Sei sulla pagina 1di 144

Abstract

A voltage dip is a reduction in the voltage magnitude with a duration


between a few cycles and several seconds. Voltage dips are considered
one of the most serious power quality problems. They lead to mal-
operation or tripping of several types of end-user equipment, e.g.
adjustable speed drives, computers, etc. A dip is often characterized by
one magnitude and one duration. This is a reasonable approximation as
long as single-phase equipment (e.g. computers) are concerned.
However, three-phase equipment (e.g. adjustable-speed drives) will
typically experience three different voltage magnitudes, as the majority
of dips are due to single-phase or phase-to-phase faults. The three-
phase voltage relation of the power supply during a dip needs to be
analyzed to assess the inuence of voltage dips on three-phase
equipment.
The three-phase unbalance of voltage dips in their characterization and
propagation is dealt with in this dissertation. A dip classication
method is proposed. The classication method is applied to analyse
voltage dip measurement from power quality survey, and to test three-
phase equipment immunity against voltage dips.
The dip classication is based on the well-proven theory of
symmetrical components. Mathematical models are developed for both
balanced and unbalanced dips, taking into account the fault types,
transformer types and load connections. The characterization results in
a so-called characteristic voltage (a generalized magnitude and phase-
angle shift) for balanced and unbalanced dips as well as a so-called
PN-factor, relating to the dynamic loads contribution to the source
impedance for unbalanced dips. The PN-factor is equal to unity if the
positive- and negative-sequence source impedance of the system are
equal and time-independent. With several acceptable assumptions, the
classication method is able to quantify all three-phase dips by one
single complex number, namely the characteristic voltage. This
signicantly simplies the study of unbalanced dip propagation in
power systems. Field measurements show that the proposed
classication method holds for both transmission systems and
distribution systems. Theoretically, large dynamic loads connected to
the system could affect the correctness of the proposed method.
However, the eld measurements show that the error introduced by
dynamic loads is negligible.

The proposed classication method helps understanding the phase
relationships of unbalanced voltage dips. The phase-angle shift
phenomenon associated with voltage dips is well explained by the
mathematical models introduced. In presenting voltage dip
measurement from power quality survey and performing three-phase
equipment immunity test, the classication method offers a platform to
exchange information between utilities, customers, and equipment
manufacturers.

Keywords:

power systems, power quality, voltage dips(sags),
symmetrical components, characterization, eld measurement,
equipment immunity test.

Preface

The work presented in this thesis has been carried out at the
Department of Electric Power Engineering at Chalmers University of
Technology. The research has been funded through the Elforsk Elektra
program which is jointly nanced by the Swedish National Board for
Industrial and Technical Development (NUTEK) and ABB Corporate
Research. The nancial support is gratefully acknowledged.
I wish to express my deepest gratitude to my supervisor, Dr. Math
Bollen, for supervising this work, for valuable comments, fruitful
discussions and for persistently revising the manuscript. I also would
like to thank my examiner, professor Jaap Daalder, for his assistance
throughout this project and many helpful comments. Many thanks to
all the colleagues at the department of Electric Power Engineering,
especially to the power system group for general assistance in different
ways.
Members of the steering group have been Ulf Grape, Mats Hger, and
Gunnar Ridell. Thank you all for fruitful discussions and valuable
comments.
Special thanks should also be given to Mats Hger of STRI, Alastair
Ferguson of Scottish Power, and Helge Seljeseth of SINTEF Energy
Research for kindly offer of the eld measurement data.
Last but not least, I would like to thank my wife Yibin for love and
support, and for patiently waiting for a husband who often come back
late from work.

LIST OF PUBLICATIONS

This thesis is based on work reported in the following papers, referred
to by Roman numerals in the text:

I L.D.Zhang, M.H.J. Bollen, A method for characterizing unbalanced
voltage dips (sags) with symmetrical components, IEEE Power
Engineering Letters, pp. 50-52, July 1998.
II L.D. Zhang, M.H.J. Bollen, Characteristics of voltage dips(sags) in power
systems, accepted by IEEE PES Transactions.
III L.D. Zhang, M.H.J. Bollen, A method for characterization of three-phase
unbalanced dips from recorded voltage waveshapes, International
Telecommunication Energy Conference(INTELEC), Copenhagen,
Denmark, June 1999.
IV M.H.J. Bollen, L.D. Zhang, Analysis of voltage tolerance of ac adjustable-
speed drives for three-phase balanced and unbalanced sags, accepted by
IEEE Transactions on Industry Applications.
V M.H.J. Bollen, J. Svensson, L.D. Zhang, Testing of grid-connected power
converters for the effects of short circuits in the grid, European Power
Electronics Conference, Lausanne, Switzerland, September 1999.

