load
terminal can be regulated.

To illustrate a power electronic device realization,


an EMTP example of a GTO operation is given below. In
this example, a GTO module is built with a controlled bidirectional current flowing switch (type-13 switch) in series
with a built-in diode device (type-11 switch). Considering
the GTO is often utilized in applications with a reactive
power carrying capability, a reversal diode (free wheeling
diode) has been included. The schematic of the module connection is shown in Fig. 7. A list of the module file is given
in Appendix A. Note that although it has not been shown in
Fig. 7, some small resistors are used to introduce intermediate nodes between the EMTP switches and snubber elements
are used in this GTO module.

type-13
switch
Fig. 9. The GTO Firing Signal.

type-11
switch

Fig. 7. An EMTP Realization of GTO Device

An example circuit given in Fig. 8 is used to illustrate how this module is used in a circuit simulation. In this
example, the GTO device is used to regulate the voltage at a
resistive load terminal. By controlling the GTO tuning-on
and off, the average voltage across the resistor can be
adjusted.

Fig. 10. The Current Flowing through GTO

2-4

Fig. 11. The Voltage across a Resistive Load

3.2. Representation of Power Electronics (PE) Systems


If every individual power electronic switching
device is represented, a system model containing power electronic applications can easily reach a complication level that
is difficult for implementation. For an example, a HVDC terminal contains tens to hundreds series and parallel SCR
devices in one converter leg for high voltage and MVA ratings. Obviously, if one wants to represent each individual
SCR device in this HVDC system model, one will have to
build a very large model.

Fig. 12. Voltage Source Converter Simulation

With these general guidelines, a voltage source converter model for a system dynamic evaluation can be built as
shown in Fig. 12. Irrespective of how many series and parallel GTO devices are used in an actual application, only two
GTO devices are used in each phase of the model to form a
converter leg. In this example, the just discussed GTO module is used as a building block to construct the converter
module.

Fortunately, (except for some failure mode analyses), for the purposes of most application simulations it is not
necessary to represent all individual devices. Usually, what
needs to be simulated is the terminal characteristics of a
power electronic subsystem and how it interfaces with the
connected system. Thus, the following procedures can be
used to reduce the modeling complexity:
Use one or a few equivalent devices to represent series
and parallel combination of a group of devices
Represent power electronic loads with similar characteristics by an equivalent load
Use the simplest device model which is appropriate for
the application
Represent a power electronic subsystem by equivalent
source injection whenever it is acceptable
Represent only the front end of the drive system when
the major concern is utility interfacing
Include the system dynamic and controls only when necessary
Use modular approach for large scale model development

The GTO firing control signals for the six-step operation is given in Fig. 13. The firing signal for the switching
cells No. 1 through No. 6 are shown in this figure by the
traces with the magnitude 1.1 through 1.6 respectively. The
device firing starts when the time reaches one 60 Hz cycle
ending point. When the gating signal is greater than zero, the
corresponding GTO device is turned on. When the gating
signal is equal to zero, the device is turned off. The resulting
line-to-neutral and line-to-line output voltages are illustrated
in Figs. 14 and 15, respectively.

Fig. 13. The GTO Firing Signal for the Six Step Converter Operation

2-5

Fig. 18. The Resulting Line-to-Neutral Voltage


Fig. 14. Six Step Operation Line-to-Neutral Voltage

Fig. 19. The Resulting Line-to-Line Voltage


Fig. 15. Six Step Operation Line-to-Line Voltage

In this guideline, detailed considerations related to


applying these system reduction are presented. Some important issues addressed are:

For this case, the GTO firing control signal is


derived by comparing the saw tooth carrier signal with a reference signal of a desired frequency of that phase. As shown
in Fig. 16, whenever, the instantaneous value of the reference
signal is greater than that of the carrier signal, the device is
turned on.

Harmonic cancellation when multiple loads are represented by their lumped equivalent
Existing system distortion.
Appropriate source topology for PE subsystem representation
System unbalance
Effects of a DC link or the inverter side connection on
the front end interface with the power supply system
Current or voltage sharing among the parallel or series
switching devices
Switching loss prediction

3.3.Representation of Power Systems


Similar to the situation in a power electronics subsystem, a power supply system can easily extend to a large
electrical and geographic radius and become too complicated
to model. Therefore, the power system needs to be simplified. The proper level of system reduction depends on the
study objectives.

Fig. 17. The PWM Firing Signal: Reference Voltage Frequency = 60


Hz, Carrier Signal Frequency = 600 Hz

The simulated line-to-neutral and line-to-line output


voltages are shown in Figs. 17 and 18, respectively. By
changing the magnitude and frequency of the reference signal, the output voltage magnitude and frequency can be
changed. The carrier frequency determines the characteristic
harmonic components of the output voltage.

If the purpose is to characterize the harmonics generated by a particular type of power electronics application,
the power system model can be significantly reduced. When
a pre-existing voltage distortion level at a power electronics
interfacing bus is low, the rest of the power system can be
2-6

satisfactorily represented by one or a set of first order equivalents connected to the bus at a higher system voltage level.
For an example, if the power electronics application interfaces with the system at the low voltage bus of a step-down
transformer, the equivalent of the system can be placed on
the high voltage bus of the transformer. When a pre-existing
voltage distortion level is greater than 2%, one needs an adequate harmonic source to properly represent the background
distortion.

If the objective is to evaluate effects of the power


electronics on a connected utility system, the model shall be
extended to cover all sensitive loads (i.e. rotating machines
and all other major power electronics) within a concerned
electrical radius. Special attention is needed if an unbalanced system condition is involved.

Extensive power system model is required for a harmonic propagation and resonant study. The main system
components and dominant topology need to be kept in the
power system model. Filter banks, nonlinear passive circuit
components, and all other harmonic injection sources should
be represented. Frequency dependent characteristics of the
system components might need to be considered.
3.4. Representation of System Controls

The system control is one of the most important


aspect of a power electronics simulation. As illustrated in
this paper, a switching device is greatly simplified. The
proper switching performance of a device is realized via
appropriate gate controls. Modeling of power electronic controls consists of three steps:
1. Monitoring and sampling
2. Signal processing and control reference derivation
3. Device gating signal generation.
Most simulation tools provide some means to
implement system controls. In some later developed programs, the control block diagram and flow-chart structures
are supported for modeling different levels of system controls. Using these tools, a user can define the specified controls in a simulated system with great flexibility. Some key
issues ensuring a correct control modeling is briefly mentioned below. These issues are more thoroughly treated in
the guideline with illustration examples.
For a time domain simulation, the highest resolution for
a signal sampling is determined by a selected time step.
In general, this presents no problem for analog control.
However, for digital control simulation, if the selected
time step is too large and if the simulated sampling resolution is significantly different from the real system sampling resolution, significant errors can be introduced and
even lead to instability.
For a time domain simulation, the computation time
does not reflect the simulated control logic response
time. User should always remember to introduce a rea-

sonable time delay to match with the limitations of the


control hardware and software.
When modeling a control response, it is important to
understand the program introduced time delay between
the primary system and the control interface. For an
example, the control model may introduce one or more
time step delay because of structure and solution method
of the program. This may not cause problems in some
simulations. However, if the modeled control logic
makes this time delay caused error to accumulate over a
period of time, it can eventually result in the solution
divergence. The problem can be corrected in most cases
by reducing the size of the time step or avoiding the possible accumulation mechanism in the control model.
Different methods may be used to synchronize power
electronics gating signals with required system references. In many cases, a real phase-locked-loop (PLL)
can be greatly simplified to reduce the modeling system
complexity. However, when the system contains significant waveform distortions, either harmonics or transient
disturbances, a practical PLL with all signal filters
should be carefully implemented in the control model in
order to accurately predict control response. This is particularly important when the objective of the simulation
is to verify control design and to evaluate the response of
a power electronic application to primary system dynamics.
All power electronic devices have their limit in switching frequency. When a load commutation or a standard
PWM type scheme is simulated, the highest switching
frequency in the simulation is controlled by the system
frequency or by a carrier frequency. Even considering a
variable carrier frequency, the number of switching per
fundamental frequency cycle is known and the highest
switching frequency can be made under a physical limit
of the simulated device. However, if the device firing is
determined by a simple comparison between the system
control reference and the system output, a device switching may take place in simulation whenever a comparison
difference is detected. Therefore, the switching frequency becomes highly dependent on the time step size,
and the average switching frequency becomes unpredictable. When using this type of firing logic, user should
always take extra measures, such as introducing a hysteresis loop, to ensure that the modeled device is working under its physical switching capability.

