You are on page 1of 29

DRAFT July 98: Note: Figure numbering to be fixed.

Some figures are too light

MODELING GUIDELINES FOR LOW FREQUENCY TRANSIENTS


Report Prepared by the Low-Frequency Transients Task Force
of the IEEE Modeling and Analysis of System Transients Working Group
Contributing Authors: R. Iravani (Chair)A.K.S. Chandhury
I.D. Hassan, J.A. Martinez, A.S. Morched,
B.A. Mork, M. Parniani, D. Shirmohammadi, R.A. Walling
Abstract: The objective of this report is to provide guidelines
for modeling and analyses of low-frequency (approximately 5 to
1000 Hz) transients of electric power systems, based on the use
of digital time-domain simulation methods. For the ease of reference, the low-frequency transients are divided in seven distinct phenomena. This report (1) briefly describes the physical
nature of each phenomenon, (2) identifies those power system
components/apparatus which either contribute to or are
affected by the phenomenon, (3) provides guidelines for digital
time-domain simulation and analyses of the phenomenon and
(4) provides sample study-system and typical digital timedomain simulation results corresponding to each phenomenon.
A comprehensive list of reference is also included in this report
to provide further in-depth information to the readers.

1.Torsional oscillations (5 to 120 Hz)

Keywords: Low-Frequency Transients, Electromechanical


Transients, Modeling, Time-Domain Analysis, Torsional
Dynamics, Turbine Vibrations, Bus-Transfer, Controller
Interactions, Harmonic Interactions, Ferroresonance

For each of the above phenomenon this report provides (1) a


brief explanation of the physical phenomenon, (2) modeling
guidelines for time-domain simulation and analyses, and (3)
typical sample systems and simulation results.

2.Transient torsional torques (5 to 120 Hz)


3.Turbine blade vibrations (90 to 250 Hz)
4.Fast bus transfer (1 to 1000 Hz)
5.Controller interactions (10 to 30 Hz)
6.Harmonic interactions and resonances (60 to 600 Hz)
7.Ferroresonance (1 to 1000 Hz)

1. INTRODUCTION
This report is intended for practicing power system engineers
who are involved in system analysis, system control, and system planning. To use the report efficiently, adequate understanding of the physical phenomenon of interest and
familiarity with the concepts and techniques of digital computer simulation approaches are necessary.

An interconnected power system can experience undesirable


oscillations and transients as a result of small-signal perturbations, large-signal disturbances, and nonlinear characteristics
of the system components. The oscillations cover a wide frequency range approximately from 0.01 Hz to 50 MHz. Oscillations in the frequency range of 0.01 to 1000 Hz are defined
in this report as low-frequency (slow) transients. We interchangeably use the terms slow transients, low frequency(LF) dynamics, and LF oscillations throughout this
report. All the issues relevant to low-frequency inter-area
electromechanical oscillations (approximately 0.1 to 1 Hz)
and classical turbine-generator swing modes (approximately
1 to 2.5 Hz) are discussed by other IEEE working groups, and
are not discussed here. A general guideline for representation
of network elements for electromagnetic transient studies
have been previously published [1.1]. The mandate of the
IEEE Low-Frequency Transients Task Force is to provide
modelling guidelines for time-domain analysis of LF oscillations within the frequency range of 5 to 1000 Hz. Low frequency dynamics are of concern with respect to power system
stability issues and/or temporary overvoltages.

Section 2 of the report deals with low-frequency transients


which involve both electrical and mechanical dynamics, i.e.,
torsional oscillations, transient torsional torques, turbineblade vibrations and fast bus-transfer. Section 3 discusses
low-frequency electrical dynamics, as a result of control systems interactions. Section 4 provides analysis guidelines for
harmonic interactions and resonance phenomena. The phenomenon of ferroresonance is discussed in Section 5.

2. LOW-FREQUENCY ELECTROMECHANICAL
DYNAMICS
This section provides modeling and analysis guidelines for low-frequency dynamics which involve electromechanical oscillations. The phenomena which are covered in
this section are torsional oscillations, transient torques, tur-

phenomena of 60 Hz power systems in the LF range are divided into the following categories:

3-1

the natural frequencies of the shaft torsional oscillatory


modes. Usually, the oscillatory mode at the first torsional frequency dominates the shaft transient oscillations. The major
incidents which result in severe shaft stresses are: line-to-line
faults, three-phase faults, fault clearing, automatic reclosures,
and out-of-phase synchronization. The amplitudes of the
shaft transient stresses can be particularly large when the network is equipped with series capacitors.

bine-blade vibrations, and bus-transfer.


2.1 DEFINITIONS
2.1.1 Torsional Oscillations [2.1, 2.2, 2.3, 2.4, 2.5]
Shaft system of a steam turbine-generator experiences torsional oscillations when one or more of its natural oscillatory
modes, usually at subsynchronous frequencies, are excited.
Sustained or negatively damped torsional oscillations occur
when a turbine-generator shaft system exchanges energy with
an electrical system at the shaft oscillatory modes. This exchange of energy can exist if the electrical system is equipped
with either series capacitors or HVDC converter stations. The
phenomenon of torsional oscillations can also exist as a result
of interaction between the shaft system of a steam turbinegenerator and

High amplitude shaft mechanical stress can induce significant


fatigue in the shaft segments and result in noticeable shaft
life-time reduction during each oscillatory cycle. Such oscillations may even result in catastrophic shaft failure. The primary purpose of time-domain investigation of turbinegenerator shaft mechanical stresses is to identify the peak
torques imposed on the shaft segments, after system disturbances. Transient shaft mechanical stresses calculated based
on time-domain simulation methods also can be used to estimate shaft loss of life as a result of system disturbances.

the generator excitation systems through either AVR or PSS


control loops,

2.1.3 Turbine-Blade Vibrations [2.6]

electronically controlled governor system,

Frequencies of turbine-blade vibrational modes are


usually within 90 to 250 Hz, and constitute supersynchronous frequency modes. Identification of supersynchronous
frequency modes and their corresponding frequencies is best
carried out by solving elasticity equation of the shaft system
as a continuum, based on the use of finite element methods.
This approach is beyond the scope of this report and usually
carried out by turbine manufacturers.

voltage control loop of an electrically close static VAR. compensator (SVC)


large electric arc furnaces.

Although AVR, PSS and governor system can excite torsional


oscillations, the excitation is primarily due to inadequate control design considerations and can be avoided by introducing
filters in the control circuitry. Thus, this report does not consider the generator controls as the main contributors to the
phenomenon of torsional oscillations (Table 2.1).

In this report, the objective is to investigate the impact of


large-signal disturbances on those supersynchronous frequency natural modes which are the reason for turbine-blade vibrations. Thus the required model is tailored to represent
particular supersynchronous modes and not all of them.

The phenomenon of torsional oscillation is referred to as subsynchronous resonance (SSR) when it is a result of interaction
between a shaft system and a series capacitor compensated
transmission line. The problems associated with the phenomenon of small-signal torsional oscillations are:

The concern with turbine-blade vibrations is fracture


and loss-of-life of the blades due to the fatigue induced in the
blades by repetitive or sustained oscillations. Vibrations of turbine-blades can be excited by large-signal electrical disturbances, e.g. faults, fault clearing, line switching, reclosure, and
out-of-phase synchronization.

i ) Sustained or even negatively damped oscillations which


are considered as small-signal instability problems, and
ii ) (loss of life of turbine-generator shaft segment(s) due to the
fatigue induced in the shaft segment(s) as a result of each
oscillatory cycle.

2.1.4 Fast Bus Transfer [2.7, 2.8, 2.9]


Motors and other loads in utility and heavy industrial applications are supplied during normal operation from a preferred
power source. An alternate power source is normally provided to supply such motors and other loads during planned shutdowns and upon loss of normal power from the preferred
power source. The process of disconnecting the motors and
other loads from one source and reconnecting to an alternate
source is commonly defined as bus transfer. Manual transfer

2.1.2 Transient Torsional Torques [2.1, 2.2, 2.3, 2.4, 2.5]


The shaft segments of turbine-generator units are exposed to
large-amplitude, oscillatory, mechanical stresses as a result of
electric network faults, and planned and unplanned switching
incidents. Frequencies of the shaft mechanical stresses are

3-2

means are normally provided to allow transferring the motors


and other loads from one power source to the other. However,
upon loss of the preferred power source, the motors and other
loads are automatically transferred to the alternate power
source. This automatic transfer is necessary to allow uninterrupted operation of the motors and other loads important to
personnel safety and process operation. This report does not
address the concept of bus transfer by means of semi-conductor switches [2.23].

mode which propagates almost through the entire of an interconnected electric network, the phenomena described in Section 2.1 are experienced only within a limited part of the
network. The section of the network which experiences the
phenomenon of interest, and must be represented in adequate
detail for the study of the phenomenon, is referred to as the
Study Zone The rest of the network is referred to as the external system The external system is represented by an
equivalent model. Identification of border nodes of the study
zone for a meshed network requires significant familiarity
with the network, as well as engineering judgment. As of
now, there is no straightforward and systematic approach to
identify the border nodes. One approach involves multiple
harmonic analyses of the system under investigation as
boundaries are extended to identify if new resonant frequencies (at the frequency range of interest) with low dampings exist.

The normal and alternate power source connections are always selected such that they are in phase. Therefore, manual
transfers can be accomplished in a make-before-break, i.e.,
the motors and loads are connected to the second power
source before the first power source is disconnected. In this
overlapping transfer, the power supply is not interrupted and
the motors are not subjected to transients. However, during
automatic transfers, the motors may be disconnected from
both power sources for a short duration depending on the type
of transfer and the associated circuit breakers operating times.
The time during which the motors are disconnected from both
power sources is termed the dead time. Dead time is usually
between two cycles to 12 cycles. If the relative angle between
the motor residual voltage and the power source voltage becomes large enough at the time of reconnection with significant residual voltage remaining, the resultant voltage

Proper determination of the study zone can exert a major impact on the investigations of torsional dynamics and transient
torques. Comparatively, the impact of the study zone on the
vibrations of turbine blades is less significant. Identification
of the study zone for bus transfer studies is relatively straightforward.
2.2.2 Component Model

between the power source and the motor will produce an inrush current. The inrush current may be significantly largely
than the normal full voltage staging current. Such high inrush
currents cause high winding stresses and transient shaft
torques which can damage the motor and/or the driven equipment.

Table 1 identifies the study zone components and their equivalent models for investigations of slow transient phenomena.
Further explanation of the system components are given in the
following sections.

2.2.2.1 Synchronous Generator Electrical System [2.10]


The most common bus transfer scheme is the fast bus transfer
scheme. In this scheme, opening of the normal power source
breaker initiates closing of the alternate power source breaker
without intentional time delay. Fast bus transfer operations
result in the motors being disconnected from both power
sources for a duration of as short as two cycles to as long as
12 or more cycles.

Figure 2.1 shows a second-order and a third-order


models of a synchronous machine. Inclusion of the differential leakage inductance Lf1d in the second-order model is
recommended. The differential leakage inductance has
noticeable influence on the damping, and the range of instability of each torsional mode, (with respect to series compensation level), particularly for a salient pole machine.
However, Lf1d does not influence the phenomenon of blade
vibrations.

Presently, there are no generic criteria to ensure


acceptable fast bus transfer operations. Therefore, it is necessary to analyze the transient behavior of motors during fast
bus transfer operations. The analysis should be on a case by
case basis to ensure that the motors will not be subjected to
excessive inrush currents and/or shaft transient torques.

Representation of machine electrical system based on


the third-order model, Fig. 2.1, is more accurate. Inclusion of
the differential leakage inductance Lf12d in the third-order
model has the same impact as that of Lf1d for the second-order
model. Magnetic saturation of a synchronous machine, both on
d-axis and q-axis, does not have any significant impact on the
phenomenon of small-signal torsional oscillations, but has pro-

2.2 MODELING GUIDELINES


2.2.1 Study Zone
In contrast to an inter-area, electromechanical, oscillatory
3-3

Component

T urbine-B lade
Vibrations
Third-Order Model
(d-q-o Model)
Including
Saturation

Fast Bus
Trans fer
Not
applicable

Not
Applicable

C onventional
Low-Frequency
Model including
Saturation
Characteristic

Detail
Mass-SpringDashpot Model
Conventional
Low-Frequency
Model including
Saturation
Characteristic

Equivalent- %
Model
Ideal Capacitor

Equivalent- %
Model
Ideal Capacitor

Equivalent-%
Model
Ideal Capacitor

Series R-L

Series R-L

Series R-L

Fixed I mpedance
Load

Fixed Impedance
Load

Fixed I mpedance
Load

Large Motor L oad

d-q-o Model of
Electrical System,
Mass-SpringDashpot Model of
Shaft System

Voltage Source
Behind Fixed
Impedance

Voltage Source
Behind Fixed
Impedance

HVDC Converter
Stat ion

Detailed Model of
C onverter and
Linearized
(Simplified)
Model of C ontrols
Detailed Model of
Power C ircuitry and
Linearized
(Simplified)
Model of C ontrols
Ideal Switch
Unimportant

Detailed Models of
Converter and
Controls

Detailed Models of
C onverter and
Controls

Detailed Model of
Power C ircuitry
and Controls

Detailed Model of
Power C ircuitry
and C ontrols

Not
Applicable

Ideal Sw itch
Unimportant

Ideal Switch
Unimportant

Unimportant

Series Capacitor
Overvoltages
Protection System

Ser ies Capacitor


Overvoltages
Protection System

Ideal Switch
Not
Applicable
Not
Applicable

Synchronous
Generators
Electrical System

Turbine-Generator
Shaft System
Power
Transfor mer

Transmiss ion Line


Series/Shunt
Capacitor
Series/Shunt
Reactor
Static Load

SVC

C ircuit Breaker
Generator
Controls
Protection System`

To rsion al
Oscillatio ns
Second-Order
Model and
Preferably ThirdOrder Model (d-q-o
Model)
Mass-SpringDashpot Model

T ra nsient
Torques
Third-Order
Model (d-q-o
Model)
Including
Saturation
Mass-SpringDashpot Model

Conventional
Low-Frequency
Model including
Saturation
Characteristic

Table 1: Component Model

Conventional
LowFrequency
Model
including
Saturation
C haracterist ic
Not
Applicable
Ideal
Capacitor
Series R-L
Fixed
Impedance
Load
d-q-o Model
of
Electrical
System,
Mass-SpringDashpot
Model of
Shaft System
Not
Applicable

erator unit must be separately represented.

nounced impact on transient torques and blade vibrations.