Contents

Abstract
Preface
Contents
Chapter 1 Introduction 1

1.1 Voltage dips and related studies ...................................................... 1
1.2 Problem of three-phase unbalance ................................................... 3
1.3 Aim and layout of the thesis ............................................................ 4

Chapter 2 Terminology 7

2.1 Voltage dips and other voltage variations ........................................ 7
2.2 Voltage dips in one phase ................................................................ 9
2.3 Example of single-phase dip characterization ................................. 11

Chapter 3 Classification of Three-phase Voltage Dips 15

3.1 Balanced faults ................................................................................. 15
3.2 Unbalanced faults ............................................................................ 17
3.2.1 Two-component symmetrical components .......................... 17
3.2.2 Unbalanced faults analysis by sequence networks .............. 20
3.3 Definition of dip types ..................................................................... 22
3.3.1 The single-line-to-ground fault (SLGF) .............................. 22
3.3.2 The line-to-line fault (LLF) ................................................. 24
3.3.3 The double-line-to-ground fault (2LGF) ............................. 25
3.3.4 The three-phase fault (3F) ................................................. 25
3.3.5 Overview of the classification ............................................. 27
3.3.6 Phase-angle shift in unbalanced dips ................................... 28
3.3.7 Symmetrical phase for unbalanced dips .............................. 29
3.4 Dip transformation through transformers ........................................ 33
3.4.1 Basic transformer models .................................................... 34
3.4.2 Effect of the basic transformer models on the basic dip types 36
3.4.3 Change of the symmetrical phase ........................................ 38
3.4.4 Physical transformers to mathematic models ...................... 40
3.5 Terminology: three-phase voltage dips ............................................ 46

Chapter 4 Voltage Dip Propagation in Power Systems 47

4.1 Voltage dip propagation in distribution systems ............................. 48
4.1.1 Voltage dip propagates upwards and downwards ................ 50
4.1.2 SLGF at medium voltage level ............................................ 54
4.1.3 Local generation .................................................................. 55
4.2 Voltage dip propagation in transmission systems ........................... 56
4.3 A single-phase scheme to study voltage dip propagation ................ 60
4.3.1 Characteristic voltage .......................................................... 60
4.3.2 PN-factor .............................................................................. 62
4.3.3 Dip type ............................................................................... 64
4.4 Loads influence .............................................................................. 64
4.4.1 Motor re-acceleration ........................................................... 65
4.4.2 PN-factor .............................................................................. 66
4.4.3 Limitations of the classification method .............................. 73

Chapter 5 Field Measurement Analysis 75

5.1 Obtaining dip characteristics ........................................................... 75
5.1.1 Principle ............................................................................... 75

Contents

5.1.2 Algorithms for dip characterization .....................................78
5.1.3 Examples ..............................................................................79
5.2 Characteristics obtained from measurements ...................................84
5.2.1 Transmission system: Sweden .............................................84
5.2.2 Distribution system: Scotland ..............................................86
5.2.3 Distribution system: Norway ...............................................87
5.3 Further application examples ...........................................................90
5.3.1 A propagating dip .................................................................90
5.3.2 Statistics from a power quality survey .................................94

Chapter 6 Equipment Immunity Tests 99

6.1 Single-phase equipment test .............................................................100
6.1.1 Test items .............................................................................100
6.1.2 Test setup .............................................................................105
6.1.3 Test example ........................................................................109
6.2 Three-phase equipment test ..............................................................111
6.2.1 Test items .............................................................................112
6.2.2 Test setup .............................................................................117

Chapter 7 Conclusions and Future Research 119

7.1 Conclusions ......................................................................................119
7.2 Future Research ................................................................................122

References 125
App.A Determination of Zero-sequence Source Impedance 131
App.B PN-factor and Characteristic Voltage 133