The block diagram for a commonly used phase locked loop is


shown in Fig. 20. The system voltage at the terminal of the
load is sampled and used to generate the synchronous reference signal. From Fig. 21 through Fig. 26, the input, intermediate output quantities and output signal are plotted to show
the each step of this signal processing.

2-7

Uref
Vsyn

Va

Vd

ABC

COS()

Vb
VCO

PI
Vc

Vq
DQ

Uref

SIN()

RESET

Fig. 20 The Block Diagram of a PLL


Fig. 24 Derived Frequency Reference

Fig. 21 Input three-phase Voltage Signal

Fig. 25 Output of VCO

Fig. 22 DQ Voltage Components

Fig. 26 Comparison of Input and Obtained Syn. Signals

The response of this control circuitry to a system


disturbance is illustrated in Fig. 27. A balanced system fault
is placed on and removed from the system, resulting a three
cycle voltage sag.
Three phase fault
Rf=0.01 ohms
Tin=0.025s
Tcl=0.075s
Kpro = 100
Kint = 8.3E-3
Fig. 23 Derived Phase Error

With above given control parameters, it takes the


PLL about three cycles to be relocked into the system voltage.

2-8

oscillations in which case it is not a concern to the simulation


engineer. Otherwise, some of the measures listed later in this
section may have to be implemented to obtain correct results.
The finite nature of the simulation time step that the
EMTP type programs use also poses another problem for
power electronic circuit simulation which necessitates the
use of snubber circuits across fast acting power electronic
switches. Note that in some situations the snubber R and C
values of the actual system may or may not work in simulations using some programs. In this case, the R and C values
of the snubbers needed for stable simulation is primarily
dependent on the time step and secondarily on system configuration (capacitors and inductors in the system) and the load
current level. Programs using special features such as variable time steps (very short time steps during switching) or
interpolate switching (simulate the switching very close to
the required instant using linear interpolation between time
steps) do not require fictitious snubber circuits. Therefore,
one of the following measures or their combinations can be
taken to prevent numerical instability in the simulation:
Select a smaller time step
Use artificial snubber circuits
Introduce a small smoothing reactor for DC links
Introduce proper stray capacitances in the system model
Provide a parallel damping for lumped system.

Fig. 27 The PLL Circuitry Response to a System Disturbance

If the circuit parameters are changed from the above


given values to the values listed below, for the same system
and fault, one can observe that the same PLL logic can be
relocked into the system voltage within a half cycle period of
the time as shown in Fig. 28.
Three phase fault
Rf=0.01 ohms
Tin=0.025s
Tcl=0.075s
With Modified:
Kpro = 1000
Kint = 8.3E-4

A model for a general, unidirectional conducting,


three terminal, controllable power electronic device with
snubber connections is shown in Fig. 29. An actual snubber
configuration can be different from one application to
another. However, if the purpose of a simulation is not to
design the snubber, a sample snubber configuration shown in
this figure can often provide satisfactory results.

ATRMNL

GTRMNL

SNUBBR

Fig. 28 The PLL Circuitry Response to a System Disturbance


SNUBBC

Note that this example of PLL logic based on the


three phase to DQ transformation is valid for three-phase balanced application. Also, its performance characteristic is
highly affected by the parameter settings.

SNUBBR
SNUBBC

3.5. Snubber Treatment in EMTP type PE Modeling


CTRMNL

The simulation programs using trapezoidal integration method are inherently prone to spurious oscillations
(also known as chatter) in capacitive and inductive circuits
when subjected to sudden changes such as step change in
voltage, current injection and switching. Some EMTP type
programs take special measures to detect and remove these

Fig. 29 A Sample Snubber Circuit

2-9

3.6.Simulation Errors and Control


AC Supply
System

Errors in a Power Electronics simulation can come


through the following sources:
1. switching device approximation and system reduction
2. added circuit elements for numerical oscillation control
3. control system simplification
4. time step related truncation
5. program structure and solution method introduced interfacing time delay
6. incorrect system initial conditions
For application simulations, some errors resulting
from the system simplification and measures of numerical
oscillation control are acceptable. The fourth and fifth items
in the list can be controlled by reducing the time step size. A
recommended time step size should not be greater than 1/5 to
1/20 of the period of the highest concerned frequency
cycle. For an example, for an IGBT inverter simulation with
5000Hz PWM switching, a selected time step could be 10
ms. However, if the objective of the simulation is to see the
detailed transient at the terminal of the induction motor
which is fed by the inverter through a section of the cable
with an 1.0 ms travel time, an adequate time step should be
0.2 ms or smaller.

DC Link

Induction
Motor
IM

Six Pulse
Diode Rectifier

PWM
Inverter

Fig. 30. Electrical Circuit Configuration of an Adjustable Speed


Drive

The built-in diode models are used to construct the


front end rectifier. The same switching devices with added
open/close controls are used to represent output inverter
IGBTs. The EMTP input data modules are use to build this
example case. Both the output reference frequency and the
PWM carrier frequency are made to be controllable. Modeling of a signal processing and firing pulse generation is illustrated in this example. The motor load of the drive is
represented by its R+jX equivalent branch. The simulated
AC input current, carrier and reference signal for the PWM
control, AC output voltage and current are presented in Fig.
31 through Fig. 33.

Errors caused by incorrect system initial conditions


can be reduced by just letting the simulation run for a period
of time to reach a corrected initial condition. This may take
more computing time, but time is saved in model construction, especially if the program allows to restart. There are
some methods developed which help to accelerate the system
into the correct initial condition quickly.
4. SUMMARY OF DOCUMENTED EXAMPLES
Summary of simulation examples on PWM Voltage
Source Inverter Adjustable Speed Drive, Voltage notching
caused by operation of Current Source Inverter, HVDC terminal and shunt TSC/TCR compensation, modeling of rotating machines and a comprehensive treatment on the voltage
source inverter based FACTS devices and its modeling techniques using EMTP are presented in this section. The examples can be either a real case study or an exercise for
illustration purposes.

Fig. 31. Simulated AC Input Current of A PWM-VSI Adjustable


Speed Drive

4.1. Simulation of the Pulse Width Modulation (PWM) Voltage Source Inverter (VSI) Adjustable Speed Drive (ASD)
The first example is a PWM-VSI AC drive simulation using EMTP. The AC drive (Fig. 30) consisting of a
three-phase diode bridge rectifier, capacitive DC link and
three-phase PWM output inverter. The switching losses of
the drive are a secondary order consideration in the analysis
and the idealized switching characteristics are used.
Fig. 32. Simulated Carrier and Reference Signals for A PWM- VSI
Adjustable Speed Drive Firing Control

2-10

Short Circuit Level =


74 MVA

144 kV
10 MVA
7.87%

25 kV Bus

2 km 266 MCM
7.5 MVA
5.75%

Fig. 33. Simulated AC Output Line-to-line Voltage of A PWM-VSI


Adjustable Speed Drive

ASD
harmonic
filters
5,7,11

4.2.Simulation of Voltage Notching Caused by Operation of


Current Source Inverter (CSI) ASD
The second example is based on a case study. The
involved system is illustrated by the one line diagram in Fig.
34. The 35 kV distribution system is supplied through a 10
MVA transformer from the 144 kV transmission system. The
customer causing the voltage notching problem has a 6000
hp induction motor supplied through a CSI adjustable speed
drive[53]. This drive is at a 4.16 kV bus supplied through a
7.5 MVA transformer. Harmonic filters (5th, 7th, 11th) are
included to control the lower order characteristic harmonics
of the six pulse drive. The actual adjustable speed drive and
motor load were represented to reproduce the notching oscillations observed in the measurements.