In most studies, the power plant under consideration is composed of more than one turbine-generator unit. If all the turbine-generator units are nominally identical, and under almost
equal loading conditions, they can be represented by a single,
equivalent turbine-generator unit. Otherwise, each turbinegenerator unit must be separately represented.

Fig. 2.1. Turbine-generator shaft system and its mass-spring-dashpot mode

Fig. 2.1. Synchronous machine 2nd-order and 3rd-order models

2.2.2.2 Turbine-Generator Mechanical System [2.11, 2.12,


2.13]
Fig. 2.2. Mass-spring dashpot model of the turbine-generator for turbineblade vibrational studies (mechanical damping is neglected)

Figure 2.2 shows a six-mass shaft system and its equivalent


mass-spring-dashpot model. The mass-spring-dashpot model
of Fig. 2.2 assumes that (1) the high-pressure turbine (HP), the
intermediate-pressure turbine (IP), the low-pressure turbines
(LPA and LPB), the generator rotor (G), and the excitor
(EXC) are rigid masses, and (2) each shaft section is composed of a spring constant (Kij) and a cyclic damping (Dij).
The main shortcoming of the model is that neither the shaft
cyclic dampings (Dijs) nor the viscous dampings (Dis) can
be directly measured or calculated. Neglecting the dampings
provides the most pessimistic dynamic response, which is often the objective of an investigation. The discussion of [2.11]
provides further description of the mass-spring-dashpot model. Figure 2.3 shows a mass-spring-dashpot model of the turbine-generator set of Fig. 2.2 for investigation of turbineblade vibrations. This model represents blades of turbine sections as lumped masses [2.6].

2.2.2.3 Power Transformer


Classical low frequency transformer model with proper connections at both HV and LV sides is adequate for representation of each power transformer within the Study Zone. Figure
2.4 shows the classical model of a single-phase transformer
for simulation of low frequency dynamics. No-load V-I magnetic saturation characteristic can be used as a fair approximation of core saturation for the phenomena of interest. A threephase transformer model is developed based on proper connections of primary and secondary windings of the singlephase model of Fig. 2.4.

In most studies, the power plant under consideration is


composed of more than one turbine-generator unit. If all the turbine-generator units are nominally identical, and under almost
equal loading conditions, they can be represented by a single,
equivalent turbine-generator unit. Otherwise, each turbine-gen-

3-5

2.2.2.8 HVDC Converter Station

Fig. 2.1. Low frequency model of a single-phase transformer.

2.2.2.4 Transmission Line

Shaft dynamics of a turbine-generator can be excited as a result of interaction between the turbine-generator and either
rectifier current-control or the inverter extinction angle (voltage) control of an HVDC link. Thus, if both the rectifier and
the inverter stations are within the study zone, both converter
stations, dc line, and the associated controls, with adequate
level of sophistication, must be represented in the system
model.

Equivalent- is an accurate model for representation of a long


or medium length transmission line for the phenomena under
investigation. In many reported studies, the shunt capacitive
branches of the line model are also neglected. Shunt capacitive branches of the line model do not have any major impact
on the
system subsynchronous frequency resonant modes, but their
effect on supersynchronous oscillatory modes can be noticeable. Shunt capacitive branches, particularly in the case of
long lines, have a significant effect on the system steady-state
conditions, e.g. the magnitude of generator power angle.
Therefore, depending on the operating conditions, they may
have a noticeable impact on the dampings of low frequency
oscillatory modes.

Each arm of a six-pulse converter is modelled by an ideal


switch including series and parallel snubber circuits. The
switch represents a group of series/parallel connected diodes
or thyristor valves. The three-phase transformer model of
Section 2.2.2.3 can adequately represent converter transformer of a six-pulse HVDC converter for low frequency studies.
Connection of two six-pulse converter models with proper
transformer models constitutes a 12-pulse HVDC converter
model. The model of each pole of an HVDC converter station
is realized by assembling an adequate number of 12-pulse
converter models. If small-signal dynamics are of concern,
e.g. torsional oscillations, a bipole HVDC link can be approximated by an equivalent monopolar link. Otherwise, e.g. for
investigation of transient torques, bipolar representation is
necessary. Models of smoothing reactors and ac/dc filters are
developed by proper connections of lumped RLC elements.
Multiple -sections is the recommended model of an HVDC
line.

2.2.2.5 Series and Shunt Capacitor Banks


Series capacitors are the main cause of severe shaft torsional
oscillations and their presence in each transmission section is
accurately represented by three lumped, ideal, capacitor
banks. Similar to the shunt capacitive branches of a transmission line, shunt capacitor banks do not have any direct impact
on the shaft dynamics. However, since shunt capacitors alter
the voltage profile of the system, they may noticeable impact
on the dampings of the oscillatory modes depending on the
operating condition. Thus, representation of shunt capacitors
in the system model, particularly under heavy loading conditions, is recommended.

Block diagram of the controls of a bipole Hvdc system for


time-domain simulation is given [2.14]. Further details of the
control blocks are available in Chapter 8 of [2.15].

2.2.2.6 Shunt Reactor


Shunt reactors can have a noticeable impact on the steadystate operating conditions, e.g. voltage profile, which can impact the dampings of the low frequency dynamics. Thus, representation of shunt reactors, particularly under light loading
conditions, is recommended.

When the inverter station is not within the Study Zone, the inverter station and the dc line can be represented by an equivalent controlled voltage source, and only the rectifier station
and its controls must be modelled in details. Similarly, the
rectifier station and the dc line can be modelled as an equivalent controlled current source and only the inverter station and
its control system be represented in detail, if the rectifier station is not within the Study Zone.

2.2.2.7 Loads
Fixed Impedance model is an adequate load representation when turbine-generator shaft dynamics are of concern.
However, if an induction motor load or a synchronous motor
load is comparable to the MVA rating of the turbine-generator
under consideration, fixed impedance representation of the load
may result in erroneous conclusions. Under such conditions, the
load is best represented by either an equivalent induction motor
or an equivalent synchronous motor.

An HVDC installation may have multiple auxiliary controls


for various purposes, e.g. damping inter-area oscillations, frequency control, and reactive power/voltage modulations. It is
recommended to represent such auxiliary controls in the system model to identify their possible adverse impacts on the
torsional oscillatory modes.

Motor loads must be represented in details for fast bus transfer


phenomenon. For these studies, parallel identical motor loads
can be lumped in an equivalent motor load.

2.2.2.9 Static VAr Compensator (SVC)


Field experience and theoretical studies indicate that possible

3-6

adverse effect of an SVC on the shaft torsional dynamics are


not as severe when compared with that of an HVDC converter
station [2.16]. However SVCs have been recognized as effective countermeasures for shaft torsional dynamics. A conventional SVC is composed of thyristor-switched capacitors
(TSCs) and thyristor-controlled reactors (TCRs) [2.17]. During small-signal dynamics, e.g. torsional oscillations, an SVC
can be approximated as fixed capacitors (FCs) and TCRs.thyristor valves in each arm of either the TCR or the TSC are
modelled as two equivalent ideal switches including the parallel snubber branch. The three-phase transformer model of
Section 2.2.2.3 can adequately represent an SVC transformer
for low frequency studies. Controlled reactor, switched capacitor and the SVC filter components are represented in the
time-domain simulation model by proper combinations of
lumped RLC elements. Chapter 9 of [2.15] and reference
[2.18] provide details of the controls of an SVC for time-domain simulation. Similar to an HVDC converter station, an
SVC may be equipped with auxiliary controls, e.g. supplemental SSR damping control. Thus, all the closed-loop controls must be represented in the simulation model to attain a
realistic time-response of an SVC.

benchmark models have been extensively used for time-domain as well as frequency-domain investigation of the phenomenon of torsional oscillations. Numerous study results,
using the benchmark model, have been published in the IEEE
PES Transactions [2.1, 2.2, 2.3, 2.4, 2.5].

Time-domain simulation and frequency-domain eigen analysis are widely used as complementary approaches for reciprocal verification of torsional studies.
2.3.2 Transient Torques
The first and the second IEEE benchmark models for SmallSignal torsional studies introduced in Section 2.3.1 also have
been extensively used for transient torque studies. Due to the
nonlinear nature of large-signal torsional oscillations, digital
time-domain simulation is the only approach to investigate
the phenomenon. There are no measurement results regarding
transient torques in the widely circulated technical literature.
Thus, simulation results cannot be readily compared with actual field tests. At this stage, a general verification rule is to
ensure that the simulation results satisfy the well understood
behavioral patterns and immediately after switching incidents.

2.2.2.10 Generator Controls


Conventional generator controls, i.e. automatic voltage regulator (AVR), power system stabilizer (PSS), and governor
system generally do not have major (positive or negative) effects on turbine-generator shaft dynamics. Although there are
reports of torsional excitation as a result of PSSs and electronically controlled governors, the adverse effect can be prevented by introducing filters in the control circuitry. Thus, the
dynamics of excitation and governor systems are neglected,
and the input mechanical power and the generator field voltage are considered as constant values for time-domain investigation of shaft dynamics. For those particular cases where
either AVR, PSS or governor may aggravate torsional oscillations [2.1, 2.2, 2.3, 2.4, 2.5], they can be represented by their
linearized models in the system model.

2.3.3 Turbine-Blade Vibrations


The radial power system of [2.6] is recommended as the test
system. The system is composed of a multi-mass tubine-generator which is connected to an infinite bus through two parallel lines. The system can be used to study blade vibrations
of low-pressure turbine sections. It should be noted that in
contrast to the shaft torsional oscillations (either small-signal
or large-signal), the blade vibrations are not readily quantifiable from time-domain responses. Thus, a frequency spectrum analysis, e.g. FFT should be conducted on the time
response to obtain the relative amplitudes and frequencies of
the blade dominant oscillatory modes.

2.2.2.11 Protection System


A qualitative verification of the simulation results can be obtained based on the comparison of the frequencies of the blade
vibrations, deduced from FFT of the simulation results, with
those provided by the turbine manufacturer.

Overvoltage protection system of series capacitor can have a


significant impact on large-signal torsional torques and turbine-blade vibrations following network transients. Thus, for
the simulation of these two phenomena, the series capacitor
overvoltage protection scheme including Zn0 varistor and
the associated bypass logic and power circuitry must be represented in the system model.

2.3.4 Bus-Transfer
The simplified system introduced in [2.7], is recommended as
the test system for bus transfer studies. Typical motor load
data for simulation studies are available in [2.19].

2.3 TEST SYSTEMS


2.3.1 Torsional Oscillations

Ideally, validating a model of a fast bus transfer operation


should include validating the individual motor models and the
circuit breakers operating times. Individual motor models can

The IEEE Working Group on Subsynchronous Resonance has


introduced two benchmark models for time-domain simulation of turbogenerator torsional oscillations [2.12, 2.13]. The
3-7

loop control system and a natural oscillatory mode of an apparatus. One practical case of controller interaction phenomenon is that of multiple SVCs [3.1]. The problem of controller
interactions attracts more attention as the number of power
electronic based devices increases.

be validated by simulating motor starting and running conditions and comparison of other simulation results to data recorded during an actual motor instantaneous current, power,
apparent power (VA), and speed. However, since a typical
bus transfer model may include 15 or more motors, it may not
be practical to validate individual motor models.

3.2 STUDY ZONE


When two or more interacting controls are identified, the
study zone encompasses those system components which
must be represented with adequate details to investigate the
interaction phenomenon. Since the frequencies of interest are
in the subsynchronous frequency range, the study zone is usually identified based on the criteria used for the study zone of
torsional oscillations, Section 2.2.1.

To establish the dead time and a range of the expected accuracy, it is recommended to perform a fast bus transfer test with
a few motors connected and simulating the test conditions using motor models based on the manufacturer supplied data.
Since measuring the transient variations in the motor shaft
torque is a complex task, it is suggested to monitor, simulate
and compare the following parameters:

3.3 DEVICE MODELS


bus instantaneous voltage

3.3.1 Generator Electrical System

individual motors instantaneous currents

If a turbine-generator controls system, i.e. governor system,


AVR, PSS, and its torsional mechanical modes do not participate in the interaction phenomenon, then the generator electrical system can be modelled as an ideal, fixed-frequency,
three-phase, voltage source behind a three-phase inductance.
Otherwise, the second-order model or the third-order model
for Section 2.2.2.1 should be used.

total instantaneous currents through the alternate source circuit


breaker
individual motors instantaneous power and apparent power
motor speed

errors can then be determined by comparison of the test data


with simulation results. A statistical measure of the expected
model accuracy may be based on the method of the root of the
sum of the squares of the individual errors (RSS). The expected error in the actual bus transfer analysis would be less than
the RSS of the errors derived due to the larger number of motors included. References [2.7, 2.8, 2.9] provide some test results which can be used as general guidelines to verify the
pattern of behaviour of the system variables due to the bus
transfer phenomenon.

3.3.2 Turbine-Generator Mechanical System


When the generator electrical system is represented either by
the second-order model or the third-order model, the shaft
system should be represented by the mass-spring-dashpot
model of Section 2.2.2.2. Otherwise, the shaft dynamics and
consequently its oscillatory modes can be ignored.
3.3.3 Power Transformer

Appendix A provides further information regarding fast bustransfer and typical time-domain simulation results.