B.1 Single-line-to-ground Fault (SLGF) ...............................................133
B.2 Line-to-line fault (LLF) ..................................................................135

Chapter 1: Introduction

1

Chapter 1 Introduction

1.1 Voltage dips and related studies

Voltage dips are short duration reductions in rms voltage, mainly
caused by short circuits and starting of large motors. Disruptive voltage
dips are mainly caused by short-circuit faults. Fault conditions on
power systems are the result of a variety of conditions, such as
lightning, wind, equipment failure, accidents, etc. The large interest in
voltage dips is due to the problems they cause on several types of
equipment [1]. Specially computers, adjustable-speed drives and
process-control equipment are notorious for their sensitivity.
Equipment used in modern industrial plants (process controllers,
programmable logic controllers, adjustable speed drives) is actually
becoming more sensitive to voltage dips as the complexity of the
equipment increases and the equipment is interconnected in
sophisticated processes. Voltage dips and short interruptions are the
most troublesome and costly type of power quality problem for most
customers. Interruptions occur when a protective device actually
interrupts the circuit serving a particular customer. This will normally
only occur when there is a fault on that circuit. Voltage dips occur
during faults in a wide part of the power system. Compared to
interruptions, voltage dips occur much more frequent. If equipment is
sensitive to these dips, the frequency of problems will be much higher
than if the equipment would be only sensitive to interruptions [2].
Over the last ten years, voltage dips have become one of the main
topics concerning power quality among utilities, customers and
equipment manufacturers. Several international standards and working
group documents have been produced to improve the understanding of
voltage dip problems[9][17][22][23].
Voltage dip related studies can be divided into the following categories:

1. Characterization of voltage dips

. These studies aim at acquiring a
knowledge of the voltage dip characteristics[20][49][51]. Both fault
propagation studies and measured data are being used for this. The
reduction in rms voltage and the duration of event are the main
characteristics. A voltage dip is normally characterized by one
magnitude and one duration [9][40][48]. However, several studies have
shown that some other characteristics associated with dips, such as

Chapter 1: Introduction

2
phase-angle shift, point-on-wave of initiation and recovery, waveform
distortion, and phase unbalance, may also cause problems for sensitive
equipment. These non-energy related characteristics received
considerable attention in recent years [7][14][15][54].

2. Equipment immunity

. The equipments sensitivity is a major part
of many voltage dip studies. Different types of electrical equipment
have different voltage tolerance. The CBEMA curve (Computer
Business Equipment Manufacturers Association) is widely quoted as a
reference, even though it only refers to mainframe computers ride-
through ability. Recently a revised CBEMA curve has been adopted
by the Information Technology Industry Council (ITIC) which is the
successor of CBEMA. The new curve is also referred to as the ITIC
curve. International standards give instructions on test set-up and
procedure to perform a dip immunity test [10][22][34]. So far, these
only concern magnitude and duration of the voltage dip [42][56]. The
effect of non-energy related characteristics has been studied for several
pieces of equipment [16][26][27], but no international standard has
considered them yet.

3. Stochastic assessment

. To be able to nd out whether a piece of
equipment is compatible with the supply, information must be obtained
on the expected number of voltage dips at the supply point. Two
methods are in use to obtain this information: monitoring the supply
for a certain period and doing a stochastic prediction[3][5]. Large scale
power quality surveys have been performed in several North American
and European countries [29][30][31][32][33]. While these surveys give
a general impression about the dip frequency, it is hard to get accurate
results in a short time [43]. The system structure changes with time,
which also affects the accuracy. Stochastic prediction is based on
historical fault frequency data and the knowledge of voltage dip
propagation in power systems. Stochastic prediction gives a quick
result and can be easily modied when the system structure is changed.
It also allows the evaluation of systems and conditions which not yet
actually exist, e.g. future expansion plans. But various approximations
are often made in these studies, which will nally affect the accuracy
[38][44].

4. Mitigation

. To mitigate the voltage dip problem, i.e. to reduce the
number of equipment problems, several methods have been proposed
and are being in use [46]: 1) Improve the power supply quality to
reduce the number of dips [1][50]; 2) Installation of compensating

Chapter 1: Introduction

3
equipment between the power system and the sensitive
equipment[45][52][53]; 3) Equipment topology modications to
reduce its susceptibility to voltage dips [19][41][57].
These four different categories of studies are quite dependent on each
other. Characterization of dips is the basic platform for other studies.
Before making a decision about which mitigation method to choose,
information is needed about the actual number of disturbing dips and
about the effectiveness of the various mitigation method. Equipment
immunity test is a way to testify the equipment susceptibility in a
certain electromagnetic environment. Most of the time, all these
studies are needed to solve a specic voltage dip problem.