6000 hp

other
loads
1500 kVA
4.7%

1500 kVA
4.7%

4.16 kV

480 Volt bus


Motor Load
(650 hp)

0.5 uF
surge capacitors

PF Correction
Capacitors
(160 kvar)

800 hp

Operation of the 6000 hp motor and drive resulted


in significant oscillations on the 25 kV supply system. These
oscillations caused clocks to run fast at the customer with the
6000 hp motor (clocks were fed separately from the 25 kV
system) and failure of surge capacitors on the 800 hp motor
at the customer located on the parallel feeder. The objective
of the simulation is to identify the power quality problem
associated with this 6000 HP CSI ASD operation. The simulation was carried out using EMTP.
The worst notching problems are associated with a
firing angle at about 70% load. The simulated waveform for
the 25 kV bus voltage is shown in Fig. 35. The oscillations at
each commutation point are in good agreement with the measurement results.

other
loads

4.16 kV
10 km
1/0

Fig. 34. One line diagram for the first example system.

Fig. 35. Simulated 25 kV system voltage with drive operating.

Fig. 36 illustrates the voltage waveform at the 4.16


kV bus where the 800 hp motor surge capacitors cause magnification of the oscillations. The potential for problems at
this location is quite evident.

2-11

nous condenser (SC) and half-and-half mix of the two


(SVC+SC).
Fig. 38 shows the typical results obtained from the
simulation for a permanent DC block. The fixed compensator case does not control the overvoltage; however, all other
options do. The SVC option is the fastest to respond followed by the SVC+SC option and lastly the SC option. This
simulation setup can be used to conduct almost any type of
performance study including a thyristor miss-fire in HVDC
valve group or in the SVC itself.
Fig. 36. Simulated waveform at surge capacitor location (4.16 kV
bus of customer on parallel feeder)

4.3. Simulation of HVDC Terminal and Shunt TSC/TCR


Compensation
The third example is an illustration case for modeling of an HVDC system with shunt TSC/TCR compensation
at the inverter bus [54]. The simulation example is made
using PSCAD/EMTDC. The schematic system shown in
Fig. 37 is a modified version of the GIGRE Benchmark
Model for HVDC Control Studies [55].
The inverter short circuit ratio has been reduced
from its original value of 2.5 to 1.5 to make the study more
interesting. The DC link is a 1000MW, 500kV, 12 pulse
monopolar system. There are damped low and high pass filters at each converter terminal to reduce the distortion on the
AC bus. The control scheme for the HVDC system consists
of a rectifier current controller with the gamma controller.

(Please ask Dr. Gole for picture)


Fig. 38 Inverter AC Voltage Following a Permanent DC Block

4.4. Modeling of Rotating Machines


1000 MW
500kV
12 Pulse

Rectifier AC
System

Two possible situations can be considered for modeling a rotating machine when simulating a power electronic
system
1. The machine is a component of a larger system where
one or several power electronic devices are operating,
for instance a synchronous machine connected to a
transmission network where FACTS devices are used to
control power flows and improve transient stability.
2. The machine is part of the power electronic system, for
instance an adjustable speed drive.

Inverter AC
System

12
12

Filters
FC
Rectifier

Inverter

Filters
FC,SC
SVC

Fig. 37. Study System

The SVC system is a -200/+300 MVAr, 12 pulse,


TCR and TSC (two stage) combination connected to the
inverter bus through a step up transformer. The SVC controls are designed to coordinate the control of TCR and TSC
in such a way that the combined susceptance of the SVC is
continuous over its entire operating range. The basic control
mode is voltage control and has as a voltage droop built into
the controls. Several studies to evaluate the recovery to full
power after a contingency were simulated [54]. The performances of several compensation options were compared.
These options included fixed capacitors (FC), SVC, synchro-

Similar modeling guidelines for representing rotating machine in both situations can be used, however some
particular considerations can be taken into account in some
cases and studies. Modeling guidelines provided in this document assume that power electronic systems operate at low
frequencies, between DC and 3 kHz. Therefore only the representation of rotating machines for this frequency range is
discussed. Regardless of the application to be simulated a
detailed modeling for the electrical and the mechanical parts
is usually required, saturation effects should be included, and

2-12

capacitance effects can be neglected. Frequency dependent


of electrical parameters, mainly rotor parameters might also
be considered. If frequencies of the transient case to be simulated are higher than 3 kHz, the simulation time is no longer
than a few milliseconds and the machine is not close to any
power electronic system (i.e. a synchronous machine connected to a transmission system), the mechanical part can be
usually neglected and the machine can be represented by
means of an ideal source behind its subtransient reactance
and a frequency-dependent resistor; capacitance effects
become important, they can be represented by a parallel
capacitance-to-ground [1].
Other representations should be considered for specific applications
A) An agg reg ated r epr ese ntatio n o f sever al
machines can be used to reduce complexity and simulation
time. Ref. [2] resents an aggregated induction model to be
used in power quality studies where short term interruptions
(i.e. sags) are of concern.
B) A very simplified representation for synchronous
and induction machines can be used in harmonic studies
when the machine is not directly connected to the harmonic
source [3].
An important aspect of the simulation of a rotating
machine is the procedure to obtain machine parameters and
the information where these parameters are derived. Electrical parameters of synchronous machines may usually be
obtained in one of the following forms : (1) data supplied
from manufacturer (conventional stability format data, standstill frequency response), (2) data from field tests (on-line
frequency response, load rejection test, other tests) and (3)
computer calculation using the finite-element method [4]. A
good discussion about methods to obtain electrical and also
mechanical parameters can be found in [5]. Data from steady
state and short circuit tests include reactances and time constants, armature resistance as well as saturation effects. Several procedures have been proposed to pass from these data
to electrical parameters which are used in the transient solution of the machine [6-8]. Although these tests and the corresponding procedures can also be used to obtain electrical
parameters of an induction machine, data conversion procedures for this type of machines are performed from different
specifications [9-10].
Frequency response tests have received much attention during the last 25 years. Several methods have been proposed to obtain parameters of d- and q-axis equivalent
circuits; they are based on standstill frequency response
(SSFR) [11-14], and on-line frequency response [15-16].
Some techniques have also been developed to account for
saturation effects [17].
4.5. Voltage Source Inverter Based FACTS Devices and
theirModeling Techniques Using EMTP

This section describes the fundamentals and the


modeling techniques of Voltage Source Inverter-based Flexible Alternating Current Transmission Systems (FACTS)
devices, namely, STATic synchronous COMpensator (STATCOM), Static Synchronous Series Compensator (SSSC), and
Unified Power Flow Controller (UPFC) using an EMTP
simulation package. The FACTS model includes all the necessary components - a voltage source inverter with a DC link
capacitor, a magnetic circuit, and a realizable controller. The
UPFC model consists of two solid-state voltage source
inverters which are connected through a common DC link
capacitor. Each inverter is coupled with a transformer at its
output. The first voltage source inverter, known as STATCOM, injects an almost sinusoidal current, of variable magnitude, at the point of connection. The second voltage source
inverter, known as SSSC, injects an almost sinusoidal voltage, of variable magnitude, in series with the transmission
line. When the STATCOM and the SSSC operate as standalone devices, they exchange almost exclusively reactive
power at their terminals. While operating both the inverters
together as a UPFC, the injected voltage in series with the
transmission line can be at any angle with respect to the line
current. The exchanged real power at the terminals of one
inverter with the line flows to the terminals of the other
inverter through the common DC link capacitor. The functionalities of the models have been verified.
4.5.1 VSI Based Facts Devices
Flexible Alternating Current Transmission Systems
(FACTS) devices, namely STATic synchronous COMpensator (STATCOM), Static Synchronous Series Compensator
(SSSC) and Unified Power Flow Controller (UPFC), are
used to control the power flow through an electrical transmission line connecting various generators and loads at its sending and receiving ends. Each of the STATCOM and the
SSSC consists of a solid-state voltage source inverter with
several Gate Turn Off (GTO) thyristor switch-based valves
and a DC link capacitor, a magnetic circuit, and a controller.
The number of valves and the various configurations of the
magnetic circuit depend on the desired quality of AC waveforms generated by the FACTS devices. When the STATCOM and the SSSC operate as stand-alone devices, they
exchange almost exclusively reactive power at their terminals. While operating both the inverters together as a UPFC,
the exchanged power at the terminals of each inverter can be
reactive as well as real. The exchanged line flows to the terminals of the other inverter through the common DC link
capacitor. The objective in this section is to describe each
component, such as a voltage source inverter, a magnetic circuit, and a controller of FACTS devices and its modeling
techniques using an EMTP simulation package. Since, the
emphasis of modeling is purely on FACTS devices, the
power system in which the FACTS devices are connected to
has been modeled in a simplistic way. A simple transmission
line, shown in Fig. 39, has an inductive reactance, Xs, and a
voltage source, , at the sending end and an inductive reac-

2-13

tance, Xr, and a voltage source, , at the receiving end, respectively. The STATCOM is connected at BUS 1 of the
transmission line as shown in Fig. 39. The STATCOM
model in EMTP consists of a harmonic neutralized voltage
source inverter, VSI1, a magnetic circuit, MC1, a coupling
transformer, T1, a mechanical switch, MS1, current and voltage sensors, and a controller. The STATCOM injects an
almost sinusoidal current at the point of connection. This
injected current is almost in quadrature with the line voltage,
thereby emulating an inductive reactance or a capacitive
reactance at the point of connection. To achieve the basic
function of a STATCOM, the inverter is operated by regulating the reactive current flow through it.