When a generator is represented by a voltage source behind an


inductance, the generator step-up transformer is represented
by a series RL branch in each phase. Otherwise, the low-frequency transformer model of Section 2.2.2.3 should be used
to represent the transformer in the overall system model. In
general, the low-frequency transformer model is an adequate
representation of a power transformer for investigation of
controller interaction phenomenon. The harmonics generated
as a result of transformer saturation have much higher frequencies than those of controller interactions. Thus the saturation does not have any major role in the controller
interaction phenomenon.Transmission Line

3. CONTROL SYSTEM INTERACTIONS


3.1 DEFINITION
Closed-loop controls associated with various power system
apparatus, e.g. SVC controls, HVDC converter controls, controls of adjustable series capacitors, generator automatic voltage regulators (AVRs), and generator power system
stabilizers (PSSs) have natural oscillatory modes at frequencies in the subsynchronous frequency range of 1 to 35 Hz.
Depending upon the electrical distance between the apparatus, the associated closed-loop controls can interact and result
in either unsatisfactory operation of the device(s), sustained
oscillations, or even small-signal instability. Another type of
controller interactions is the interaction between a closed-

Per-phase equivalent- model is an adequate representation of


a line for investigation of the phenomenon of control interactions.

3-8

trolled current source and only the inverter station and its control system be represented in details, if the rectifier station is
not within the Study Zone.

3.3.4 Series and Shunt Capacitor Banks


Presence of series capacitors in a transmission line can alter
the level of controller interactions or even excite the interaction mode(s) [3.2]. Impacts of parallel (shunt) capacitors on
the controller interactions is significantly less than that of a series capacitor. Both series and shunt capacitors can be adequately represented by three-phase lumped capacitor banks
for investigation of controller interactions.

3.3.8 Static VAR Compensator (SVC)


A conventional SVC, which is composed of thyristor-controlled reactor (TCR) and fixed capacitor (FC), can interact
with an HVDC converter station or other SVCs through their
closed-loop controls and excite the phenomenon of controller
interaction. An SVC model for control interaction studies
should accurately represent the SVC and its control system in
the frequency range of 5 to about 45 Hz. The steady-state
continuous controls including all the auxiliary loops, e.g. SVC
voltage control and SSR damping control, must be represented in the simulation model. Further details of an SVC smallsignal model are available in Section 4.3.9.

3.3.5 Shunt Reactor


Similar to shunt capacitors, fixed, shunt inductors do not have
a major impact on controller interactions. Nevertheless, shunt
reactors are adequately represented by three-phase lumped inductances for investigation of controller interactions.
3.3.6 Loads

3.3.9 Generator Controls


Fixed Impedance model of loads within the study zone provides accurate representation of the loads for investigation of
controller interaction phenomenon. Very large load areas can
also be represented by an infinite bus with proper phase angle to draw the required power at the fundamental frequency.
The impacts of various load models on the phenomenon of
controller interactions have been neither adequately investigated nor reported in the literature.

Conventional synchronous generator controls, i.e. governor


system, AVR, and PSS are designed to perform corresponding tasks at very low frequencies (0.1 to 2.5 Hz), and are not
the prime cause of controller interactions. Thus the dynamics
of the generator controls often can be neglected for the investigation of controller interaction phenomenon. However, if
their presence in the overall system model is required, their
conventional low-frequency, linearized models would suffice.

3.3.7 HVDC Converter Station


Rectifier or inverter firing angle controls can interact with
other system controllers, e.g. SVC controls, and excite control
interaction phenomenon. Contribution of an HVDC converter station to the controller interaction phenomenon is primarily as a result of the natural oscillatory modes of its control
loop(s) and not due to the harmonics generated by the valve
switchings. If both inverter and rectifier are within the study
zone, both converter stations, the connecting dc link, and all
the associated controls must be represented in the study model. Further details on representation of each 12-pulse converter are given in Section 4.3.8.

3.3.10 Harmonic Filters


Harmonic filters of SVCs are adequately represented by
lumped RLC circuits. Similarly, ac side and dc side harmonic
filters of HVDC converter stations are represented by lumped
RLC circuits.
3.4 TEST SYSTEM
Figure 3.1 shows the recommended test system for the investigation of controller interactions [3.1] of multiple SVCs. Depending upon the operating conditions and parameters, the
voltage control loops of the SVCs can interact and exhibit
small-signal instability. Inclusion of control limits in the
model is not necessary since the control interaction constitutes a linear phenomenon and nonlinearities are not involved.
icomp in Fig. 3.1 is the total current of each TCR and the associated capacitor bank. The systems data and initial conditions are given in [3.3, 3.4].

All the steady-state continuous controls of rectifier and inverter stations, e.g. DC current control, DC voltage control, AC
voltage control or reactive power control, real power control,
and frequency control must be represented in the model. The
control model must adequately represent firing and synchronization schemes used for the converter values.

When the inverter station is not within the Study Zone, the inverter station and the dc line can be represented by an equivalent controlled voltage source, and only the rectifier station
and its controls be modelled in detail. Similarly, the rectifier
station and the dc line can be modelled as an equivalent con-

3-9

4.2 STUDY ZONE


Those system apparatus which either generate or interact with
the frequencies of interest must be represented in adequate details, and they identify the study zone. Also transmission lines
which connect the apparatus within the study zone must be
represented with adequate accuracy in the frequency range of
interest in the system model. The remainder of the system
which neither generates nor interacts with the harmonics can
be simplified and represented by its frequency dependent
equivalent model [4.3].
4.3 DEVICE MODEL
4.3.1 Generator Electrical Model

Fig. 3.1. Test systems for investigation of controller interaction phenomena

Rotating machines within the study zone do not contribute to


the harmonic interaction phenomenon and can be represented
by equivalent voltage sources behind fixed RL elements.

3.5 VERIFICATION OF SIMULATION RESULTS


Small-signal controller interactions also can be investigated
based on the linearized model of the system under investigation, using eigen analysis approaches [3.6, 3.7. 3.8]. Both
time-domain simulation and the eigen analysis of controller
interactions are conducted for qualitative comparison of the
results and their mutual verifications.

4.3.2 Turbine-Generator Mechanical System


dynamics do not play any noticeable role in the harmonic interaction phenomenon. Thus, the shaft model can be readily
discarded from the overall system model.
4.3.3 Power Transformer

4. HARMONIC INTERACTION AND RESONANCE


4.1 DEFINITION
Operation of power electronic converters, e.g. an HVDC converter station, is characterized by generation of current and/or
voltage harmonics. These harmonics are classified as characteristic and noncharacteristic harmonics. In contrast to characteristic harmonics, amplitudes and orders of
noncharacteristic harmonics cannot be accurately predicted
by conventional analytical techniques, e.g. Fourier analysis.
Time-domain simulation methods provide an alternative approach for the analysis of noncharacteristic harmonics. References [4.1] and [4.2] provide a comprehensive description
of the physical phenomena resulting in harmonic interactions.

The main concerns with the presence of noncharacteristic harmonics are (1) harmonic interactions and/or resonance [4.1],
and (2) the interference phenomenon [4.2].

Radio and telephone interference as a result of dc side harmonics of HVDC converters is a well known phenomenon.
Also, second and third harmonic instability of ac systems due
to harmonic modulation characteristic of HVDC converter
has been encountered in the existing installations [4.1].

Both stray capacitances and magnetic saturation characteristics of power transformers within the study zone can have significant impact on power system harmonics. The magnetic
saturation characteristic has a deterministic impact on the second harmonic instability and can be fairly represented by the
no-load V-I characteristic in the magnetization branch of the
transformer. The winding stray capacitances to the tank have
a noticeable effect on the interference phenomenon [4.2]. The
stray capacitance can be adequately modelled by a single capacitance from the winding terminal to the ground [4.2].
4.3.4 Transmission Lines
Transmission lines within the study zone are best represented
as distributed parameter lines including parameter frequency
dependency. However, if the frequency range of interest does
not cover high frequencies (more than 300 Hz), each transmission line can be represented by multiple sections.
4.3.5 Series and Shunt Capacitor Banks
Series and shunt capacitors have deterministic impacts on series and parallel resonant frequencies of the system and must
be represented in the overall system model for harmonic studies. Both series and shunt capacitors are adequately represented by lumped three-phase capacitor banks.

3-10

4.3.6 Shunt Reactor

Similar to series and shunt capacitors, shunt reactors also influence the system natural resonant frequencies and must be
represented in the system model. A shunt reactor is adequately represented by a three-phase lumped reactor bank.
4.3.7 Loads
Fixed Impedance model is a valid representation for loads
within the study zone, unless the load is known to have particular resonant frequency or generates particular harmonic(s)
which can affect the harmonic phenomenon of interest.
4.3.8 HVDC Converter Station
The HVDC converter station is one of the major
sources for generation of harmonics which cause interference
and/or instability of electrical power networks [4.1]. The
required model of an HVDC converter station for studying interference and harmonic interaction phenomena is the same as the
model described in Section 3.3.8.

capacitors, Fig. 4.2.


Magnetic saturation characteristics of converter transformers must be included in the model [4.4].

4.3.9 Static VAR Compensator (SVC)


Static VAR compensators have not been reported as a source
of interference phenomenon and harmonic interactions.
However, in the vicinity of HVDC converter stations and
FACTS devices, a static VAR compensator can aggravate
harmonic related issues [4.4]. The required SVC model for
time-domain investigation of harmonic problems is the same
as the model described in Section 3.3.9, except for the following differences:
Snubber circuits of each valve chain must be included in
the simulation model.

The model of valve firing circuitry must be capable of generating exact firing instants.
Operating point and parameter values of a SVC can readily
influence series/parallel resonant frequencies of a network
and consequently tune the system for resonant conditions, e.g.
second harmonic resonance [4.4]. The above model can also
be used for this class of resonant conditions which normally
occur at noncharacteristic harmonics generated by power
electronic circuits.
4.3.10 Generator Control
Automatic voltage regulator, power system stabilizer, and
governor system do not influence harmonic related problem.
Thus, their model can be excluded from the system model for
time-domain harmonic studies.
4.3.11 Harmonic Filters
SVC and HVDC harmonic filters must be modelled as described in Section 3.3.11.
4.4 TEST SYSTEM

Fig. 4.1. Lumped equivalent of the stray capacitances of a 12-pulse HVDC


converter and the converter transformers

Exact parameters of the snubber circuits of each valve


chain should be included in the model [4.2]. It should be
noted that in some transients programs, the exact parameters of snubber circuits cannot be used. Unrealistic
snubber circuits are required by these programs to avoid
numerical problems.
The model used for the valve firing circuitry should generate actual firing instants. Otherwise, the amplitudes and
orders of noncharacteristic harmonics will be noticeably
distorted as a result of improper firing instants [4.4].
Stray capacitances of the converter transformers, valve
structure, and smoothing reactor must be adequately represented in the system model [4.2]. The impact of stray
capacitances can be represented by a set of lumped

The HVDC-AC system of Fig. 4.3 is proposed as the test system for the investigation of harmonic interactions phenomena
and the second harmonic instability issues.

3-11

Fig. 4.1. HVDC-AC test systems for time-domain simulation of harmonic


interaction phenomena, and the second harmonic instability

The HVDC link is a 450-kV, 936-km, 2000-MW, 12-pulse,


bipole configuration. Each pole is equipped with ac side and
dc side filters. The inverter neutral is equipped with a neutral
filter. The rectifier neutral is solidly grounded close to the station. Parameters and control system of the Manitoba Hydros
Bipole-2 HVDC system [4.5, 4.6] are adopted for the test system of Fig. 4.3.

The rectifier ac system, Fig. 4.3, is composed of an equivalent


26-kV source which is connected to the rectifier ac bus
through a 26/235-kV transformer and a short 230-kV line.
The effective short circuit ratio (ESCR) of the rectifier ac system is 3.6.

modelling approach of [4.15] are reported in [4.14]. References [4.16, 4.17, 4.18] provide a comprehensive and fundamental description of the harmonic interaction phenomenon.
However, there are not that many measurements and investigation of the harmonic interaction phenomenon to establish a
method for verification of time-domain simulation studies.
Reference [4.19] introduces an alternative approach based on
frequency scanning method for identification of harmonic instabilities in HVDC systems. This approach may be used for
qualitative verification of digital time-domain simulation approach.

5. FERRORESONANCE
The inverter ac system consists of a 230-kV ac source which
is connected to the inverter station through a 500-kV, 832-km
transmission system. The transmission line is equipped with
240/525-kV Y - Y connected transformer at the source side.
The ac line is divided in three sections, Fig. 4.3. Each intermediate station is equipped with a 400 MVA capacitor bank
for voltage profile improvement. Loads #1, #2, and #3 are rated at 920-MVA, 400-MVA and 360-MVA respectively. The
inverter station is also equipped with an SVC which can adjust its reactive power from 180-MVAR inductive to 510MVAR capacitive. Electrical parameters of the inverter ac
system are given in [4.7]. The ESCR of the inverter ac side is
2.2.

References [4.8] and [4.9] provide various HVDC/ac benchmark models that also can be used for the analyses of harmonic interactions and resonance phenomena. The first HVDC
benchmark model [4.8] proposed by CIGRE WG 14-02 also
exhibits second harmonic resonance and can be adopted for
investigation of harmonic instability phenomenon. This system is less complicated as compared with that of Fig. 4.3.
Reference [4.10] provides a very simple circuit configuration
which exhibits instability due to switching characteristic of
thyristor-controlled reactor (TCR). A set of time-domain
simulations results of the test systems of Fig. 4.3 is given in
[4.4].
4.5 VERIFICATION OF SIMULATION RESULTS
There are several technical papers which deal with analysis
and measurement of noncharacteristic harmonics of HVDC
converter stations [4.2, 4.11, 4.12, 4.13]. The primary concern in these papers is the dc side triplen harmonics which
cause interference and not the second harmonic instability
problem. Reference [4.15] provides a modelling approach for
representation of a six-pulse converter with respect to the second harmonic for eigen analysis. Such eigen analysis approach can be used as an alternative technique for validation
of simulation results. Eigen analysis studies based on the

In this section, ferroresonance is introduced and a general


modeling approach is given. An overview of available literature and contributors to this area is provided. A simple case
of ferroresonance in a single phase transformer is used to illustrate this phenomenon. Three phase transformer core
structures are discussed. Ferroresonance in three phase
grounded-wye distribution systems is described and illustrated with waveform data obtained from laboratory simulations.
Representation of the study zone is discussed, modeling techniques are presented, and implementation suggestions are
made. Three case studied are presented. Transformer representation is critical to performing a valid simulation. The direction of ongoing research is discussed, and the reader is
advised to monitor the literature for ongoing rapid improvements in transformer modeling techniques.