1.2 Problem of three-phase unbalance

Because unbalanced faults (single-phase, phase-to-phase) constitute
the majority of power system faults, unbalanced dips, with different
voltages in the three phases, occur much more frequently than
balanced dips [12]. From the loss of energy viewpoint, an unbalanced
dip is generally considered less severe than a balanced dip if their
lowest phase voltage is the same. But unbalanced dips show larger
values for the non-energy related characteristics, such as phase
unbalance and larger phase-angle shift[7]. This could also cause
problems to specic equipment.
As mentioned before, a voltage dip is often characterized by one
magnitude and one duration. Dip data is often presented with
susceptibility curves overlaid in magnitude-duration plots. While such
a plot could make sense to single-phase equipment, it will almost
certainly give misleading result for three-phase equipment.
One common way is to present only the lowest of the three phase
voltages for each event. This implies a three-phase load that is sensitive
to the lowest of the three phase voltages. This is unlikely to be
generally true. Simulations have shown this to be incorrect for three-
phase adjustable-speed drives [28]. No study is known to the author
suggesting such a sensitivity to the lowest voltage. A single-phase
voltage dip down to a certain level may not affect any equipment, yet a
three-phase or two-phase dip of the same magnitude may cause all end-
use equipment to malfunction. Several other characterization methods
have been proposed, including reporting the average voltage of the

Chapter 1: Introduction

4
three-phases, which assumes three-phase loads sensitive to the average
voltage. This magnitude does not match any of the three individual
phase dip voltages. This method has the same problem as the rst
approach. For a certain piece of equipment, it is not that
straightforward to choose a single value to quantify a dip. Reporting
each of the three-phases separately will obviously give a complete
picture, but will make it hard to use a dip coordination chart.
For a stochastic prediction study, it is more complicated to do a
prediction for unbalanced dips than for balanced dips. In the latter case
a single-phase scheme can be used. When a certain unbalanced dip
propagates in the power network, it behaves differently than a balanced
dip, e.g. a voltage drop in one phase becomes a drop in two phases
through a delta-wye connected transformer, and vice versa[2][6]. A
load connected in delta experiences different voltages than a star-
connected load due to the same reason.
To perform a three-phase equipment immunity test, a certain amount of
knowledge of the phase relationships of the three phases is needed.
Besides the different magnitudes, an unbalanced dip typically
experiences different phase-angle shift in three phases[14][15]. This
requires an understanding of different types of unbalanced fault and
their propagation in power systems, as well as statistics about their
occurence. As mentioned above, neither existing methods for
stochastic prediction nor power quality survey treat unbalanced dips in
a satisfactory way. A practical three phase equipment immunity test
can not be properly performed without a correct treatment of three-
phase unbalance.

1.3 Aim and layout of the thesis

This thesis intends to introduce a new concept in dealing with the
three-phase unbalance problem in voltage dip studies.
A classication of voltage dips as experienced by three-phase
equipment is proposed based on the well-known symmetrical
components. The classication is valid for all balanced and unbalanced
dips, taking into account the different fault types, transformer types
and load connections. The classication results in three dip types, one
for balanced dips and other two for unbalanced dips. Those two types
of unbalanced dips can be further classied as six types if their phase