9V

(

%86

9

06

;U

9U

,19
,
(

(

L '&

06

,

'&

(

9

'&

L
06

&RQWUROOHU
0&

96,

;U

9

9U

,
(6

(

96,

67$7&20

9

,19

,19

3U  4 U
,

7

;V

9V

7

0&

06

9GT

9

7

%86

83)&

3U  4 U

%86 
;V

The UPFC which is connected to the simple transmission line is shown in Fig. 41. The UPFC model in EMTP
consists of two harmonic neutralized voltage source inverters, VSI1 and VSI2, two magnetic circuits, MC1 and MC2,
two coupling transformers, T1 and T2, four mechanical
switches, MS1, MS2, MS3, and MS4, two electronic switches,
ES2 and ES22, current and voltage sensors, and a controller.
The voltage source inverters are connected through a common DC link capacitor. In a basic operation of a UPFC, the
STATCOM is operated by regulating the reactive current
flow through it and the SSSC is operated by injecting a voltage in series with the transmission line.

(6
'&

(

'&

06

96,

0&

&RQWUROOHU

&RQWUROOHU

Fig. 39 A Static Synchronous Compensator Model in EMTP

Fig. 40 shows an SSSC connected in series with the


simple transmission line between BUS 1 and BUS 2. The
SSSC model in EMTP consists of a harmonic neutralized
voltage source inverter, VSI2, a magnetic circuit, MC2, a
coupling transformer, T2, a mechanical switch, MS2, two
electronic switches, ES2 and ES22, current and voltage sensors, and a controller. The SSSC injects an almost sinusoidal
voltage, of variable magnitude, in series with the transmission line. This injected voltage is almost in quadrature with
the line current, thereby emulating an inductive reactance or
a capacitive reactance in series with the transmission line.

%86

%86

666&
06

9T

9V

3T 4 T

9

,19
9

Fig. 42 shows a single phase inverter circuit,


referred to as a 3-level pole, which consists of a positive
valve, A+, a negative valve, A-, and an AC valve, A AC.
When a pole is connected across a series of capacitors which
are charged with a total DC voltage of vDC and the valves are
closed and opened alternately, the pole output voltage, vAO,
at the midpoint of the pole A with respect to the midpoint, O,
of the capacitor is a quasi square wave containing a positive
sequence fundamental component and all the odd harmonic
components, such as the zero sequence third, the negative
sequence fifth, and the positive sequence seventh, etc.

9

;U

(6

'&

Y'&

'&

96,

$

Y'&

$ $&
$

2))
21

2))

$ $& 21

(6

(
Y

21

$
/HYHO ,QYHUWHU 3ROH $

9U

,
( 

4.5.2 DESCRIPTION OF THE INVERTER

7

;V

Fig. 41 A Unified Power Flow Controller Model in EMTP

2))

21

YGF
Y$2

Y$

2))

21

+  

YGF

0&

&RQWUROOHU

Fig. 42 A 3-Level Inverter Pole and its Output Voltage


Fig. 40 A Static Synchronous Series Compensator Model in EMTP

2-14

The amplitude of any odd multiple of fundamental compo-

an inverter pole voltage is time shifted by an angle of -, the


fundamental and all the harmonic components of the pole
voltage get a phase shift by an angle of + in the positive
direction, irrespective of their sequence.

nent is
VA O, n =

2
n

(1)

v DC cos n

where is the dead period during which the AC


valve operates in each quarter cycle and the pole output voltage is zero and n = 2k + 1 for k = 0, 1, 2, 3, etc. For ( = 0, the
fundamental as well as all the harmonic components have the
highest possible amplitudes.

,QYHUWHU '()



9 )


'

9 '

9 (

Y '&
Y '&
Y '&

&

Y '&

Y $

Y '&

Y $1

Y %1

Y &1

Y '&

Y 1


'&

<

9 &

&

9$

Y &1

,QYHUWHU $%&

9 %

Fig. 44 A 12-Pulse Harmonic Neutralized Inverter Configuration


with 3-Level poles

Fig. 43 shows three poles A, B, and C which are


connected across the same DC link capacitor and the pole
outputs are connected to a 3-phase load whose neutral point,
N, is not connected to the midpoint of the capacitor. The
poles A, B, and C which form a 6-pulse inverter are operated
in such a way that the pole voltages, vAO, vBO, and vCO, are
time shifted from one another by one third of the time period
of the pole voltage. Therefore, the fundamental phasors
and

<

&,5&8,7

Y %1

Fig. 43 A 6-Pulse Inverter with 3-Level poles and its Output Voltages

V A, 1, V B, 1

0$*1(7,&

2
9

9 '&

Y &

Y '&
Y 1

Y $1

Y %

Y '&

Y
;

V C, 1 are 120 o apartS ince, the zer o

sequence components of each pole current have no return


path to the midpoint of the DC link capacitor, the zero
sequence components of each pole voltage, vNO = (vAO + vBO
+ vCO)/3, appear between the neutral point and the midpoint
of the DC link capacitor. Therefore, each phase of the load
voltages, vAN = vAO - vNO, vBN = vBO - vNO, and vCN = vCO vNO, consists of only a fundamental component and odd harmonic components (n) given by the equation (1) where n = 6k
1 for k = 1, 2, 3, etc.
Fig. 44 shows two 6-pulse inverters (ABC and DEF)
which are operated from the same DC link capacitor. On the
AC side, they are connected to a 3-phase load (XYZ) through
a magnetic circuit. The poles D, E, and F are operated in
such a w ay th at t he po le v ol tag e f un dam e nta l p hasors V E, 1, V C, 1
and
V F, 1 and are 120 o apart and the
fundamental voltage phasor set of the DEF inverter lags the
fundamental voltage phasor set of the ABC inverter by 30o.
The displacement angle between two consecutive 6-pulse
inverters in a multipulse inverter arrangement is 2/6m,
where m is the total number of 6-pulse inverters used. The
configuration of the magnetic circuit in Fig. 44 is such that if

Table 1 shows the time shifted A and D pole voltages first twenty five harmonic components final phase
angles after appropriate phase shift. The pole voltages from
the ABC inverter exhibits a 6-pulse harmonic neutralized
waveform with harmonic components n = 6k 1 for k = 1, 2,
3, etc. Similarly, the pole voltages from the DEF inverter
exhibits a 6-pulse harmonic neutralized waveform whose
harmonic components (n = 6k 1 for k = 1, 2, 3, etc.) have
the same magnitudes as the corresponding harmonic components of the ABC inverters 6-pulse harmonic neutralized
waveform. However, the harmonic components (n = 6k 1
for k = 1, 3, 5, etc.) are in opposite phases while the harmonic
components (n = 6k 1 for k = 2, 4, 6, etc.) are in phases with
the corresponding harmonic components of the ABC
inverters 6-pulse harmonic neutralized waveform. Therefore, if all the outputs from each 6-pulse inverter are combined by connecting the corresponding phases in series, a 12pulse harmonic neutralized waveform is obtained. The
resulting output voltage exhibits a fundamental component
and odd harmonic components (n) given by the equation (1)
where n = 12k 1 for k = 1, 2, 3, etc. Note that the output
voltage of a 12-pulse inverter with 3-level poles is referred to
as a 12-pulse waveform when the poles are operated with
dead angle = 0.
Fig. 45 shows a possible configuration of the magnetic circuit which can be used to generate a 12-pulse harmonic neutralized voltage. The ABC 6-pulse inverter
voltage is fed to a Y-Y transformer and the DEF 6-pulse
inverter voltage is fed to a (-Y transformer. The inverter side
A winding and DE winding will have per turn fundamental
component voltages which are of same magnitude and in
phase and the fifth and the seventh harmonic components
each of which are of same magnitude but in opposite phase.
Therefore, if the line side of the transformer windings are
connected in series, the phase-X voltage will exhibit only a

2-15

fundamental component and 12-pulse harmonic components.