5.1 INTRODUCTION TO FERRORESONANCE

Research involving ferroresonance in transformers has been


conducted over the last 80 years. The word ferroresonance
first appears in the literature in 1920 [5.7], although papers on
resonance in transformers appeared as early as 1907 [5.4].
Practical interest was generated in the 1930s when it was
shown that use of series capacitors for voltage regulation
caused ferroresonance in distribution systems [5.9], resulting
in damaging overvoltages.
The first analytical work was done by Rudenberg in the 1940s
[5.36]. More exacting and detailed work was done later by
Hayashi in the 1950s [5.17]. Subsequent research has been
divided into two main areas: improving the models used to
predict the behavior of the transformers, and studying ferroresonance involving transformers installed in power systems.

3-12

An understanding of the nonlinear parameters describing a


transformer core is prerequisite to dealing with ferroresonance. Swift [5.47] and Jiles [5.20] have provided insight into
transformer core behavior and the separation of hysteresis and
eddy current losses. Frame [5.15] and others have developed
piecewise-linear methods of modeling the nonlinearities in
saturable inductances.
Hopkinson [5.19] performed system tests and simulations on
the effect of different switching strategies on the initiation of
ferroresonance in three phase systems. Smith [5.38] categorized the modes of ferroresonance in one type of three phase
distribution transformer based on the magnitude and appearance of the voltage waveforms. Arturi [5.2] and Mork [5.29]
have demonstrated the use of duality transformations to obtain transformer equivalent circuits. Mork [5.27] and Kieny
[5.21] have shown that the theories and experimental techniques of nonlinear dynamics and chaotic systems can be applied to better understand ferroresonance and limitations
inherent in modeling a nonlinear system. Developments in
the near future are expected to be in the areas of developing
improved transformer models and applying nonlinear dynamics to the simulation of ferroresonance.

5.2 FERRORESONANCE IN A SINGLE PHASE TRANSFORMER


In simple terms, ferroresonance is a series "resonance" involving nonlinear inductance and capacitances. It typically
involves the saturable magnetizing inductance of a transformer and a capacitive distribution cable or transmission line connected to the transformer. Its occurrence is more likely in the
absence of adequate damping. A simple case of ferroresonance is presented here as an illustration.
When rated voltage is applied to an unloaded single phase
transformer, only a very small excitation current flows (Fig.
5.1). In this case, the 120-volt winding of a 120-240 volt 1.5
kVA dry-type transformer is energized, resulting in an exciting current, whose peak amplitude is 0.05 per unit. Referring
to the equivalent circuit shown, it is seen that this current consists of two components: the magnetizing current and the core
loss current. The magnetizing current, which flows through
the nonlinear magnetizing inductance LM, is required to induce a voltage in the secondary winding of the transformer.
The core loss current, flowing through RC, makes up the eddy
current losses and hysteresis losses in the transformer's steel
core.

2 0 0 .0

1 .0

1 5 0 .0

0 .7 5

1 0 0 .0

0 .5

5 0 .0

0 .2 5

0 .0

0 .0

-5 0 .0

-0 .2 5

-1 0 0. 0

-0 .5

-1 5 0. 0

-0 .7 5

-2 0 0. 0
0 .0 s

1 0 . 0m s

2 0 . 0m s

3 0 . 0m s

4 0 . 0m s

-1 .0
5 0 . 0m s

T IM E

Fig. 5.1. Unloaded single phase transformer with rated voltage applied.Solid
waveform is applied voltage; dashed waveform is exciting current

Although usually assumed linear, RC is dependent on voltage


and frequency. The excitation current contains high order odd
harmonics, due to transformer core saturation. RW and LL
are the winding resistance and winding leakage inductance,
respectively. They are assumed to be linear parameters. Their
magnitudes are relatively small compared to LM and RC and
so are usually ignored in no-load situations [5.3,5.24].
If a capacitor is placed between the voltage source and the unloaded transformer, ferroresonance may occur (Fig. 5.2). An
extremely large exciting current (1.92 per unit peak) is drawn
and the voltage induced on the secondary may be much larger
than rated (1.44 per unit peak). The high current here is due
to resonance between CS and LM; ferroresonance in most
practical situations results in smaller exciting currents. Any
operating "modes" which result in a significantly distorted
transformer (inductor) voltage waveform are typically referred to as ferroresonance, although the implication of resonance in a classical sense is arguably a misnomer. Even
though the "resonance" occurring does involve a capacitance
and an inductance, there is no definite resonant frequency,
more than one response is possible for the same set of parameters, and gradual drifts or transients may cause the response
to jump from one steady-state response to another.
High-order odd harmonics are characteristic of the waveforms, whose shapes might be conceptually explained in
terms of the effective natural frequency 1 LMCS as LM goes
in and out of saturation. Steep slopes (fast changes) occur
when LM is saturated, and flat slopes occur when LM is operating in its linear unsaturated region.
Due to nonlinearity, two other ferroresonant operating modes

3-13

are possible, depending on the magnitudes of source voltage


and series capacitance. In this case, all modes are seen to produce periodic voltage waveforms on the transformer secondary [5.26,5.29]. In general, gradual changes in source voltage
or capacitance will cause state transitions. A reversal to conditions that caused a transition will not reverse the transition,
due to nonlinearity of LM [5.36]. Transients can also trigger
transition from mode to mode.
In modern terms, these jumps are referred to as bifurcations
[16,27,29,45], and may be better understood by applying the
theory of nonlinear dynamics and chaos. A long-used intuitive explanation of these jumps, based on a graphical method,
is given by Rudenberg [5.36]. However, this method is not a
good analytical tool since it is based only on the fundamental
frequency and neglects harmonics.
Damping added to the circuit will attenuate the ferroresonant voltage and current. Some damping is always
present in the form of resistive source impedance, transformer
losses, and also corona losses in high voltage systems, but most
damping is due to the load applied to the secondary of the transformer.

Ferroresonance can lead to heating of transformer, due to high


peak currents and high core fluxes. High temperatures inside
the transformer may weaken the insulation and cause a failure
under electrical stresses. In EHV systems, ferroresonance
may result in high overvoltages during the first few cycles, resulting in an insulation coordination problem involving frequencies higher than the operating frequency of the system.
Because of nonlinearities, analytical solution of the ferroresonant circuit must be done using time domain methods. Typically, a computer-based numerical integration method is
applied using time domain simulation programs such as the
EMTP.

5.3 MAGNETIC BEHAVIOR OF THREE PHASE TRANSFORMERS


It is incorrect to assume that a three phase transformer
core is magnetically equivalent to three single phase transformers, i.e. that the three phases have no direct magnetic coupling.
Such an assumption can lead to serious errors, especially if one
is investigating a transformer's behavior under transient or
unbalanced conditions.

Fig. 5.2. Same transformer as in Fig. 5.1, fed through a 75F capacitance,operating in ferroresonance. Solid waveform is terminal voltageof
transformer; dashed waveform is the current.

Damping added to the circuit will attenuate the ferroresonant


voltage and current. Some damping is always present in the
form of resistive source impedance, transformer losses, and
also corona losses in high voltage systems, but most damping
is due to the load applied to the secondary of the transformer.
Therefore, a lightly-loaded or unloaded transformer fed
through a capacitive source impedance is a prime candidate
for ferroresonance.
This elementary type of ferroresonance is similar to that
which occurred in the series capacitor compensated distribution systems of the 1930s. It can also occur, from different
sources of capacitance, in today's single phase distribution
transformers and voltage instrument transformers [5.1,5.18].
It can also occur in series-compensated transmission lines.

Fig. 5.1. Core configurations commonly used in three phase transformers.Only one set of windings is shown.

The only type of core that displays magnetic characteristics


similar to three single phase transformers is the triplex core.
Although the cores share the same tank, they are magnetically
isolated (except for leakage fluxes). Core laminations can be
stacked or wound. Zero sequence fluxes will circulate individually in each core, and tank heating is not a problem. Under normal balanced operation, exciting currents in each
phase are identical, except for their 120 shift in phase angle.
All of the other core configurations provide direct flux linkages between phases via the magnetic core. Simply stated, applying a voltage to any one phase will result in voltages being

3-14

induced in the other phases (only in the adjacent phase(s) in


the case of the five-legged wound core). Further, the degree
of saturation in each limb of the core affects the way flux
flows divide. The apparent reluctance seen by each of the
windings changes depending on the degree of saturation in
each of the limbs of the transformer core. Therefore, exciting
currents vary from phase to phase, even under balanced operation. A brief discussion of each of these core types follows:
Core-form transformers require the least amount of core material to manufacture. Laminations are stacked. Their worst
problem is that unbalanced operation results in zero sequence
fluxes which cannot circulate in the core. These zero sequence fluxes are forced through the insulation surrounding
the core and through the transformer tank. Tank steel is not
laminated like the core is, so eddy currents can heat the tank
and cause damage. Therefore, this type of core should only be
used where load currents are balanced.
The shell-form core provides a magnetic path for zero sequence flux, and is much better-suited for unbalanced operation. Laminations are stacked. There is a large base of
transformers with this type of core (about half of the installed
three phase power transformers in the US).

lines, capacitor banks, coupling capacitances between double


circuit lines or in a temporarily-ungrounded system, and voltage grading capacitors in HV circuit breakers. Other possibilities are generator surge capacitors and SVCs in long
transmission lines. Due to the multitude of transformer winding and core configurations, system connections, various
sources of capacitance, and the nonlinearities involved, the
scenarios under which ferroresonance can occur are seemingly endless [5.5].
System events that may initiate ferroresonance include single
phase switching or fusing, or loss of system grounding. The
ferroresonant circuit in all cases is an applied (or induced)
voltage connected to a capacitance in series with a transformer's magnetizing reactance.
Fig. 5.4 gives three examples of ferroresonance occurring in a
network where single phase switching is used. A wye-connected capacitance is paralleled with an unloaded wye-connected transformer. The capacitance could be a capacitor
bank or the shunt capacitance of the lines or cables connecting
the transformer to the source. Each phase of the transformer
is represented by jXm, since ferroresonance involves only the
magnetizing reactance.

The four-legged core also provides a magnetic path for zero


sequence flux. This type of core design is not very common.
It is the only type of core whose outer phases do not exhibit
like behavior.
The five-legged stacked core also provides a magnetic path
for zero sequence flux, but has a more symmetric core. This
type of core is often specified where a low-profile is desirable
for shipping or for visual appearance in urban substations.
The five-legged wound core is made up of four concentrically-laminated cores. The unique feature of this core is that only
adjacent phases are directly linked via a magnetic path. Assuming no flux leakage between cores, the two outer winding
assemblies are not magnetically coupled. Tank heating is
minimized, since there are zero sequence flux paths in the
core. Because of its low cost, this type of transformer core is
widely used in distribution systems.
The winding configuration used does not have any effect on
the transformer core model. Delta, wye, or zig-zag winding
connections are made outside of the model of the core equivalent. However, behavior of the transformer is strongly dependent on the winding configuration.
5.4 FERRORESONANCE IN THREE PHASE SYSTEMS
Ferroresonance in three phase systems can involve large power transformers, distribution transformers, or instrument
transformers (VTs or CVTs). The general requirements for
ferroresonance are an applied (or induced) source voltage, a
saturable magnetizing inductance of a transformer, a capacitance, and little damping. The capacitance can be in the form
of capacitance of underground cables or long transmission

Fig. 5.1. Three examples of ferroresonance in three phase systems.

If one or two poles of the switch are open and if either the capacitor bank or the transformer have grounded neutrals, then
a series path through capacitance(s) and magnetizing reac-

3-15

tance(s) exists and ferroresonance is possible. If both neutrals


are grounded or both are ungrounded, then no series path exists and there is no clear possibility of ferroresonance. In all
of these cases, the voltage source is the applied system voltage. Ferroresonance is possible for any of the core configurations of Fig. 5.3 (even for triplexed or a bank of single phase
transformers).

Depending on the type of transformer core, ferroresonance


may be possible even when there is no obvious series path
from the applied voltage through a capacitance and a magnetizing reactance. This is possible with three phase core types
which provide direct magnetic coupling between phases,
where voltages can be induced in the open phase(s) of the
transformer. To illustrate, a grounded-wye to grounded-wye
transformer typical of modern distribution systems is considered. A recent survey in the US showed that 79% of underground rural distribution systems use this configuration, so
ferroresonance problems in this type of installation are of special interest [5.23,5.25,5.40,5.41]. A simplified schematic of
such a system is shown in Fig. 5.5. The distribution line is
represented by its RLC pi equivalent, with no interphase coupling. Three phase circuit breakers and gang-operated
switches are used at the substation where distribution lines
originate, but single phase switching and interrupting devices
are used outside of the substation.

Whether ferroresonance occurs depends on the type of


switching and interrupting devices, type of transformer, the
load on the secondary of the transformer, and the length and
type of distribution line. A long underground line is much
more capacitive than a short overhead line. However, due to
nonlinearities, increased capacitance does not necessarily
mean an increased likelihood of ferroresonance. Operating
guidelines based on linear extrapolations of capacitance may
not be valid. Also, as mentioned previously, the smaller the
load on the transformer's secondary, the less the system damping is and the more likely ferroresonance will be. Therefore,
a highly capacitive line and little or no load on the transformer
are prerequisites for ferroresonance. Binary loads (either full
load or no load) such as irrigation, are essentially zero most of
the time and cannot be relied upon to damp ferroresonance.
Ferroresonance is rarely seen provided all three source phases
are energized, but may occur when one or two of the source
phases are lost while the transformer is unloaded or lightly
loaded. The loss of one or two phases can easily happen due
to clearing of single phase fusing, operation of single phase
reclosers or sectionalizers, or when energizing or deenergizing using single phase switching procedures.
If one of the three switches of Fig. 5.4 were open, only two
phases of the transformer would be energized. If the transformer is of the triplex design or is a bank of single phase
transformers, the open phase is simply deenergized and the
energized phases draw normal exciting current. (Existence of
capacitor banks or significant phase to phase capacitive coupling could still result in ferroresonance, but that possibility is
not addressed here).
However, if the transformer is of the three-, four- or fivelegged core type, a voltage is induced in the "open" phase.
This induced voltage will "backfeed" the distribution line
back to the open switch. If the shunt capacitance is significant, ferroresonance may occur. The ferroresonance that occurs involves the nonlinear magnetizing reactance of the
transformer's open phase and the shunt capacitance of the distribution line and/or transformer winding capacitance. It has
been shown that the ferroresonant circuit is a series combination of the shunt cable capacitance and the magnetizing inductance of one of the transformer's wound cores [5.23]. The
equivalent circuit for this transformer is derived later in this
paper.