Chapter 1: Introduction

5
symmetry is considered. The characterization of three-phase dips
results in a so-called characteristic voltage (leading to a generalized
magnitude and phase-angle shift) for both balanced and unbalanced
dips as well as a factor, related to the rotating machine contribution to
the source impedance for unbalanced dips. It is shown that this factor is
close to unity, so that both balanced and unbalanced dips can be
characterized in many cases with one phasor.
The proposed method of classication solves the three-phase
unbalance problem mentioned in Section 1.2. Since both balanced dips
and unbalanced dips can be quantied by one voltage, the generally-
used susceptibility curves can still be used, the only difference being
separation of the three dip types. An equipment immunity curve can be
also obtained for a piece of equipment for each type of dip.
In Chapter 2 the characteristics of voltage dips and other voltage
disturbances on a single phase are dened for single-phase equipment,
which provides a consistent terminology for the other chapters. A
method of classication of three-phase dips is proposed in Chapter 3
The propagation of three-phase dips in power systems is studied in
Chapter 4 using the classication introduced in Chapter 3. In Chapter
5, the concept of dip classication is applied to data obtained from a
power quality survey. Field measurement data is used to verify the
theory. In Chapter 6, equipment immunity tests against voltage dips are
described. The test for three-phase equipment is based on the proposed
dip classication theory. The laboratory setup needed for performing
the test is also discussed.

Chapter 1: Introduction

6

Chapter 2: Terminology

7

Chapter 2 Terminology

The terminology used in describing voltage dips and other voltage
disturbances is developed in several international standard documents
[8][9]. Unfortunately these documents are not always consistent and
some phenomena are not dened in these standards. Many terms have
been used in the power quality literature have multiple or unclear
meaning. To avoid confusion, a list of terminology is included in this
chapter. The terms are put into three categories: 1) voltage dips and
other voltage variations; 2) voltage dip in one phase; 3) three-phase
voltage dips. Denitions related to three-phase voltage dips are based
on the theory in Chapter 3, these denitions are therefore put in Section
3.5 of Chapter 3.

2.1 Voltage dips and other voltage variations

RMS

: Root-mean-square value of voltage or current over one cycle or
one half-cycle of the fundamental power frequency (50 Hz or 60 Hz).

Overvoltage

: An increase in the

RMS

voltage to greater than 110% for
a duration longer than 1 min. Also called a long overvoltage to
distinguish it from a

voltage swell

(a short overvoltage). Another value
than 110% can be used if the normal operating voltage limits are
different from 90-110%.

Undervoltage

: A decrease in the

RMS

voltage to less than 90% at the
power frequency for a duration longer than 1 min. Also called a long
undervoltage to distinguish it from a

voltage dip

(a short
undervoltage). Another value than 90% can be used if the normal
operating voltage limits are different from 90-110%.

Short interruption

: An decrease in the

RMS

voltage to less than 1%
for a duration not exceeding 1 min. A cause-based denition would be:
the total loss of supply followed by automatic restoration of the supply.

Long interruption

: A decrease in the RMS voltage to less than 1% for
a duration in excess of 1 min. A cause-based denition would be: the
total loss of supply followed by manual restoration of the supply.

Voltage dip (sag)

: A short-duration (typically less than 1 minute)
reduction in

RMS

voltage due to a short circuit fault, motor starting, or
the switching of a large load. An event with zero voltage is normally

Chapter 2: Terminology

8
not called a voltage dip but an interruption. In case such an event is due
to a short circuit fault we will still refer to it as a voltage dip.

Voltage swell

: A short-duration (typically less than 1 minute) increase
in the

RMS

voltage due to a short circuit fault, a switching action in
the system, or the switching of a large load.

Voltage unbalance

: A condition in which the three-phase voltages
differ in magnitude, are displaced from their normal 120 degree phase
relationship or both.

Magnitude unbalance

: The maximum deviation among the three
phases from the average three-phase voltage divided by the average the
three-phase voltage.

Phase-angle unbalance

: The maximum deviation of the angular
difference between the three phases divided by 120

0

.

Negative-sequence unbalance ratio

: The ratio of the negative-
sequence component to the positive sequence component, usually
expressed as a percentage.

Zero-sequence unbalance ratio

: The ratio of the zero-sequence
component to the positive-sequence component, usually expressed as a
percentage.

Subcycle overvoltage

: A sudden voltage increase within a short
duration (less than half cycle) that is unidirectional in polarity.

Subcycle undervoltage

: A sudden voltage decrease within a short
duration (less than half cycle) that is unidirectional in polarity.

Subcycle oscillatory disturbance:

A sudden increase or decrease in
voltage followed by an oscillation of the voltage. The oscillation
frequency is well above the fundamental power system frequency.

Voltage notch

: A switching (or other) disturbance of the normal power
voltage waveform, lasting less than one half-cycle, which is initially of
opposite polarity than the waveform.