Note that the inverter side ( winding has 3 times the turns
as the inverter side Y winding has. This is needed in order to
keep the same volts per turn in both windings. The line side
inverter windings can have any turns ratio other than 0.5 to
increase or decrease the output voltage.
n

time shift

5
7
11
13
17
19
23
25

pole A
-5*(0)
+7*(0)
-11*(0)
+13*(0)
-17*(0)
+19*(0)
-23*(0)
+25*(0)

phase
shift

final
phase
angle

0
0
0
0
0
0
0
0

time shift

pole D
-5*(-/6)
+7*(-/6)
-11*(-/6)
+13*(-/6)
-17*(-/6)
+19*(-/6)
-23*(-/6)
+25*(-/6)

0
0
0
0
0
0
0
0

phase
shift

final
phase
angle

+/6
+/6
+/6
+/6
+/6
+/6
+/6
+/6

0
0

0
0

tralized inverter voltages, e 1 , which are fed to the source


section. In an actual simulation case, the gating signals are
used to operate the pole valves of an inverter structure such
as one shown in Fig. 42.

*HQHUDO &RQVWDQWV

6HQVLQJ
9ROWDJHV
&XUUHQWV

&RQWURO  7$&6
0HDVXULQJ
6ZLWFKHV

,QYHUWHU
9ROWDJHV

%UDQFK
7UDQVPLVVLRQ /LQH
7UDQVIRUPHU

6RXUFHV
7$& &RQWUROOHG
,QGHSHQGHQW

Table 1 Phase Angles of a 12-Pulse Inverter Phasors

2XWSXWV
0$*1(7,& &,5&8,7
&

)

(
Y)


&

Fig. 46 EMTP Modeling Structure

)




9'



'

Y;

Y$

&

Y<

=


Y=







'

Y(

9 $







<

Fig. 45 A Magnetic Circuit for a 12-Pulse Harmonic Neutralized Inverter

The 12-pulse inverter configuration, shown in Fig.


45, presents a 3-phase voltage which contains a fundamental
component and odd harmonic components (n) where n = 12k
1 for k = 1, 2, 3, etc. The presence of 12-pulse harmonic
components in the inverter output voltage may not be acceptable in many applications. Therefore, an inverter with a
higher pulse output voltage should be considered [56-58].

4.5.3 MODELING TECHNIQUE


ig. 46 shows the block diagram of the EMTP simulation program layout. Sample EMTP program files are
given in [56-58]. First, some general constants are defined.
Next, the control or the Transient Analysis of Control Systems (TACS) section receives its input signals from the sensors or measuring switches. This section generates the gating
signals for the pole valves on the fly. The ideal pole voltages are mathematically combined to produce harmonic neu-

Each valve, located in the switch section, can be


modeled with a number of GTO thyristors connected in
series each having an antiparallel diode and appropriate
snubber circuits. The pole output voltages are fed to a magnetic circuit, located in the branch section, which produces a
3-phase voltage set. In this way, the effects of a nonideal
magnetic circuit, which includes leakage reactance, magnetic
saturation, etc. can be studied. However, in this paper, the
valves and the magnetic circuit are assumed to be ideal. The
voltage, vDC, across the DC link capacitor is maintained by
the power balance equation at both AC and DC sides of the
inverter. This modeling technique gives more insight to the
operation of the power circuit which produces a 3-phase voltage set. The source section has some independent voltage
sources which establish the power flow in a transmission
line. Next, the controlled and the independent sources are
fed to the branch section which contains the transmission line
and the coupling transformer. The line voltage set, v 1 , at
BUS 1, the inverters current sets, i1 and i2, and the line current, i, are measured by the measuring switches. Finally, the
output section is defined. In reality, the magnetic circuit can
also serve as the coupling transformer. Therefore, there is no
need for an additional coupling transformer.
The modeling may be done at various levels. For
example, to study the functionality of a FACTS device on an
elaborated power system network, a FACTS device with a
simplified model consisting of sinusoidal voltage sources
and detailed control and protection schemes may be adequate. For magnetic circuit and valve designers, the primary
focus should be on the modeling of the detailed power circuit. The modeling techniques described in this section are

2-16

useful tools to the FACTS designers.


The various control techniques of FACTS devices
and simulation results are described in the next section. In
each case, an instantaneous 3-phase set of line voltages, v1, at
BUS 1 is used to calculate the reference angle, which is
phase-locked to the phase a of the line voltage, v1a.
A. STATCOM
The controller of a STATCOM is used to operate the
inverter in such a way that the phase angle between the
inverter voltage and the line voltage is dynamically adjusted
so that the STATCOM generates or absorbs desired VAR at
the point of connection [56]. Fig. 47 shows the control block
diagram of the STATCOM. An instantaneous 3-phase

reactive current control operation of a STATCOM. Between


0 and 50 ms, the mechanical switch, MS1, stays open, disconnecting the STATCOM from the transmission line. The DC
link capacitor is precharged. The inverter output 12-pulse
voltage of phase a, e1a, and the line voltage of phase a, v1a,
are in phase. At 50 ms, MS1 closes and the quadrature current demand, I 1q * , of the inverter is set to zero. Since the
inverter current is zero, the inverter voltage of phase a, e1a,
and the line voltage of phase a, v1a, have equal amplitudes.
At 125 ms, the quadrature current demand, I 1q * , of the
inverter is set to one per unit capacitive, which means the
STATCOM should see the system as an inductive reactance and the inverter current of phase a, i1a, lags the line

voltage of phase a, v1a, by almost 90 .

&RQWUROOHU

9 $


, T
, G

L

Y '&

, T

5RWDWLQJ
)UDPH
7UDQVIRUPHU

Y

SX

(UURU
$PSOLILHU

3KDVH/RFNHG
/RRS


, T

*DWH
3DWWHUQ
/RJLF

, T

WLPH

UDG

,QYHUWHU

PV



H


Fig. 47 Control Block Diagram of a Static Synchronous Compensator

quadrature component, I1q*, of the inverter current is defined


to be either positive if the STATCOM is emulating an inductive reactance or negative if it is emulating a capacitive reactance. The DC link capacitor voltage, vDC, is dynamically
adjusted in relationship with the inverter voltage. The control scheme used in this section shows the implementation of
the inner current control loop which regulates the reactive
current flow through the inverter regardless of the line voltage. However, if one is interested in regulating the line voltage, an outer voltage control loop must be implemented. The
outer voltage control loop will automatically determine the
reference reactive current for the inner current control loop
which, in turn, will regulate the line voltage.
Fig. 48 shows the digital simulation results from the

L D



D

set of measured inverter currents, i1, is decomposed


into its real or direct component, I1d, and reactive or quadrature component, I 1q , respectively. The quadrature component is compared with the desired rference value, I1q *, and
the error is passed through an error amplifier which produces
a relative angle, , of the inverter voltage with respect to the
line voltage. The phase angle, 1, of the inverter voltage is
calculated by adding the relative angle, , of the inverter
voltage and the phase-lock-loop angle, . The reference

D





Fig. 48 Performance of a Static Synchronous Compensator with a


12-Pulse Harmonic Neutralized Inverter Operating in Capacitive and
Inductive Modes

The inverter voltage set, e1, is greater than the line


voltage set, v1. At 175 ms, the quadrature current demand,
I1q* , of the inverter is set to one per unit inductive, which
means the STATCOM should see the system as a capacitive reactance and the inverter current in phase a, i1a, leads

the line voltage at phase a, v1a, by almost 90 . The inverter


voltage set, e1, is less than the line voltage set, v1. At 250 ms,
the quadrature current demand, I1q*, of the inverter is set to
one per unit capacitive and the transition takes place in a subcycle time. The phase angle, , between the inverter voltage
and the line voltage is dynamically adjusted so that the
inverter maintains proper DC link capacitor voltage.
Fig. 49 shows the expanded view of two sections of
Fig. 48. The inverter voltage and current show the presence

2-17

of 12-pulse harmonic components.