Fig. 5.2. Typical distribution system supplying a three phaseload through a


grounded-wye to grounded-wye transformer.

Either overhead lines or underground cables connect transformers to the system. Cables have a relatively large shunt capacitance compared to overhead lines, so this type of
ferroresonance most often involves underground cables, but is
also possible due solely to transformer winding capacitance.
Three phase or single phase transformers can appear at the
end of a distribution line or at any point along the line. Three
phase transformers may have any one of the several core types
discussed in the previous section.

An example of ferroresonant voltage and current waveforms


occurring under this scenario is shown in Fig. 5.6. In this
case, rated voltage was applied to X2 and X3, while X1 was
unenergized and had 9F attached to simulate a length of underground distribution cable.
Whether in ferroresonance or not, this backfeed situation can
be dangerous, as operating personnel may assume that the
load side of the open switch is deenergized and safe to work
on, when in fact a high voltage is present. Also, it can be seen
that single phase loads connected along this backfed phase

3-16

will continue to be supplied, although with dangerously high


or low voltage levels and with poor power quality.
Therefore, use of single phase interruption and switching
practices in systems containing the five-legged core transformers is the main operating tactic responsible for initiating
ferroresonance. Replacement of all single phase switching
and interrupting devices with three phase devices would eliminate this problem, although economics discourages such
large scale upgrades. An alternate solution would be to replace all five-legged core transformers with single phase
banks or triplex designs wherever there is a small load factor.
System wide operation and design implications of this problem have been more fully addressed in prior work [5.25].

Fig. 5.1. Measurement of ferroresonance in a three phase grounded-wye to


grounded-wye five-legged core transformer. Voltage waveform is solid; current waveform is dashed.

The "higher energy modes" [5.1] of ferroresonance involving


relatively large capacitances and little damping can produce a
nonperiodic voltage on the open phase(s). These voltage
waveforms can be quite similar to those of Duffing's equation
[5.45], which describes a nonlinear forced oscillator commonly used to illustrate the behaviors of a nonlinear dynamical
system. Transitions between periodic and nonperiodic modes
occur due to gradual changes in circuit parameters or to transients. And as with Duffing's equation, initial conditions determine the mode that operation stabilizes in after the
transients die down.
The recognition that ferroresonance is a nonlinear and sometimes chaotic process opens up many possibilities. The newly-developed techniques for analysis of nonlinear dynamical
systems and chaos are being evaluated for use with ferroresonance [5.27,5.21]. Use of geometric graphical methods like
phase plane projections and Poincar sections can be applied
to obtain a better understanding of ferroresonance.

5.5 NONLINEAR DYNAMICS AND CHAOS APPLIED


TO FERRORESONANCE

Ferroresonant circuits can be analyzed as damped nonlinear


systems driven by sinusoidal forcing function(s) [5.27]. The
nonlinear behavior of ferroresonance falls into two main categories. In the first, the response is a distorted periodic waveform, containing the fundamental and higher-order odd
harmonics of the fundamental frequency. The second type is
characterized by a nonperiodic, or chaotic, response. In both
cases the response's power spectrum contains fundamental
and odd harmonic frequency components. In the chaotic response, however, there are also distributed frequency harmonics and subharmonics. A good conceptual introduction to
chaos and nonlinear dynamics is given by [5.16], and a good
theoretical introduction can be found in [5.45].
At least 2 different periodic responses are possible for
a single phase transformer [5.26], similar to that of Fig.5.1. Ferroresonance in the above three phase five-legged core distribution transformer can be periodic or nonperiodic. "Lower energy
modes" [5.1] (involving relatively low energy oscillations
between the inductance and capacitance, similar to the waveforms shown in Fig. 5.5) produce periodic voltages on the secondary. Some of the periodic modes of ferroresonance may
contain subharmonics, but still have strong power frequency
components, but take longer than one fundamental cycle to
repeat. This occurs more typically for very large values of C.

5.6 MODELLING AND ANALYSIS OF FERRORESONANCE


5.6.1 Overview
Ferroresonance has never been well-understood. Therefore,
there is a great deal of misinformation on ferroresonance in
the literature. A good example of this concerns the application of grounded-wye to grounded-wye five-legged core distribution transformers. As recently as 1989, specification of
this type of transformer was recommended to eliminate or
minimize the possibility of ferroresonance [5.14,5.35]. This
misinformation is gradually being corrected [5.25,5.32], but
engineers must be cautious and continue to update themselves.
Efforts in past years seem focused on refining equivalent circuit models for transformers and performing simulations using a transient circuit analysis program such as EMTP.
Although these programs use fairly robust methods of numerical integration, such as the trapezoidal rule, results are only
as good as the models used (and the initial conditions if the
onset of ferroresonance is a concern). Simulation results have
a great sensitivity to the model used and errors in nonlinear
model parameters. Unfortunately, determining the model's
nonlinear parameters is probably the biggest modeling difficulty. Three phase transformer modeling has not progressed
as far as single phase modeling. A different model is required
for each type of core, and a different means of determining the
model parameters.
Ideally, use of a correct transformer model would allow an engineer to simulate situations where ferroresonance is likely.

3-17

Simulation results could then be used to avoid this problem


when designing a distribution system. Difficulties in determining an adequate model and in simulating every possible
combination of initial condition and transient make prediction
less than certain.
5.6.2 The Study Zone
Parts of the system that must be simulated are the source impedance, the transmission or distribution line(s), the transformer, and any capacitance not already included. Source
representation is not generally critical. Unless the source contains nonlinearities, it is sufficient to use the steady-state thevenin impedance and open-circuit voltage. The distribution
line or transmission line can be assumed to be an RLC coupled pi-equivalent, cascaded for longer lines. Shunt or series
capacitors may be represented as a standard capacitance, paralleled with the appropriate dissipation resistance. Stray capacitance may also be incorporated either at the corners of an
open-circuited delta transformer winding or midway along
each winding. Other sources of capacitance are transformer
bushings and interwinding capacitances, and possibly busbar
capacitances.
One of the most critical parts of any ferroresonance study is
the transformer model. The transformer contains the nonlinearities, and modeling results are most sensitive to correct
representation of magnetic saturation and core loss. The rest
of this discussion focuses mainly on how the transformer
should be modeled. Many are dissatisfied with the transformer modeling capabilities in today's modeling packages. There
has been much discussion recently as to what improvements
can be made in modeling techniques [5.6,5.13,5.46].
5.6.3 More on Single Phase Transformer Models And
Parameters
Single phase transformers are typically modeled as shown in
Fig. 5.2. This model is topologically correct only for the case
where the primary and secondary windings are not concentrically wound. LL2 is essentially zero for concentric coils. Errors in leakage representation are not significant, however,
unless the core saturates. Obtaining the linear parameters for
this 2-winding transformer may be difficult. Short circuit
tests give total impedance (R1 + R2) + j(X1 + X2). A judgement must be made as to how it is divided between the primary and secondary windings.
If the transformer has three or more windings, the Rs and Xs
for the individual windings of each phase may be separated.
Sometimes one of Xs is negative, but this will not usually
cause a problem in the time domain transient simulation. This
approach satisfactorily separates the winding resistances, but
may not correctly account for mutual inductive coupling. To
solve this problem, a coupled L representation for the short
circuit inductances is recommended [5.11]. Binary short circuit (shorting two windings at a time while leaving all others

open) tests for all possible combinations of windings must be


performed to obtain the inductance matrix. Additional developments are still needed, however, since the core equivalent
cannot be correctly incorporated with this representation (the
only place it can be connected is on one of the external transformer terminals).
Model performance depends mainly on the representation of
the nonlinear elements RC and LM. RC has traditionally been
modeled as a linear resistance. Such a core loss representation, if it represents the average losses at the level of excitation being simulated, may in fact yield reasonable results.
Due to eddy current losses and hysteresis losses being nonlinear, calculation of a linear core loss resistance RC gives different values for each level of excitation. Using the value of
RC closest to rated voltage may be a good enough estimate.
Past research has shown low sensitivities to fairly large
changes in RC [5.29] for single phase transformers, but a high
sensitivity for three-phase cores.
LM is typically represented as a piecewise linear -i characteristic [5.22], or perhaps as a hysteretic inductance
[5.15,5.20,5.33]. The linear value of LM (below the knee of
the curve) does not much affect the simulation results [5.8],
although great sensitivities are seen for the shape of the knee
and the final slope in saturation.
Factory test data provided by the transformer manufacturer is
often insufficient to obtain the core parameters. Open circuit
tests should be made for 0.2 to 1.3 pu (or higher) instead of the
typical 0.8 to 1.14 pu range. It is important that open circuit
tests be performed for voltages as high as the conditions being
simulated, or the final -i slope of LM must be guessed. Some
thought should be given to the requirements of test reports
when specifying new transformers.
A method proposed by Dommel [5.11,5.22] is often used to
convert the RMS V-I open circuit characteristic to the -i characteristic of LM. To successfully use this method, the first
(lowest) level of excitation must result in sinusoidal current,
or errors will result in the form of an S-shaped -i curve. Also,
the V-I characteristic must extend as high as the highest voltage that will be encountered in the simulation. An extension
on this method has been proposed to obtain a nonlinear v-i
representation of RC [5.31], but the resulting flux-linked vs.
IEX loop does not seem to correctly represent the core losses.
Modern low-loss transformers have comparatively large inter-winding capacitances which can affect the shape of the excitation curve [5.47]. This can cause significant errors when
the above method is being used to obtain core parameters. In
these cases, factory tests must be performed to get the -i curve
before the coils are placed on the core. A means of removing
the capacitive component of the exciting current has also been
developed [5.29].
5.6.4 Three Phase Transformer Models And Model Parameters

3-18

For three phase transformers, it is possible to make a simplified model by connecting together three of the above single
phase models. If this is done, a triplex core configuration is
assumed (see Fig. 5.3). A delta-wye transformer of this type
is shown in Fig. 5.7. It is postulated that zero sequence (homopolar) effects are included almost entirely by the leakage
inductance of the delta windings [5.11,5.22].

The model might be improved by using a coupled inductance


matrix to model the short circuit characteristics of three phase
transformers. Binary short circuit tests involving all windings
of all phases must be performed. Problems can arise for RMS
short circuit data involving windings on different phases,
since the current may be nonsinusoidal. A problem also exists
with connecting the core equivalent. Three single phase core
equivalents are often attached to the windings closest to the
core, and may provide acceptable results in some cases, especially in the case of the three-legged stacked core. Questions
exist as to this method's validity, especially depending on the
type of core being analyzed. The most important question is,
however, what is the topology of the core equivalent? A
method of obtaining topologically correct models is presented
in the next section.
5.6.5 Use of Duality Transformations to Obtain Equivalent
Circuits
This method is based on the duality between magnetic and
electrical circuits. It was originally developed by Cherry
[5.10] in 1949 and Slemon [5.37] in 1953. Using duality
transformations, equivalent circuit derivations reduce to exercises in topology. These methods did not receive much attention at first, presumably since computers were not available.
Researchers have recently begun to use duality to provide
equivalent circuit models which are more topologically correct [5.2,5.29,5.30,5.34,5.39,5.42,5.44]. This approach results in models that include the effects of saturation in each
individual leg of the core, inter-phase magnetic coupling, and
leakage effects. Results are promising, and ongoing work
seems most focused on developing and improving dualitybased models.

Fig. 5.1. Model of a delta to wye transformer bank made up of three single
phase transformer models [22].

If the transformer does not have any delta windings, zero sequence effects may be included by adding a set of delta windings to the model whose total leakage impedance is equal to
the transformer's zero sequence inductance. This may work
for a three-legged core transformer that has an air path for
zero sequence flux, but is highly questionable in the case of
transformers having a saturable zero sequence flux path.
Factory three-phase excitation test reports will not provide the
information needed to get the magnetizing inductances for
this model. Note that standards require the exciting current to
be stated as the "average" value of the RMS exciting currents
of the three phases. Unless it is a triplexed core, this is meaningless, since the currents are not sinusoidal and they are not
the same in every phase. Therefore, the waveforms of the applied voltage and exciting currents in all three phase should be
given by the manufacturer for all levels of applied voltage.

To illustrate the method, a duality derivation used to obtain


the model for the five-legged wound core transformer [5.28]
is done here and a case study is presented later in this paper.
A section view of this type of transformer is shown in Fig. 5.8.
The magnetic flux paths and assumed leakage flux paths are
labeled. In the equivalent magnetic circuit, windings appear
as MMF sources, leakage paths appear as linear reluctances,
and magnetic cores appear as saturable reluctances.

The next step is the duality transformation itself. Using the


symbol to denote the transformation between electrical and
magnetic circuit elements, MMF I (MMF = NI), d /dt V, and
L (L = N2/ ). In terms of topology, meshes and nodes in the
magnetic circuit transform into nodes and meshes respectively in the electrical circuit. The resulting equivalent circuit is
given in Fig. 5.9.
To make the model practically useful, each current source resulting from the transformation has been replaced with an ideal transformer to provide primary-to-secondary isolation and
coupling to the core, while preserving the overall primary to
secondary turns ratio. Turns ratios are chosen so that core pa-

3-19

rameters are referenced to the low voltage windings. The portion of the model inside the coupling transformers represents
the core and leakages. Winding resistance and interconnection of the windings appears external to the coupling transformers. The advantage to this is that the derived core
equivalent can be used independently of winding configuration (delta, wye, zig-zag, etc.). Winding resistance, core losses, and capacitive coupling effects are not obtained directly,
but can be added to this topologically-correct equivalent electrical circuit.

voltage transformer (VT) connected to the delta side of the


power transformer, ferroresonance can occur (see figure).
The capacitance in this case comes from whatever "stray"
coupling capacitance exists between the delta windings and
earth. Adding a resistive burden to the VT can eliminate the
problem.