Chapter 2: Terminology

9

2.2 Voltage dips in one phase

Here we assume that only the voltage in one phase is of interest, e.g. to
study the voltage tolerance of a single-phase device.

Dip magnitude

: The remaining RMS voltage in percent or per unit of
pre-fault voltage during fault. In case of a

non-rectangular

dip, the dip
magnitude is a function of time.

Remaining complex voltage

: A complex number which represents the
voltage dip in one phase. Its absolute value is the

dip magnitude

and
its argument is the

phase-angle shift

of the voltage.

Minimum magnitude

: The lowest value of the

dip magnitude

between dip initiation and voltage recovery.

Voltage drop

: The difference between the pre-event

RMS

voltage and
the

RMS

voltage during the event, expressed in volt, pu. or percent.

Maximum voltage drop

: The largest value of the

voltage drop

between dip initiation and voltage recovery.

Missing voltage

: Difference between the actual voltage during the
event and the voltage as it would have been if the event had not taken
place.

Complex missing voltage

: A complex number which represents the
missing voltage of a voltage dip in one phase. It is dened as the
difference in the complex plane between the pre-event voltage and the
voltage during the dip.

Magnitude of the missing voltage

: The

RMS

value of the missing
voltage. In case of a non-rectangular dip, the magnitude of the missing
voltage is a function of time.

Maximum magnitude of the missing voltage

: The maximum
magnitude of the missing voltage between dip initiation and voltage
recovery.

Phase-angle shift

(Phenomenon): A voltage dip caused by a short
circuit in a system not only has a drop in voltage magnitude but also a
shift in the phase angle of the voltage. Two phenomena contribute to
phase-angle shift. A difference in X/R ratio between the source and the
faulted feeder, results in a phase-angle shift at the point of common

Chapter 2: Terminology

10
coupling (PCC) between the fault and the load. Phase unbalance due to
unbalanced faults.

Phase-angle shift

(Quantied): The displacement in time of the
during-event voltage-waveform relative to the pre-event waveform. A
positive phase-angle shift indicates that the phase angle of during-event
voltage leads the pre-event voltage. A negative phase-angle shift
indicates that the phase angle of during-event voltage lags the pre-
event voltage.

Maximum phase-angle shift

: The maximum

phase-angle shift

in
case the phase shift is not constant during the fault.

Point-on-wave of dip initiation:

Phase angle of the voltage at the
moment the voltage waveshape shows a signicant drop compared to
its normal waveshape. The phase angle is measured compared to the
last upward zero-crossing of the voltage. It will not for each dip be
possible to recognize a point-on-wave of dip initiation.

Point-on-wave of dip recovery

: phase angle of the voltage at the
moment the voltage waveshape shows a signicant recovery. It will not
for each dip be possible to recognize a point-on-wave of dip recovery.
Dip duration (1): duration of RMS reduction of a voltage dip. It is
calculated as the persistent time that the phase with lowest magnitude
is lower than 90% of the nominal voltage.
Dip duration (2): The duration of the dip between the point-on-wave
of dip initiation and point-on-wave of dip recovery.
Post-fault dip: The phenomenon with a voltage dip due to a short
circuit fault, that the voltage remains outside the normal operating
range, even after the fault has been cleared. The re-acceleration of
motor may cause an extended dip if the motor load is large with respect
to the system impedance after the fault is cleared. The post-fault dip
can last up to several seconds and the voltage will be between 60% and
90%. Post-fault dip extends the dip duration and can cause tripping of
equipment which survived the during-event dip.
Non-rectangular dip: A voltage dip where the dip magnitude vs. time
is not consistent. The dynamic loads, e.g. induction motors, are often
the cause of such phenomena.
Chapter 2: Terminology
11
2.3 Example of single-phase dip characterization
Figure 2.1 shows a measured voltage dip in one phase with negative
phase-angle shift. Figure 2.2 and Figure 2.3 plot the dip magnitude and
phase-angle shift according to the denition in Section 2.2. Table 2.1
gives numerical result of the denition in Section 2.3.
Duration (2)
Negative phase-angle shift
Figure 2.1 A measured voltage dip in one phase, obtained from scottish
power.
0 0.05 0.1 0.15 0.2 0.25
-2.0
-1.0
0.0
1.0
2.0
Time [s]
V
o
l
t
a
g
e