&RQWUROOHU

9 $

D

L D

SX

D

$%6


9'&

(UURU

LQY

Y
,

'&

&DOFXODWRU
,G





)UDPH
L

PV

L D



Y


H

D



7UDQVIRUPHU

,T

D

5RWDWLQJ

WLPH

  

DQG $QJOH



0DJQLWXGH



$PSOLILHU

3KDVH/RFNHG

6*1

LU
L

*DWH
3DWWHUQ

,QYHUWHU

/RJLF

/RRS

Fig. 50 Control Block Diagram of a Static Synchronous Series Compensator


WLPH
PV



Fig. 49 Waveforms from a Static Synchronous Compensator with a


12-Pulse Harmonic Neutralized Inverter Operating in Capacitive and
Inductive Modes

B. SSSC
An SSSC controller uses a solid-state voltage source
inverter to inject an almost sinusoidal voltage, of variable
magnitude, in series with a transmission line. This injected
voltage is almost in quadrature with the line current. A small
part of the injected voltage which is in phase with the line
current provides the losses in the inverter. Most of the
injected voltage which is in quadrature with the line current
emulates an inductive or a capacitive reactance in series with
the transmission line. This emulated variable reactance,
inserted by the injected voltage source, influences the electric
power flow in the transmission line. If an SSSC is operated
with an energy storage system, the controller becomes an
impedance compensation controller which can compensate
for the transmission line resistance as well as reactance. The
reactance compensation controller is used to operate the
inverter in such a way that the injected alternating voltage in
series with the transmission line is proportional to the line
current with the emulated reactance being the constant of
proportionality [57]. When an SSSC injects an alternating
voltage leading the line current, it emulates an inductive
reactance in series with the transmission line causing the
power flow as well as the line current to decrease as the level
of compensation increases and the SSSC is considered to be
operating in an inductive mode. When an SSSC injects an
alternating voltage lagging the line current, it emulates a
capacitive reactance in series with the transmission line causing the power flow as well as the line current to increase as
the level of compensation increases and the SSSC is considered to be operating in a capacitive mode. An SSSC controller can also be used for stable reversal of power flow in the
transmission line.

Fig. 50 shows a control block diagram of an SSSC.


An instantaneous 3-phase set of measured line currents, i, is
first decomposed into its real or direct component, I d , and
reactive or quadrature component, Iq, and then the amplitude,
I, and the relative angle, ir, of the line current with respect
to the phase-lock-loop angle, , are calculated. The phase
angle, i, of the line current is calculated by adding the relative angle, ir, of the line current and the phase-lock-loop
angle, . The calculated amplitude, I, of the line current
multiplied by the compensating reactance demand, Xq*, is the
insertion voltage amplitude demand, Vq*. The phase angle,

v, of this insertion voltage demand is either i+90 if the

demanding compensating reactance is inductive or i-90 if


the demanding compensating reactance is capacitive. The
DC link capacitor voltage is dynamically regulated in relationship with the insertion voltage amplitude demand. The
insertion voltage amplitude demand, Vq* , and the DC link
capacitor voltage demand, VDC*, are related by the inverter
DC-to-fundamental AC amplitude gain factor (Kinv = 2/ for
a true harmonic netralized voltage source inverter). The DC
link capacitor voltage demand, VDC*, and the measured DC
voltage, vDC, are compared and the error is passed through an
error amplifier which produces an angle, . The phase angle,
2, of the inverter voltage is calculated by adding the angle,
, of the DC voltage regulator and the phase angle, v, of the
insertion voltage demand. The compensating reactance
demand, Xq*, is either negative if the SSSC is emulating an
inductive reactance or positive if it is emulating a capacitive
reactance. In another application, the insertion voltage
amplitude demand, Vq * may directly be specified and the
SSSC will inject the desired voltage almost in quadrature
with the line current.

2-18

9 $ ; 3

LD

SX

9 $ ; 3
3T

SX

Y '&

LD
3T

'&





D

4T


H

D

WLPH
;
T

PV


3T

4T

LD


;T


; T

WLPH


PV






Y '&

Fig. 51 Performance of a Static Synchronous Series Compensator


with a 24-Pulse Harmonic Neutralized Inverter Operating in Inductive and Capacitive Modes

Fig. 51 shows the digital simulation results when an


SSSC emulates a reactance in series with the transmission
line. At the beginning of the operation, the mechanical
switch, MS2, and the electronic switch, ES22, are open and
the electronic switch, ES2, is closed. The inverter, VSI2,
injects no voltage. The DC link capacitor voltage, v DC, is
zero. At 50 ms, an inductive reactance compensation of 0.15
per unit is requested. The inverter output 24-pulse voltage,

e2a, of phase a leads the line current, ia, by almost 90 . At


175 ms, the inductive reactance demand is increased to 0.3
per unit. As the inductive reactance demand increases, the
line current, ia, and the power flow, Pq and Qq, in the transmission line decrease. At 300 ms, a capacitive reactance
compensation of 0.1 per unit is requested. The inverter volt
age, e2a, lags the line current, ia, by almost 90 . At 450 ms,
the capacitive reactance demand is increased to 0.15 per unit.
As the capacitive reactance demand increases, the line current, ia, and the power flow, Pq and Qq , in the transmission
line increase. In reality, the SSSC would encounter power
losses in the valves and in the magnetic circuit. Therefore,
there will always be a small part of real current component,
I1d, flowing into the inverter and the inverter voltage will be

almost 90 out of phase with the line current. The instantaneous DC link capacitor voltage is proportional to the amplitude of the inverter voltage.
Therefore, when an SSSC emulates a reactance in
series with the transmission line, the power flow in the transmission line always decreases if the emulated reactance is
inductive. Also, the power flow always increases if the emulated reactance is capacitive.
Fig. 52 shows the expanded view of the two sections of Fig.
51. The inverter voltage show the presence of 24-pulse harmonic components.






WLPH

D

PV


4T

Fig. 52 Waveforms from a Static Synchronous Series Compensator


with a 24-Pulse Harmonic Neutralized Inverter Operating in Inductive and Capacitive Modes

C. UPFC
The stand alone operations of the STATCOM and
the SSSC, as just described, only allow the inverters to
exchange almost exclusively reactive power at their terminals. However, if both the inverters are operated from a common DC link capacitor, the injected voltage by the SSSC can
be at any angle with respect to the line current. The real
power exchanged at the terminals of the SSSC appears at the
terminals of the STATCOM through the DC link capacitor.
The STATCOM can still be used to control the reactive current flow through it independently [58]. The current injected
by the STATCOM has two components. First, a real or direct
component, which is in phase with the line voltage, absorbs
or delivers the real power exchanged by the SSSC with the
line. Second, a reactive or quadrature component, which is
in quadrature with the line voltage, emulates an inductive or
a capacitive reactance at the point of connection with the
transmission line.
The SSSC can be operated in many different modes,
such as voltage injection, phase angle shifter emulation, line
impedance emulation, automatic power flow control, etc. In
each mode of operation, the final outcome is such that the
SSSC injects a voltage in series with the transmission line
[58]. In this section, the SSSC is operated in a voltage injection mode. The control block diagram for the SSSC is shown
in Fig. 53.
The desired peak fundamental voltage, Vdq*, at the
output of the inverter and its relative angle, , with respect to
the reference phase-lock-loop angle are specified. The phase
angle, 2, of the inverter voltage is calculated by adding the
relative angle, , of the inverter voltage and the phase-lock-

2-19

loop angle, . The dead angle of each pole is calculated in


accordance with the operation of 24-pulse quasi harmonic
neutralized inverter [58].