Fig. 5.1. Duality derived equivalent circuit with current sources replaced by
ideal coupling transformers. Winding resistances have also been added

Fig. 5.1. Development of magnetic circuit for grounded-wye togroundedwye five-legged wound core transformer. At top, transformer core sectional
view used as a basis for duality derivation. Leakageflux paths are labeled.
Bold dividing lines mark division in corereluctances. Equivalent magnetic
circuit is shown at bottom.

Tests have been developed to determine the parameters for


this model [5.28].
5.7 CASE STUDIES
5.7.1 Case Study #1: VT Ferroresonance on Floating Systems
It is possible that parts of a power system can be operated for short times without system grounding. One common
example is the no-load energization of the wye side of a wye to
delta power transformer.

The delta side will "float" with respect to earth, until some
load or other source of grounding is connected. If there is a

A recent problem occurring in a 50-kV network in the Hafslund area near Moss, Norway, serves as an excellent example
[5.18]. The clearing of a short circuit removed the only remaining source of grounding on the system. After the fault
was cleared, the only remaining zero sequence impedance
was due to capacitive coupling to earth. After operating in
this way for only 3 minutes, ferroresonance had destroyed 72
of the VTs used for measurement and protective relaying. All
72 of the damaged VTs were from the same manufacturer.
The VTs of two other manufacturers that were also in service
during this time were not damaged.
Fig. 5.10 shows the typical VT arrangement used in this system. The VTs have two low voltage windings. The secondary
is used for measurement and protective relaying purposes.
The burden on that winding has a very high impedance and its
effects can be ignored when considering ferroresonance. It is
the tertiary windings which are shown in Fig. 5.10. These
windings are connected in open delta and loaded with a damp-

3-20

ing resistance RO. The purpose of this damping resistance is


to damp out ferroresonance, and this design has been commonly used for many years.

EMTP model. System positive and negative sequence impedances were found to be very small compared to the primary
impedances of the VTs, and could be neglected. The zero
sequence impedance ZO consists almost entirely of the stray
capacitance of the floating system, and is therefore very important. Values of ZO varied from 0.6 - j219 to 0.2 - j221 , depending where in the system. ZO therefore becomes the only system
impedance needed in the model, and the positive sequence voltage sources can be modeled as stiff sources. The core losses of
the VTs were also neglected, their values being much higher
than the damping resistance RO.

Fig. 5.2. Typical VT connection in 50-kV Norwegian subtransmission system.

Since some of the VTs were damaged and the others weren't,
the VTs of different manufacturers obviously must have different characteristics. The problem at Hafslund therefore
forced a re-evaluation of the specification and application of
voltage transformers. EMTP was used to simulate the system
conditions that caused the VT failures. VT model parameters
were obtained from the manufacturers. Parameters are shown
in Table 1. Saturation characteristics were calculated based
on core material B-H data, core dimensions, and number of
primary turns. Data for the damaged VTs are listed as VT #1.

Fig. 5.3. Comparison of the saturation characteristics of the three VTs. Note
the much lower saturation level of VT #1, the ones that were damaged.

The designed flux densities BM at rated voltage vary. As a


more uniform basis of comparison, the flux densities were
converted to flux-linked values (Fig. 5.11). Note that VT #1
will saturate out at lower levels than the other VTs, and one
might guess this to be one of the reasons these failed and the
others didn't. But this can only be confirmed from simulation
results.

RP

XP

XT

N 1 :N 3

B M AX

VT # 1

32 50 6

2 500 6

0.0 1 6

20 k:2 3

1. 05T

VT # 2

32 18 6

3 094 6

0.0 1 6

~3 6k:42

0. 77T

VT # 3

75 88 6

4 833 6

0.0 1 6

25 k:2 9

0. 83T

Table 1: Linear parameters used to model the VTs at Hafslund

Fig. 5.12 shows the reduced equivalent used in the

Fig. 5.4. Reduced system equivalent, neglecting line impedances and lumping all VTs in each phase into an aggregate jXM.

Many simulations were run, with various combinations of


VTs and values of R0. It was found that ferroresonance occurred in most cases where RO was set to the 60 value typically used in system design. It was also seen that the high
magnetizing currents drawn by VT #1 while in ferroresonance
caused high IR losses in the windings, which thermally destroyed those VTs. If all of the VTs from manufacturer #1
were replaced with different VTs and if RO was reduced to 10
, ferroresonance would not occur. It was therefore recommended that the failed VTs be replaced with those of either
VT #2 or VT #3. A decrease in the value of RO standardly

3-21

being used was also recommended.

5.7.2 Case Study #2: Ferroresonance in Distribution Systems


This case involves the verification of the 75-kVA five-legged
wound-core distribution transformer model developed earlier.
Ferroresonance was staged on the secondary windings in the
laboratory. Balanced 3-phase voltage was applied to the secondary windings, and then one or two phases of the supply
were removed and replaced with various values of shunt capacitance. Scenarios investigated were loss of one source
phase to the center or an outer winding, and loss of two source
phases to either the two outer windings or to the center winding and one outer winding [5.27].
Measured waveforms were then compared to EMTP simulations. The transformer equivalent circuit used was essentially
that of Fig. 5.9. Details of model development and parameter
values are given in [5.29].
Since many ferroresonant modes are possible, bifurcation
simulations were first run. A bifurcation is essentially a jump
from one mode of ferroresonance to another. A simulation
technique was developed to very slowly ramp the capacitance
[5.12,5.28] and record jumps from one mode to another. Fig.
5.13 gives one bifurcation diagram for the case where a
ramped capacitance is connected to unenergized winding X1
and rated positive sequence voltage is applied to X2 and X3.
Due to nonlinearities, it is important to ramp the capacitance
both upward and downward, to ensure that as many ferroresonant modes are discovered as possible.

correctly predicts the existence of all modes of ferroresonance at


the correct values of capacitance. The actual waveforms simulated are very close for the periods one, two, and three. Period
five is generally correct, with slightly lower than actual peak
amplitudes predicted. The chaotic response predicted is slightly
higher than actual. The model used a simplistic linear resistance
to represent the core losses of each core. The model's accuracy
could be improved by implementing a more correct (complex)
core loss representation.

5.7.3 Case Study #3: Ferroresonance of Autotransformer


This case is taken from the Ontario Hydro system
where the Cataraqui 230/115-kV autotransformer T2, fed by line
X3H, was experiencing ferroresonance upon deenergization of
line X3H and the 115-kV bus (Fig. 5.15). The deenergizing circuit breaker was also experiencing a high recovery voltage. It
was deduced that capacitive coupling between line X3H and the
still-energized lines X4H and X522A was driving the autotransformer into ferroresonance. Damping resistors were added to
the tertiary of T2, but it was not certain whether the resulting
damping was sufficient to limit the duration of ferroresonance
and the related recovery voltage.

Using the bifurcation diagram as a road map, ferroresonance


for capacitances of 5F, 10F, 22.5F, 14.6F, and 18 F
was simulated. This corresponds to waveforms of periods 1,
2, 3, 5, and chaotic (nonperiodic). "Period 3" simply means
that the waveform takes three periods of the forcing function
to repeat -- it contains 1/3 harmonics.

Fig. 5.1. Period 3 ferroresonance, 22.5F connected X1

Fig. 5.2. Ontario Hydro 230-kV system. Ferroresonance involving line X3H
and connected transformer at Cataraqui Transformer Station.
Fig. 5.1. Sample bifurcation diagram. Shunt capacitance on X1 is ramped
from 0 to 30 F. Blurred areas correspond to chaos.

Fig. 5.14 shows the result of one of the EMTP simulations and compares it to the actual measurements. The model

Several EMTP simulations were run, with Y-connected resistive loads of zero, 133 kW/phase, and 266 kW/phase attached
to the tertiary of T2. In each case, the 115-kV breaker of T2
was assumed to open last. Two double-circuit 230-kV lines,
an existing 500-kV line, and a future 500-kV line were includ-

3-22

ed in the corridor, resulting in an 18-phase coupled-circuit


transmission equivalent (Fig. 5.16).
Fig. 5.17 shows the circuit breaker recovery voltage
for one of the cases.
It is interesting to note that a 133-kW/phase load did an
effective job of damping ferroresonance in T2, but resulted in a
higher recovery voltage than no damping at all. The circuit
breaker was marginally able to handle the recovery voltage
when the load was doubled to 267 kW/phase. Simulations were
also performed for deenergization of T1, with similar but less
severe behaviors noted. Recommendations were made to add
267 kW/phase loads to both transformers, and add surge arresters to the high and low voltage terminals of both transformers.

3.0E+5
CATARAQUI T2 FERRORESONANCE
EXTERNAL DAMPING = 133 kW / phase

2.0E+5
1.0E+5
0.0
-1.0E+5
-2.0E+5
-3.0E+5

402 kV peak
-4.0E+5
0.0s

100.0ms

200.0ms

300.0ms

400.0ms

TIME

Fig. 5.4. Cataraqui (T2) Autotransformer Ferroresonance. HV terminal voltage on Phase C is 2.0 per unit, with 133 kW/phase of damping.

5.8 RECOMMENDATIONS
Is seen that many different types of ferroresonance can and
do occur. Because of the nonlinear nature of ferroresonance,
it is difficult to predict if and where it might next occur. The
power system engineer should be aware, however, that it is
possible for lightly-loaded transformers operating in the presence of source or shunt capacitance to experience ferroresonance. Capacitance can be present in the form of cables,
series or shunt capacitor banks, or even stray capacitances in
inadequately-grounded portions of the system.

Fig. 5.3. Sequence of development of the transmission right-of-way

It is interesting to note that a 133-kW/phase load did an


effective job of damping ferroresonance in T2, but resulted in a
higher recovery voltage than no damping at all. The circuit
breaker was marginally able to handle the recovery voltage
when the load was doubled to 267 kW/phase. Simulations were
also performed for deenergization of T1, with similar but less
severe behaviors noted. Recommendations were made to add
267 kW/phase loads to both transformers, and add surge arresters to the high and low voltage terminals of both transformers.

Transient simulations are helpful in confirming or predicting the likelihood of ferroresonance, but only if a correct model is used. Per phase simulations of three phase systems will
not give correct results, due to various possible transformer
core configurations and winding connections. A complete
three phase model must be used. Therefore, the key to transient modeling is use of the proper transformer model. Development and use of acceptable transformer models should be a
priority task.
The development of improved topologically correct models is
a significant advancement, but model performance still depends on improving the way in which the cores are represented. Transformer core configuration must be considered and
saturation characteristics must be accurately known to operating levels well above rated voltage.
At this time, it is seen that modeling of ferroresonance is as
much an art as a science. As such, it is important if possible
to verify the results by checking the simulations against system measurements. It is highly recommended that anyone active in this area must continually monitor the literature for
improvements in modeling techniques.
6. SUMMARY
This document provides a set of general guidelines for digitalcomputer time-domain simulation of low-frequency (approximately 5 to 1000Hz) transients of electric power systems.

3-23

The report is intended for practicing engineers who are involved in analysis, control and system planning issues related
to electronic power systems. It is assumed that the reader has
(1) a fair understanding of the physical phenomena and (2) an
adequate knowledge of digital simulation techniques. The
guidelines are provided for seven transient torsional torques,
(3) turbine-blade vibrations, (4) fast bus transfer, (5) controller interactions, (6) harmonic interactions and resonance, and
(7) ferroresonance. For those phenomena which have extensively discussed in the literature, i.e. (1) to (4), general guidelines are provided and the reader is frequently referred to the
technical literature for further in-depth modeling and simulation issues. The emphasis of this document is on phenomena
(5), (6) and particularly (7).

7. REFERENCES

[1.1]CIGRE, Guidelines For Representation of Network Elements When Calculating Transients, CIGRE Working
Group 33.02, 1990.
[2.1] IEEE Torsional Issues Working Group, Fourth Supplement To A Bibliography For The Study of Subsynchronous
Resonance Between Rotating Machines and Power Systems,
IEEE Trans., Vol. PWRS-12, No. 3, pp. 1276-1282, August
1997.
[2.2] IEEE SSR Working Group, A Bibliography for the
Study of Subsynchronous Resonance Between Rotating Machines and Power Systems, IEEE Trans., Vol. PAS-95, No.
1, pp. 216-218, Jan. - Feb. 1976.

Transfer Studies, IEEE Trans. on Energy Conversion, Vol.


EC-5, No. 3, pp. 470-476, 1990.
[2.9]P.L. Young, W.L. Snider, T.A. Higgins, H.J. Holley,
Report on Bus Transfer: Part III, Full Scale Testing and
Evaluation, IEEE Trans. on Energy Conversion, Vol. EC-5,
No. 3, pp. 477-484, 1990.
[2.10]P.L. Dandeno, M.R. Iravani, Development and Application of Third-Order Turbogenerator Electrical Stability
Models With Particular Reference To Subsynchronous Resonance Studies, IEEE Trans., Vol. EC-10, No. 1, pp. 78-86,
March 1995.
[2.11]G. Gross, M.C. Hall, Synchronouos Machine and Torsional Dynamic Simulation in Computation of Electromagnetic Transients, IEEE Trans. on Power Apparatus and
Systems, Vol. PAS-97, No. 4, pp. 1074-1086, 1978.
[2.12]IEEE SSR Working Group, First Benchmark Model
for Computer Simulation of Subsynchronous Resonance,
IEEE Trans. on Power Apparatus and Systems, Vol. PAS-96,
No. 5, pp. 1565-1572, 1977.
[2.13]IEEE SSR Working Group, Second Benchmark Model
for Computer Simulation of Subsynchronous Resonance,
IEEE Trans. on Power Apparatus and Systems, Vol. PAS104, No. 5, pp. 1057-1066, 1985.
[2.14]D.A. Woodford, Validation of Digital Simulation of
DC Links, IEEE Trans. on Power Apparatus and Systems,
Vol. PAS-104, No. 9, pp. 2588-2595, September 1985.
[2.15]Manitoba HVDC Research Centre, EMTDC Users
Manual, 1988.