[
p
u
.
]
0.0 0.05 0.1 0.15 0.2 0.25
0.2
0.6
0.8
1.0
0.4
0.0
V
o
l
t
a
g
e

(
p
u
.
)
Time (s)
Duration (1)
Figure 2.2 Dip magnitude of Figure 2.1.
Chapter 2: Terminology
12
The dip magnitude in Figure 2.2 is calculated as the RMS voltage over
a window of one cycle, which was 96 samples for the recording used.
Each point in Figure 2.2 is the RMS voltage over the preceding 96
points:
(2.1)
with N = 96 and v
i
the sampled voltage in time domain.
The phase-angle shift in Figure 2.3 is obtained from the phase angle of
the fundamental component of the voltage compared to the pre-fault
voltage. The complex fundamental component was obtained from a
Fast Fourier Transformation (FFT). Let V(t) be the complex
fundamental voltage over the period [t, T] with T one cycle of the
fundamental frequency, and V
0
be the complex voltage at t = 0. The
synchronous voltage has an angle
0
+ t with the angular speed of
the fundamental frequency. The phase-angle shift as plotted in
Figure 2.3 can be calculated from:
(2.2)
The oscillation of the phase angle around dip initiation and voltage
recovery are due to the shift of the window in and out of the dip. It
takes one cycle before the phase-angle shift reaches a reliable value. In
0 0.05 0.1
0.15 0.2 0.25
-30
-20
-10
0
10
Time (s)
P
h
a
s
e
-
a
n
g
l
e

s
h
i
f
t

(
d
e
g
r
e
e
)
Figure 2.3 Phase-angle shift of the dip in figure 2.1
V
rms
k ( )
1
N
---- v
i
2
i k N 1 + =
i k =