&RQWUROOHU

9 GT

'HDG $QJOH
&DOFXODWRU

Y '&

3KDVH/RFNHG
/RRS

Y

*DWH
3DWWHUQ
/RJLF

,QYHUWHU

Fig. 53 Control Block Diagram of a Static Synchronous Series Compensator

Fig. 54 shows the digital simulation results from the


voltage injection mode of operation of an SSSC while the
STATCOM is operated to deliver no reactive current.

9 $ 3 4
SX

Y '&
3U


, T


4U





injects no voltage. The voltage, v12a, at the terminals of the


coupling transformer, T2, is the voltage across its leakage
reactance. The mechanical switch, MS1, is open, disconnecting the STATCOM from the transmission line. The DC link
capacitor is precharged. At 50 ms, MS1 closes and the
quadrature current demand, I1q*, of the inverter is set to zero.
At 100 ms, a series voltage injection of 0.2 per unit at an

angle of 60 leading the reference phase-lock-loop angle is


requested. The series inverter output voltage, e2a, of phase a
leads the line current, i a , by an angle . The real power
absorbed by the series inverter appears at the BUS 1 through
the STATCOM. The shunt inverter output voltage, e1a, of
phase a is in phase with the current, i1a, flowing through it.
The power delivered at the receiving end decreases. At 175
ms, the injected voltage request is increased to 0.4 per unit
while maintaining the same angle. As the voltage injection
demand increases, the line current, ia, and the power flow, Pr
and Qr, in the transmission line decrease. By injecting a voltage by the SSSC of any magnitude, within the rating of the
inverter, and at any angle with respect to the line current, the
real power, Pr, and the reactive power, Q r, at the receiving
end of the transmission line can be increased, decreased or
even reversed selectively.
Fig. 55 shows the expanded view of two sections of
Fig. 54. The inverter voltage and current show the presence
of harmonic components.


WLPH

PV

D

D

9 $
SX


L D

LD

D

H
D

WLPH



PV









LD

D

D

D

D







WLPH
PV





LD

Fig. 55 Waveforms from a Unified Power Flow Controller with a


24-Pulse Quasi Harmonic Neutralized Inverter with 3-Level Poles
Operating in a Voltage Injection Mode

Fig. 54 Performance of a Unified Power Flow Controller with a 24Pulse Quasi Harmonic Neutralized Inverter with 3-Level Poles Operating in a Voltage Injection Mode

At the beginning of the operation, the mechanical


switch, MS2, and the electronic switch, ES22, are open and
the electronic switch, ES2, is closed. The inverter, VSI2,

4.5.4 SUMMARY
FACTS devices - STATCOM, SSSC, and UPFC,
have been modeled using an EMTP simulation package. The
UPFC consists of two voltage source inverters - one injects
an almost sinusoidal voltage in series with the transmission

2-20

line and the other injects an almost sinusoidal current at the


point of connection. The injected voltage can be at any angle
with the line current. The injected current has two parts.
First, the real part, which is in phase with the line voltage,
delivers or absorbs real power to the line that is exchanged by
the injected voltage source plus losses in the UPFC. Second,
the reactive part, which is in quadrature with the line voltage,
emulates an inductive reactance or a capacitive reactance at
the point of connection. The SSSC model has been operated
injecting a voltage in series with the transmission line. The
STATCOM model has been operated regulating the reactive
current flow through it and the transition from one mode of
operation to the other mode takes place in a subcycle time.
The operation of the model is verified with the model connected to a simple transmission line which can easily be
replaced by the utilitys existing more complex power system
network.

3.

5. CONCLUSIONS

10.

The appropriate characterization of the power electronics is


very important in power system simulations involving power
electronics operations. In most of these simulations, detailed
representations of the power electronics are not necessary.
Depending on the objective of a study, the involved power
electronics subsystem can be always reduced to some extend
with minimal loss of accuracy.
Numbers of the digital computation tools are capable of simulating power electronics cases. However, in power systems
engineering community, the EMTP type of programs are
more commonly used. This results mainly from the great capabilities and flexibility of these programs in handling conventional power system dynamics and electromagnetic
transients beside their capabilities of handling power electronics. With an adequate power electronics device and circuit
simplification, the EMTP type of programs are powerful for
modeling various types of power electronics applications.
Because these programs are based on the time domain solution method, the dynamic interaction between the power electronics and the rest of the system can be easily incorporated in
simulation.
The important considerations for simulating power electronics applications have been summarized in this guidelines.
Several modeling examples including a comprehensive treatment of voltage source inverter based FACTS device and its
modeling techniques using EMTP type of programs were presented. The procedures used to implement power electronics
models in these examples are valuable for using other digital
simulation tools.
6. REFERENCES
1. CIGRE Working Group 02 (SC 33), "Guidelines for Representation of Network Elements when Calculating Transients", 1990.
2. J. Blanes, A. Quijano, S. Catal n, "Simulation of induction motors behaviour during voltage disturbances using
an aggregated model", European EMTP/ATP Users
Group Meeting, November 9-11,1997, Barcelona.

4.
5.
6.
7.

8.
9.

11.
12.

13.

14.

15.

16.

17.

18.

19.
20.

2-21

P. F. Ribeiro, "Distribution system and other elements


modeling", Chapter 3 of Tutorial on Harmonics Modeling and Simulation, IEEE TP-125-0, 1998.
"Symposium on Synchronous Machines Modelling for
Power System Studies", IEEE Power Engineering Society, Publication 83THO101-6-PWR, 1983.
P. M. Anderson, B. L. Agrawal and J. E. Van Ness, Subsynchronous Resonance in Power Systems, IEEE Press,
1990.
"IEEE Guide: Test Procedures for Synchronous Machines", IEEE Std. 115, 1983.
F. L. Alvarado and C. Canizares, "Synchronous machine
parameters from sudden-short tests by back-solving",
IEEE Trans. on Energy Conversion, Vol. 4, no. 2, pp.
224-236, June 1989.
I. M. Canay, "Modelling of alternating-current machines
having multiple rotor circuits", IEEE Trans. on Energy
Conversion, vol.8, no. 2, pp. 280-296, June 1993.
"IEEE Guide: Test procedure for polyphase induction
motors and generators", IEEE Std. 112-1978.
G. J. Rogers and D. Shirmohammadi, "Induction machine modelling for electromagnetic transient program",
IEEE Trans. on Energy Conversion, vol. 2, no. 4, pp. 622628, December 1987.
"IEEE Standard Procedures for Obtaining Synchronous
Machine Parameters by Standstill Frequency Response
Testing", IEEE Std.115A, 1987.
I. M. Canay, "Determination of the model parameters of
machines from the reactance operators xd(p), xq(p)
(Evaluation of standstill frequency response test)", IEEE
Trans. on Energy Conversion, vol. 8, no. 2, pp. 272-279,
June 1993.
A. Keyhani and H. Tsai, "Identification of high-order
synchronous generator models from SSFR test data", presented at the 1994 IEEE/PES Winter Meeting, Paper no.
94 WM 022-4 EC, New York, January 30-February 3,
1994.
J. R. Willis, G. J. Brook and J. S. Edmonds, "Derivation
of induction motor models from standstill frequency response", IEEE Trans. on Energy Conversion, vol. 4, no.
4, pp. 608-615, December 1989.
P. L. Dandeno, P. Kundur, A. T. Poray and H. M. Zein Eldin, "Adaptation and validation of turbogenerator model
parameters through on-line frequency response measurements", IEEE Trans. on Power Apparatus and Systems,
vol. 100, no. 4, pp. 1656-1645, April 1981.
P. L. Dandeno, P. Kundur, A. T. Poray and M. E. Coultes,
"Validation of turbogenerator stability models by comparison with power system tests", IEEE Trans. on Power
Apparatus and Systems, vol. 100, no. 4, pp. 1637-1645,
April 1981.
F. P. de Mello, L. N. Hannett, J. R. Willis, "Determination of synchronous machine stator and field leakage inductances standstill frequency response tests", IEEE
Trans. on Power Systems, vol. 3, no. 4, pp. 1625-1632,
November 1988.
L. Dube, H.W. Dommel, Simulation of control system
in an Electromagnetic Transient Program with TACS,
IEEE Trans. on Power Industry and Computer Applications, 1977
EMTP Rule Book, EPRI/DCG Version 1.0.
D. Goldsworthy, J. J. Vithayathil, EMTP model of an

21.
22.