[2.3]IEEE SSR Working Group, First Supplement to a Bibliography for the Study of Subsynchronous Resonance Between Rotating Machines and Power Systems, IEEE Trans.,
Vol. PAS-98, No. 6, pp. 1872-1875, Nov. - Dec. 1979.

[2.16]N. Rostamkolai, et al, Subsynchronous Torsional Interactions with Static VAR Compensators - Influence of
HVDC, IEEE Trans. on Power Systems, Vol. PWRS-6, No.
1, pp. 255-261, Feb. 1991.

[2.4]IEEE SSR Working Group, The Second Supplement to


a Bibliography for the Study of Subsynchronous Resonance
Between Rotating Machines and Power Systems, IEEE
Trans., Vol. PAS-104, No. 2, pp. 321-327, Feb. 1985.

[2.17]L. Gyugyi, Fundamentals of Thyristor-Controlled


Static VAr Compensators in Electric Power System Applications, IEEE Publication 87TH0187-5 PWRS, pp. 8-27, 1987.

[2.5]IEEE SSR Working Group, Third Supplement to a Bibliography for the Study of Subsynchronous Resonance Between Rotating Machines and Power Systems, IEEE Trans.,
Vol. PWRS-6, No. 2, pp. 830-834, May 1991.
[2.6]T.P. Tsao, C. Chgn, Restriction of Turbine Blade Vibrations in Turbogenerators, IEE Proceedings, Vol. 137, Part C,
No. 5, pp. 339-342, 1990.
[2.7]J.D. Gill, Transfer of Motor Loads, Between Out-ofPhase Sources, IEEE Trans. on Industrial Applications, Vol.
IA-15, No. 4, pp. 376-381, 1979.
[2.8]T.A. Higgins, P.L. Young, W.L. Snider, H.J. Holley,
Report on Bus Transfer: Part II, Computer Modelling for

[2.18]A.N. Vasconcelos, et al, Detailed Modeling of an Actual Static VAR Compensator for Electromagnetic Transient
Studies, IEEE Trans. on Power Systems, Vol. PWRS-7, No.
1, pp. 11-19, 1992.
[2.19]H.J. Holley, T.A. Higgins, P.L. Young, W.L. Snider,
A Comparison of Induction Motor Models for Bus Transfer
Studies, IEEE Trans. on Energy Conversion, Vol. EC-5, No.
2, pp. 310-319, 1990.
[2.20]EPRI Report, Bus Transfer Studies for Utility Motors, EERI EL-4286, Vol. 2, Project 1763-2, October 1986.
[2.21]G.J. Rogers, D. Shirmohammadi, Induction Machine
Modelling For Electromagnetic Transients Program, IEEE
Trans., Vol. EC-2, No. 4, pp. 622-628, December 1987.

3-24

[2.22]R.H. Doherty, Analysis of Transient Electrical


Torques and Shaft Torques in Induction Motors as a Result of
Power Supply Disturbances, IEEE Trans., Vol. PAS-101,
No. 8, pp. 2826-2834, August 1982.

No. 1, pp. 335-344, January 1996.

[2.23]Y. Pavlyuk, Application of Static Transfer Switch For


Induction Motor Load Transfer, M.A.Sc. Thesis, University
of Toronto, 1997.

[4.6]D.A. Woodford, Validation of Digital Simulation of DC


Links, IEEE Trans., Vol. PAS-104, No. 9, pp. 2588-2595,
September 1985.

[3.1]A.J.P. Ramos, H. Tyll, Dynamic Performance of a Radial Weak Power System with Multiple Static VAr Compensators, IEEE Trans., Vol. PWRS-4, pp. 1316-1325,
November 1989.

[4.7]M. El-Marsafaway, Use of Frequency Scan Techniques


for Subsynchronous Resonance Analysis of a Practical Series
Compensated AC Network, IEE Proceedings, Vol. 135, No.
1, pp. 28-40, 1983.

[3.2]E.V. Larsen, et al, Basic Aspects of Applying SVCs to


Series Compensated AC Transmission Lines, IEEE Trans.,
Vol. PWRD-5, No. 3, pp. 1466-1473, July 1990.

[4.8]M. Szechtman, T. Wess, C.V. Thio, H. Ring, L. Pilotto,


P. Kuffel, First Benchmark Model for HVDC Control Studies, Electra, No. 135, pp. 54-73, April 1991.

[3.3]IEEE TF Report, Modelling and Analysis Guidelines


For Slow Transients, IEEE Trans., Vol. PWRD-11, No. 3,
pp. 1672-1677, July 1996.

[4.9]CIGRE Working Group 14.02 of Study Committee 14,


The CIGRE HVDC Benchmark Model - A New Proposal
with Revised Parameters:, Electra, No. 157, pp. 60-66, December 1994.

[3.4]M. Parniani, Small-Signal Stability Analysis And Robust Control Design of Static VAR Compensators, Ph.D
Thesis, University of Toronto, 1995.
...[3.5]M. Parniani, M.R. Iravani, Voltage Control Stability
and Dynamic Interaction Phenomena of Static VAR Compensators, IEEE Trans., Vol. PWRS-10, No. 3, pp. 1592-1597,
August 1995.
[3.6]G. Gross, C.F. Imparato, P.M. Look, A Tool for Comprehensive Analysis of Power System Dynamic Stability,
IEEE Trans., Vol. PAS-101, No. 1, pp. 226-234, January
1982.
[3.7]R.M. Hamouda, M.R. Iravani, R. Hackam, Coordinated
Static VAR Compensators and Power System Stabilizers for
Damping Power System Oscillations, IEEE Trans., Vol.
PWRS-2, pp. 1059-1067, 1987.
[3.8]M. Parniani, M.R. Iravani , Computer Analysis of
Small-Signal Stability of Power System Including Network
Dynamics, Proceeding IEE, Gen-Trans-Distrib, Vol. 142,
No. 6, pp. 613-617, November 1995.
[4.1]A.E. Hammad, Analysis of Second Harmonic Instability for the Chateauguay HVDC/SVC Scheme, IEEE Trans.,
Vol. PWRD-7, No. 1, pp. 410-415, January 1992.
[4.2]E.V. Larsen, M. Lublich, S.C. Kapoor, Impact of Stray
Capacitance on HVDC Harmonics, IEEE Trans., Vol.
PWRD-4, No. 1, pp. 637-645, 1989.
[4.3]A.S. Morched, J.H. Ottevangers, L. Marti, Multi-Port
Frequency Dependent Network Equivalents for the EMTP,
IEEE Trans., Vol. PWRD-8, No. 3, pp. 1402-1412, July 1993.
[4.4]A. Sarshar, M.R. Iravani, J. Li, Calculation of Noncharacteristic Harmonics of HVDC Station Using Digital TimeDomain Simulation Method, IEEE Trans., Vol. PWRD-11,

[4.5]Manitoba HVDC Research Center, EMTDC User Manual, pp. 12.44-12.50, 1988.

[4.10]L.A.S. Pilotto, J.E.R. Alves, E.H. Watanabe, A.E. Hammad, A Non-Linear Switching Function Model for Static
VAR Compensators, Proceedings of International Power
Electronics Conference (IPEC), pp. 627-631, Yokohama,
April, 1995.
[4.11]N.L. Shore, G. Andersson, A.P. Canelhas and G. Asplund, A Three-Pulse Model of DC Side Harmonic Flow in
HVDC Systems, IEEE Trans., Vol. PWRD-4, No. 3, pp.
1945-1954, July 1989.
[4.12]D.L. Dickmander, K.J. Peterson, Analysis of DC Harmonics Using the Three-Pulse Model for the Intermountain
Power Project HVDC Transmission, IEEE Trans., Vol.
PWRD-4, No. 2, pp. 1195-1204, April 1989.
[4.13]T.F. Garrity, I.D. Hassan, K.A. Adamson, J.A.
Donahue, Measurement of Harmonic Currents and Evaluation of DC Filter Performance of the New England - Hydro
Quebec Phase I HVDC Project, IEEE Trans., Vol. PWRD-4,
No. 1, pp. 779-786, January 1989.
[4.14]D.L. Dickmander, S.Y. Lee,, G.L. Desilets, M. Granger, AC/DC Harmonic Interaction in the Presence of GIC for
the Hydro Quebec - New England Phase II HVDC Transmission, IEEE Trans., Vol. PWRD-9, No. 1, pp. 68-78, January
1994.
[4.15]H. Stemmler, HVDC Back-to-back Interties on Weak
AC Systems - Second Harmonic Problems Analysis and Solution, CIGRE Symposium, Paper 300-08, Boston, 1987.
[4.16]R.H. Lasseter, L.J. Bohmann, Harmonic Interactions
in Thyristor Controlled Reactor Circuits, IEEE Trans., Vol.
PWRD-4, No. 3, pp. 1919-1925, July 1989.
[4.17]L.J. Bohmann, R.H. Lasseter, Stability and Harmonics
in Thyristor Controlled Reactors, IEEE Trans., Vol. PWRD-

3-25

5, No. 2, pp. 1175-1181, April 1991.

28-30, 1989.

[4.18]S.G. Jalali, R.H. Lasseter, A Study of Nonlinear Harmonic Interaction Between a Single Phase Line-Commutated
Converter and a Power System, IEEE Trans., Vol. PWRD-9,
No. 3, pp. 1616-1624, July 1994.

[5.14]D.G. Fink and H.W. Beatty, Standard Handbook for


Electrical Engineers, 11th Ed., McGraw-Hill Book Company, New York, NY, copyright 1978.

[4.19]X. Jiang, A.M. Gole, A Frequency Scanning Method


for the Identification of Harmonic Instabilities in HVDC Systems, IEEE PES paper 95WM222-0 PWRD.
[5.1]R.G. Andrei and B.R. Halley, "Voltage Transformer Ferroresonance from an Energy Standpoint", IEEE Trans. Power
Delivery, vol. 4, no. 3, pp. 1773-1778, July, 1989.
[5.2]C.M. Arturi, "Transient Simulation and Analysis of a
Five-Limb Generator Step-Up Transformer Following an
Out-of-Phase Synchronization", IEEE Trans. Power Delivery, vol. 6, no. 1, pp. 196-207, January 1991.
[5.3]F.R. Bergseth and S.S. Venkata, Introduction to Electric
Energy Devices, Prentice Hall, Inc., Inglewood Cliffs, NJ,
copyright 1987.
[5.4]J. Bethenod, "Sur le Transformateur et Rsonance",
L'Eclairae Electrique, pp. 289-296, November 30, 1907.
[5.5]J.L. Blackburn, Protective Relaying Principles and Applications, Marcel Dekker, Inc., New York, NY, pp. 231-237,
9th printing, copyright 1987.
[5.6]M.H.J. Bollen, "The Search for a General Transformer
Model", 16th European EMTP Users Group Meeting, paper
89-07, pp. 1-20, May 28- 30, 1989.
[5.7]P. Boucherot, "Existence de Deux Rgimes en Ferro-rsonance", R.G.E., pp. 827-828, December 10, 1920.
[5.8]V. Brenner, "Subharmonic Response of the Ferroresonant Circuit with Coil Hysteresis", AIEE Transactions, vol.
75 I, pp. 450-456, September 1956.
[5.9]J.W. Butler and C. Concordia, "Analysis of Series Capacitor Application Problems", AIEE Trans., vol. 56, pp.
975-988, August, 1937.
[5.10]E.C. Cherry, "The Duality Between Interlinked Electric
and Magnetic Circuits and the Formation of Transformer
Equivalent Circuits," Proceedings of the Physical Society,
Part B, vol. 62, pp. 101-111, 1949.
[5.11]H.W. Dommel with S. Bhattacharya, V. Brandwajn,
H.K. Lauw and L. Marti, EMTP Theory Book, 2nd Ed., Microtran Power System Analysis Corporation, Vancouver, BC,
May 1992.
[5.12]L. Dub and B.A. Mork, "Variable Capacitances and
Inductance in ATP," EMTP News, vol. 5, no. 1, pp. 33-36,
March 1992.
[5.13]G. Empereur, "Miscellaneous - Transformers", 16th
European EMTP Users Group Meeting, Paper 89-09, May

[5.15]J.G. Frame, N. Mohan and T. Liu, "Hysteresis Modeling in an Electro-Magnetic Transients Program", IEEE Trans.
PAS, vol. PAS-101, no. 9, pp. 3403-3411, September, 1982.
[5.16]J. Gleick, Chaos: Making a New Science, Viking, New
York, NY, copyright 1987.
[5.17]C. Hayashi, Nonlinear Oscillations in Physical Systems, McGraw-Hill Book Company, New York, NY, copyright 1964.
[5.18]T. Henriksen and O. Rrvik, "Ferroresonans i 50-kV
Nett til Hafslund", Energiforsyningens Forskningsinstitutt A/
S, Trondheim, Norway, ISBN 82-594-0229-7, EFI TR 3779,
December 19, 1990 (in Norwegian).
[5.19]R.H. Hopkinson, "Ferroresonance During Single-Phase
Switching of 3-Phase Distribution Transformer Banks", IEEE
Trans. PAS, vol. PAS-84, no. 4, pp. 289-293, April 1965.
[5.20]D.C. Jiles and D.L. Atherton, "Theory of Ferromagnetic
Hysteresis," Elsevier Science Publishers B.V., Journal of
Magnetism and Magnetic Materials, vol. 61. pp. 48-60, January 21, 1986.
[5.21]Kieny, "Application of Bifurcation Theory in Studying
and Understanding the Global Behavior of a Resonant Electric Power Circuit," IEEE PES Summer Meeting, SM 265-9
PWRD, July 1990.
[5.22]K.U. Leuven EMTP Center, Alternate Transients Program Rule Book, Leuven EMTP Center, Heverlee, Belgium,
Revised July, 1987, copyright 1987.
[5.23]D.D. Mairs, D.L. Stuehm and B.A. Mork, "Overvoltages on Five-Legged Core Transformers on Rural Electric Systems", IEEE Trans. on Industrial Applications, vol. 25, no. 2,
pp. 366-370, March, 1989.
[5.24]L.W. Match and J.D. Morgan, Electromagnetic and
Electromechanical Machines, Third Edition, Harper & Row
Publishers, Inc., New York, copyright 1986.
[5.25]R.D. Millet, D.D. Mairs and D.L. Stuehm, "The Assessment and Mitigation Study of Ferroresonance on GroundedWye to Grounded-Wye 3-Phase Padmounted Transformers",
Summary Report, NRECA Project 86-7, July, 1987.
[5.26]B.A. Mork, Ferroresonant Modeling Using EMTP, MS
Thesis, North Dakota State University, September 1981.
[5.27]B.A. Mork and D.L. Stuehm, "Application of Nonlinear
Dynamics and Chaos to Ferroresonance in Distribution Systems," IEEE Trans. Power Systems, vol. 9, no. 2, pp. 10091017, April 1994.