23.

24.
25.

26.

27.

28.

29.

30.
31.

32.
33.
34.
35.
36.

37.
38.

HVDC transmission system, Proceedings of the IEEE


Montech 86 Conference on HVDC Power Transmission,
September 26-October 1, 1986, pp. 39-46
L. X. Bui, S. Casoria, G. Morin, Modeling of digital
controls with EMTP, CEA Meeting, March 25-29,
1989, Montreal, Canada
J. Reeve and S. P. Chen, Versatile interactive digital
simulator based on EMTP for AC/DC power system transient studies, IEEE Trans. on Power Apparatus and Systems, Vol. 103, No. 12, December 1984, pp. 3625-3633
K. G. Fehrle, R. H. Lasseter, Simulation of control systems and application to HVDC converters, IEEE Tutorial Course 81 EHO173-PWR on Digital Simulation of
Electrical Transient Phenomena, 1981.
L. X. Bui, G. Morin, J. Reeve, EMTP TACS-FORTRAN interface development for digital controls modeling, 91 SM 417-6 PWRS
G. Morin, L. X. Bui, S. Casoria, J. Reeve, Modeling of
the Hydro-Quebec - New England HVDC system and
digital controls with EMTP, IEEE Trans. on Power Delivery, Vol. 8, No. 2, April 1993, pp. 559-566.
R. H. Lasseter and S. Y. Lee, Digital simulation of static
var system transients, IEEE Trans. on Power Apparatus
and Systems, Vol. PAS-101, No. 10, pp. 4171-4177, October 1982.
A. M. Gole and V. K. Sood, A static compensator model
for use with electromagnetic transients simulation programs, IEEE Trans. on Power Delivery, Vol. PWRS-5,
No. 3, pp. 1398-1407, July 1990
A. N. Vasconcelos et. al. Detailed modeling of an actual
static Var compensator for electromagnetic transients
studies, IEEE Trans. on Power Systems, Vol. PWRS-7,
no. 1, pp. 11-19, February 1992
S. Y. Lee et al., Detailed modeling of static Var compensators using the Electromagnetic Transients Program
(EMTP), IEEE Trans. on Power Delivery, Vol. 7, no. 2,
pp. 836-847, April 1992
S. Lefebvre and L. Gerin-Lajoie, A static compensator
model for the EMTP, IEEE PES Meeting, San Diego,
July 28-August 1, 1991, Paper 91 SM 461-4 PWRS.
L. Dube and I. Bonfanti, MODELS: A new simulation
tool in the EMTP, European Transactions on Electrical
Power Engineering, Vol. 2, no. 1, pp. 45-50, January/
February 1992.
Leuven EMTP Center (ed.), ATP Rule Book, 1990.
J. A. Martinez, Simulation of a microprocessor-controlled SVC, 21th European EMTP Meeting, June 5-7,
1992, Crete (Greece).
H.W. Dommel, EMTP Reference-Manual (EMTP Theory Book), BPA, 1986.
J. A. Martinez, Simulation of power electronics using
the EMTP, Part I: Power converters, A survey, UPEC94, September 14-16, 1994, Galway.
G. A. Capolino, H. Henao, ATP simulation for power
electronics and AC drives, 15th European EMTP Users
Group Meeting, Paper 88R-027, October 17-18, 1988,
Leuven.
G. A. Capolino, H. Henao, Simulation of electrical machine drives with EMTP, 18th European EMTP Users
Group Meeting, Paper M7, May 28-29, 1990, Marseille
J. A. Martinez, G. A. Capolino, TACS and MODELS:
Drive simulation languages in a general purpose pro-

39.
40.

41.

42.

43.

44.

45.

46.

47.

48.

49.

50.

51.

52.

2-22

gram, Proc. MCED91, Marseille, July 1-2, 1991, pp.


R1-R13.
G. A. Capolino, H. Henao, ATP advanced usage for
electrical drives, EMTP Summer Course, July 5-8,
1993, Leuven.
H. Knudsen, Extended Parks transformation for 2 by 3phase synchronous machine and converter phasor model
with representation of harmonics, IEEE PES Summer
Meeting, Paper 94 SM 350-9 EC, July 24-28, 1994, San
Francisco.
M. Mazzucchelli, G. Sciutto, Digital simulation of AC
electrical drives based on field-oriented control method
using a general purpose program, Proceedings PCIM,
pp. 350-364, 1986, Munchen
Z. Daboussi, N. Mohan, Digital simulation of field-oriented control of induction motor drives using EMTP,
IEEE Trans. on Energy Conversion, Vol. 3, pp. 667-673,
September 1988.
L. Tang, M. McGranaghan, Modeling an active power
line conditioner for compensation of switching transients, Proceedings of First International Conference on
Power Systems Transients (IPST95), Lisbon (Portugal),
pp. 403-408.
X. Z. Meng, J. G. J. Sloot, H. Rijanto, Modelling of
semiconductor fuses in EMTP, Proceedings of First International Conference on Power Systems Transients (IPST95), Lisbon (Portugal), pp. 481-486.
J. A. Martinez-Velasco, R. Abdo, G.A. Capolino, Advanced representation of power semiconductors using the
EMTP, Proceedings of First International Conference
on Power Systems Transients (IPST95), Lisbon (Portugal), pp. 505-510.
S. Lefebvre, R. D. Rangel, Modeling of power electronics devices in EMTP-TACS, Proceedings of First International Conference on Power Systems Transients
(IPST95), Lisbon (Portugal), pp. 511-516.
G. A. Capolino, H. Henao, D. Leduc, V. T. Nguyen
Phuoc, CAD of field-oriented induction motor drives
using a general purpose program, Proceedings PCIM,
1989, Munchen.
E. Bassily, G. A. Capolino, H. Henao, Simulation of discrete DC drive using sliding mode control, 22nd European EMTP Users Group Meeting, Paper 92R-018,
November 9-10, 1992, Leuven.
E. Bassily, G. A. Capolino, H. Henao, Simulation and
design of brushless motor drive control with fuzzy pi regulator, Proceedings of First European Conference on
Power Systems Transients (EPST93), Lisbon (Portugal), pp 84-91, June 1993.
A. K. Khan, R. Dwyer, M. Mcgranaghan, L. Tang, Evaluation of Harmonic Impacts from Compact Fluorescent
Lights on Distribution Systems, IEEE PES 95 WM 1057-PWRS.
J. A. Martinez, EMTP simulation of digitally-controlled
static Var system for optimal load compensation, IEEE
PES Summer Meeting, Paper 94 SM 452-3 PWRD, July
24-28, 1994, San Francisco.
J. A. Martinez and G. A. Capolino, EMTP simulation of
power electronics and drives using data modularization,
Proceedings First International Aegean Conference on
Electrical Machines and Power Electronics, Kucadasi
(Turkey), May 27-29, 1992.

53. L. Tang, M. F. McGranaghan, R. A. Ferraro, S. Morganson, b. Hunt, Voltage notching interaction caused by
large adjustable speed drives on distribution systems with
low short circuit capacities, IEEE PES 95 SM 388-9PWRD.
54. O. B. Nayak, A. M. Gole, D. G. Chapman, and J.B.
Davies, Dynamic performance of static and synchronous compensators at an HVDC inverter bus in a very
weak AC system, IEEE Trans. on Power Systems, Vol.
9, NO. 3, August 1994, pp. 1350-1358.
55. M. Szechtman, T. Wess, C. V. Thio, First benchmark
model for HVDC control studies, Electra, No. 135,
April 1991.
56. K. K. Sen, STATCOM - STATic synchronous COMpensator: Theory, Modeling, and Applications, 99WM607,
IEEE PES Winter Meeting, New York, USA, 1999.
57. K. K. Sen, SSSC - Static Synchronous Series Compensator: Theory, Modeling, and Applications, IEEE Transactions on Power Delivery, vol. 13, No. 1, January 1998.
58. K. K. Sen and E. J. Stacey, UPFC - Unified Power Flow
Controller: Theory, Modeling, and Applications, IEEE
Transactions on Power Delivery, vol. 13, No. 4, October
1998.

2-23

You might also like