3-26

[5.28]B.A. Mork, "Five-Legged Wound-Core Transformer


Model: Derivation, Parameters, Implementation, and Evaluation," Submitted for IEEE PES Winter Meeting, 98WM414,
January 1998.
[5.29]B.A. Mork, Ferroresonance and Chaos - Observation
and Simulation of Ferroresonance in a Five-Legged Core
Distribution Transformer, Ph.D. Dissertation, North Dakota
State University, May 1992. Publication No. 9227588, UMI
Publishing Services, Ann Arbor, MI, 48106, (800) 521-0600.

Delivery, vol. 4, no. 3, pp. 1786-1793, July, 1989.


[5.42]D.L. Stuehm, "Final Report - Three Phase Transformer
Core Modeling," Bonneville Power Administration Award
No. DE-BI79-92BP26700, February 28, 1993. Copies available from BPA, Dept. EOHC, (503) 230-4404.
[5.43]G.W. Swift, "Power Transformer Core Behavior Under
Transient Conditions," IEEE Trans. PAS, vol. PAS-90, no. 5,
pp. 2206-2209, Sep/Oct 1971.

[5.30]A. Narang and R.H. Brierley, "Topology Based Magnetic Model for Steady-State and Transient Studies for Three
Phase Core Type Transformers," IEEE Trans. PWRS, vol. 9,
no. 3, pp. 1337-1349, August 1994.

[5.44]E.J. Tarasiewicz, A.S. Morched, A. Narang, E.P. Dick,


"Frequency Dependent Eddy Current Models for Nonlinear
Iron Cores," IEEE Trans. PWRS, vol. 8, no. 2, pp. 588-597,
May 1993.

[5.31]W.L.A. Neves and H.W. Dommel, "On Modelling Iron


Core Nonlinearities," IEEE Trans. Power Systems, vol. 8, no.
2, pp. 417-425, May 1993.

[5.45]J.M.T. Thompson and H.B. Stewart, Nonlinear Dynamics and Chaos - Geometrical Methods for Engineers and Scientists, John Wiley and Sons, New York, NY, (Reprinted
October 1987), copyright 1986.

[5.32]NRECA, Underground Distribution System Design and


Installation Guide, Section 6, National Rural Electric Cooperatives Association, 1994.
[5.33]O.E. Radulescu, "EMTP Transformer Model with
Type-96 Element - Computation of the Inrush Current",
EMTP Newsletter, vol. 6, no. 2, pp. 25-33, June, 1986.
[5.34]J. Rougin and V. Ranjamina, "Modeling of Magnetic
Circuits with EMTP", EMTP Newsletter, vol. 7, no. 1, pp.
8-28, March 1987.

[5.46]J. Usaola and G. Empereur, "Comparison Between Different Transformer Models in EMTP," EMTP News, vol. 2,
no. 2, pp. 25-34, June 1989.
[5.47]R.A. Walling, K.D. Barker, T.M. Compton, and L.E.
Zimmerman, "Ferroresonant Overvoltages in Grounded WyeWye Padmount Transformers with Low-Loss Silicon-Steel
Cores", IEEE Trans. Power Delivery, vol. 8, no. 3, pp. 16471660, July, 199.

[5.35]R.J. Rush and M.L. Good, "Wyes and Wye Nots of


Three-Phase Distribution Transformer Connections", IEEE
Conference Paper 89CH2709-4-C2, 1989.

APPENDIX A

[5.36]R. Rudenberg, Transient Performance of Electric Power Systems, McGraw-Hill Book Company, New York, NY,
chapter 48, copyright 1950.

FAST BUS TRANSFER TRANSIENTS


Introduction

[5.37]G.R. Slemon, "Equivalent Circuits for Transformers


and Machines Including Non-Linear Effects," Proceedings
IEE, vol. 100, part IV, pp. 129-143, 1953.
[5.38]D.R. Smith, S.R. Swanson and J.D. Borst, "Overvoltages with Remotely-Switched Cable-Fed Grounded Wye-Wye
Transformers", IEEE Trans. PAS, vol. PAS-94, no. 5, pp.
1843-1853, Sep/Oct 1975.
[5.39]P.L. Sorenson, "Simulation of Faults and Switchings in
Electrical Distribution Networks", ATV - NESA - Electrical
Engineering Department, DTH - DEFU, Industrial Research
Project EF186, pp. 1- 120, April, 1988.
[5.40]D.L. Stuehm, B.A. Mork and D.D. Mairs, "Ferroresonance with Three Phase Five-Legged Core Transformers",
Minnesota Power Systems Conference, Minneapolis, MN,
October 3, 1988.
[5.41]D.L. Stuehm, B.A. Mork and D.D. Mairs, "Five-Legged
Core Transformer Equivalent Circuit", IEEE Trans. Power

Motors and other loads in utility and heavy industrial applications are supplied during normal operation from a
preferred power source. An alternate power source is normally provided to supply such motors and other loads during
planned shutdowns and upon loss of normal power from the
preferred power source. The process of disconnecting the
motors and other loads from one source and reconnecting to
an alternate source is commonly defined as bus transfer.
Manual transfer means are normally provided to
allow transferring the motors and other loads from one power
source to the other. However, upon loss of the preferred
power source, the motors and other loads are automatically
transferred to the alternate power source. This automatic
transfer is necessary to allow uninterrupted operation of the
motors and other loads important to personnel safety and process operation.

3-27

The normal and alternate power source connections

are always selected such that they are in phase. Therefore,


manual transfers can be accomplished in a make-beforebreak, i.e., the motors and loads are connected to the second
power source before the first power source is disconnected.
In this overlapping transfer, the power supply is not interrupted and the motors are not subjected to transients. However, during automatic transfers, the motors may be
disconnected from both power sources for a short duration
depending on the type of transfer and the associated circuit
breakers operating times. The time during which the motors
are disconnected from both power sources is termed the
dead time. It is commonly longer than two cycles and can
be as long as 12 cycles. While motors are disconnected from
both power sources they decelerate. The deceleration rate
depends on the motor-load inertia and the synchronizing
power flowing between motors due to their differing characteristics. As the motor decelerates, the relative angle
between the motor internal voltage and the power source
voltage changes. Also, the motor residual voltage decays at a
rate which depends on the motor magnetic characteristics,
speed and initial loading. If the relative angle between the
motor residual voltage and the power source voltage
becomes large enough at the time of reconnection with significant residual voltage remaining, the resultant voltage
between the power source and the motor will produce an
inrush current. The inrush current may be significantly
larger than the normal full voltage starting current. Such
high inrush currents cause high winding stresses and transient shaft torques which can damage the motor and/or the
driven equipment.
The most common bus transfer scheme is the fast
bus transfer scheme. In this scheme, opening of the normal
power source breaker initiates closing of the alternate power
source breaker without intentional time delay. The fast bus
transfer operations result in the motors being disconnected
from both power sources for a duration of as short as 2 cycles
to as long as 12 or more cycles.
Presently, there are no generic criteria to ensure
acceptable fast bus transfer operations. Therefore, it is necessary to analyze the transient behavior of motors during fast
bus transfer operations. The analysis should be on a case by
case basis to ensure that the motors will not be subjected to
excessive inrush currents and/or shaft transient torques.
A.2Modeling and Analysis
A three-phase model of the motors and the power
distribution system is required. This is to permit simulating
the breaker individual pole interruption at separate current
zeros and analyzing the effect of unbalanced faults on the
motor behavior. The model must simulate the motor stator
and rotor dynamics, the load dynamics, and the power source
dynamics when available. The larger motors should be individually modeled; smaller motors unless for the motor being
studied, if any, may be lumped together and modeled by one

equivalent motor with typical characteristics. The distinction


between large and small motors should be made on a case by
case basis.
A.2.1 Motor Electrical System
The motor electrical system may be modeled by the
differential equations describing the stator and rotor quantities and flux linkages [2.20] or by the two-axis model [2.21].
A single rotor motor model may be adequate since the motor
speed usually does not drop significantly during the time a
fast transfer is accomplished. The model should account for
saturation in the magnetizing, stator and rotor leakage reactances.
A.2.2Loads
The mechanical load should be modeled by its
torque-speed characteristics and moment of inertia. Common centrifugal and axial pumps and fans may be modeled
by a quadratic torque-speed characteristic.
Non-motor loads may be lumped together and modeled by an equivalent resistance-inductance circuit. Nonmotor loads would be included in the model to account for
their damping effects on the motors during the dead time.
A.2.3Motor-Load Shaft Torsional Model
The shaft system should be modeled by the motor
rotor mass connected to the load rotor mass by a flexible
spring representing the shaft [2.22]. The motor air gap
torque excites the mass representing the motor rotor while
the load torque excites the mass representing the load rotor.
The shaft torsional model should include the effect of damping and shaft flexibility. The effect of shaft flexibility is particularly important in applications where loads have a large
inertia relative to that of the motor. An example of this application is torsional study of a large boiler fan. Under such
conditions, the shaft flexibility may cause the shaft torque to
be higher than the motor air gap torque [2.22].
A.2.4Circuit Breakers
The circuit breakers should be modeled as a three
pole switch which can be opened or closed at a preset time.
The three poles of the circuit breaker connecting the alternate
source must be modeled to close simultaneously. the individual poles of the circuit breaker disconnecting the normal
source must be modeled to open only at the respective current zero. In analysis involving transfers caused by high
level electrical faults, the individual poles may be modeled to
open at the respective first current zero following the end of
the breaker arcing time. This is a conservative approach
which, in effect, models a zero resistance arc.
A.2.5Power Sources

3-28

Generally, events which initiate bus transfers such


as the loss of the generator in a generating station also initiate
disturbances to the connected power system. The dynamic
variations of the power system voltage magnitude and phase
angle are normally determined as part of system stability
studies.
Ideally, the normal and alternate power source models should reflect the dynamic variations in the voltage magnitude and phase angle following the initiating event. This
can be accomplished by a point to point representation of the
system voltage magnitude and phase angle profiles. Alternately, the system voltage magnitude and phase angle may be
modeled by polynomials fitting their profiles for the short
duration of interest. However, in the event that data on the
dynamic behavior of the power system is not available, the
normal and alternate power sources may be modeled as ideal
sources in phase.
A.2.6Transformers
Transformer dynamics have a little or no effect on
bus transfer operations. Therefore, a transformer may be
modeled as an ideal transformer in series with a lumped
resistance in series with a lumped inductance representing
the transformer equivalent impedance.

torques should be investigated.


The motor air gap torque at the instant of closing the
alternate source breaker is determined by the motor residual
voltage magnitude and phase angle. The magnitude of the
motor residual voltage decreases with time while the phase
angle increased with time. This causes the magnitude of the
air gap torque to be cyclic. It has a minimum value at some
short bus dead time, peaks as the dead time increases and
then decreases as the dead time increases further. The duration of the dead time at which the air gap torque attains minimum and maximum values are system specific and depends
on the connected motors characteristics and load levels.
The account for the above considerations, the following fast bus transfer operations should be simulated and
analyzed.
Transfers caused without high level electrical faults
with motors operating at their highest loadings when the
alternate source voltage is at its maximum level.
Transfers caused without high level electrical faults
as in 1) above except without the largest high inertia motor
running.
Transfers caused by high level electrical faults (line
to ground and three line to ground faults).
The following parameters should be monitored during the simulations:

A.2.7 Cables/Lines
Cable and Lines may be modelled by their -equivalents.
A.2.8Simulation and Analysis
In selecting conditions to be analyzed, the following
should be taken into consideration:

bus instantaneous voltages


bus voltage phase angle
individual motors instantaneous currents
individual motor air gap torques
individual motors shaft torques (when modeled)
individual motors speed

A.3 Model Validation


The motor initial loading (prior to the transfer) has a
significant effect on the rate of change of the motor internal
voltage phase angle. Higher loads cause faster drop in rotor
speed and faster rates of change in the phase angle. Therefore, the worst case transfer results when the motors are operating at their highest loading.
The motor-load inertia also has a significant effect
on the internal voltage phase angle rate of change. Motors
with a high inertia have a slower rate of change than motors
with low inertia. Therefore, the effect of fast bus transfer
operation with and without such high inertia motors should
be evaluated.
Bus transfers are initiated by low and high level
electrical faults. The motor residual voltage decays at a high
rate until the fault is cleared by opening the source breaker.
This causes the transient shaft torque produced upon closing
the alternate source breaker to be relatively low. However,
faults such as line to ground faults cause the motor to experience a high oscillatory torque before the fault is cleared. The
effect of the torsional stress caused by such high oscillatory

Ideally, validating a model of a fast bus transfer


operation should include validating the individual motor
models and the circuit breakers operating times. Individual
motor models can be validated by simulating motor starting
and running conditions and comparison of the simulation
results to data recorded during an actual motor starting test.
Parameters be compared include motor instantaneous current, power, apparent power (VA), and speed. However,
since a typical bus transfer model may include 15 or more
motors, it may not be practical to validate individual motor
models. As an alternative, the bus transfer model can be
based on modeling motors using unadjusted manufacturers
supplied data and establishing a range of the expected accuracy.

3-29

You might also like