V t ( )
V
0
e
jwt
-----------------
( ,
, (
j \
arg =
Chapter 2: Terminology
13
calculating the maximum phase-angle shift listed in Table 2.1, the
oscillation areas are skipped.
Table 2.1: Numerical characterization results of dip in Figure 2.1
Some comments on the denition of characteristics for dips in a single
phase:
1. In IEC standards and UNIPEDE documents, the severity of a dip is
quantied by the voltage drop during the dip [9][17], while in IEEE
standards the magnitude refers to the remaining magnitude during the
dip [8]. In IEC 61000-4-11[10], another term, test level,
corresponding to the IEEE denition magnitude, is used. The
magnitude of a dip can be described as a percentage of nominal voltage
or as a percentage of pre-fault voltage. In analysing voltage dip
measurement and assessing appropriate voltage regulation at a piece of
equipment, the pre-fault voltage reference is more convenient.
However, end-use equipment is rated based on its nameplate rating or
nominal voltage, so that nominal voltage is a good measure. In
reporting voltage dips, it is important to clarify the notation used.
2. The RMS plot of dip magnitude (Figure 2.2) uses a one cycle
moving window for the calculation. Although the actual waveform
drops rather abruptly, the RMS voltage shows a smooth transition to its
during-event value. The same phenomenon occurs upon recovery.
Thus, the duration (1) dened by RMS drop over-estimates the dip
duration by up to one half cycle. The similar phenomenon occurs for
the phase-angle shift, with an oscillation occurs during the transition.
The phase-angle is obtained by FFT with a one-cycle moving window.
Denition Value Denition Value
Minimum dip
magnitude (pu.)
0.3884 Dip duration (2)
(time,second)
0.09
Maximum voltage drop
(pu.)
0.6116 Point-on-wave of
initiation (time, second)
0.079
Maximum missing
voltage (pu.)
0.6235 Point-on-wave of
initiation (degree, angle)
86
0
Maximum phase-angle
shift (degree)
-13
0
Point-on-wave of
recovery (time, second)
0.169
Dip duration (1) 0.097 Point-on-wave of
recovery (degree, angle)
75
0
Chapter 2: Terminology
14
3. Duration (2), dened by the point-on-wave gives a better estimation
of the exact time that the fault lasts. But duration (2) under-estimates
the dip duration in case the event is associated with a deep post-fault
dip.
4. Dip magnitude and voltage drop are two opposite ways to describe
the severity of a dip. Their absolute value addition is equal to the pre-
fault voltage, which is assumed to be 1 pu.
5. Remaining complex voltage and complex missing voltage are two
opposite ways to describe the severity of a dip. Their vector addition is
equal to the pre-fault voltage, which is assumed to be 1 pu.
d
ip
m
a
g
n
itu
d
e
voltage drop
pre-fault voltage
dip magnitude
m
issing voltage
remaining voltage
missing voltage
Figure 2.4 Phasor relationship of complex remaining voltage and
complex missing voltage. Bold --- complex number, Italic ---
scalar number.
Chapter 3: Classication of Three-phase Voltage Dips
15
Chapter 3 Classication of Three-phase Voltage
Dips
Voltage dips are mainly caused by short circuit faults in power systems.
In this chapter, a simple voltage divider model is used to illustrate the
short circuits and voltage dips, and how this leads to a drop in rms
voltage and a phase-angle shift. Dips caused by unbalanced faults are
analyzed by symmetrical components. Based on the fault analysis
using symmetrical components, a classication of three-phase voltage
dips is introduced. Mathematical models for transformers are
developed to study the changes in the dip characteristics when a dip
goes through a transformer. Finally, the terminology related to three-
phase dips is dened as a complement of Chapter 2.
3.1 Balanced faults
To explain the origin of voltage dips and the associated phase-angle
shifts due to a system fault, a voltage divider model is often used. The
voltage divider is shown in Figure 3.1.
Assuming that a three-phase fault(3) occurs at position F, the
(complex) voltage remaining at the point-of-common coupling (PCC)
during the fault is
(3.1)
Figure 3.1 Voltage divider model
PCC
F
Load
Z
f
Z
s
V
dip
Z
f
Z
s
Z
f
+
------------------ = pu.
Chapter 3: Classication of Three-phase Voltage Dips
16
where Z
f
is the feeder impedance, and Z
s
is the source impedance. The
pre-fault voltage is here assumed to be .
Besides the voltage drop (the absolute value of V
dip
), there is often
phase-angle shift present. The origin of the phase-angle shift can be
understood as follows:
Let
(3.2)
and
(3.3)
The argument of the remaining complex voltage is
(3.4)
The argument of expression (3.4) is the difference between the pre-
fault and the during-fault voltage phase angle. Due to the symmetry of
the three-phases during a 3, we can easily introduce a phasor. We will
use the term characteristic voltage, for the remaining complex
voltage, V
dip
in (3.1). Note that the remaining complex voltages are the
same in all three phases during a three-phase fault to quantify a
balanced voltage dip.
The magnitude of a dip depends on the distance between the fault and
the PCC and on the fault level at PCC. A stronger system, where Z
s
is
smaller, results in an increase of the dip magnitude (a less severe
event). The phase-angle shift associated with a dip is determined by the
X/R ratio difference between the source and the feeder. In a
distribution system, the feeder has a smaller X/R ratio compared to the
source. Thus a negative phase-angle shift is often accompanying a dip
due to a fault in distribution systems.
In a mesh-connected network, such as in transmission systems, the
concept of voltage divider model is still useful [3]. However, the
feeders and the source impedance are not easy to identify. This
normally requires a computer program for network fault analysis. In
transmission systems, there is no big difference in the X/R ratio
1 0
0

Z
S
R
S
jX
S
+ =
Z
f
R
f
jX
f
+ =
arc
X
f
R
f
------
( ,
j \
tan arc
X
S
X
f
+
R
S
R
f
+
-------------------
( ,
j \
tan =
Chapter 3: Classication of Three-phase Voltage Dips
17
between the source and the feeders. Thus no signicant phase-angle
shifts occur for balanced dips due to faults on transmission lines.
3.2 Unbalanced faults
The three phase voltages during an unbalanced system fault generally
show different magnitudes and phase-angle shifts. Furthermore, the
three phase quantities show different changes when they move through
various types of transformers and load connections [6]. Thus we can
not directly use any single one of the three phase voltages to quantify
an unbalanced dip.
3.2.1 Two-component symmetrical components
The method of two-component symmetrical components is proposed
by P.M. Andersson in [12] to simplify calculations by reducing the
numbers of components. The traditional symmetrical component
method (also called three-component method) requires three values,
where the two-component method only needs two. The two-component
method is based on the idea that the positive- and negative- source
impedance are equal (Z
s1
= Z
s2
). This assumption holds for any static
circuit, such as transmission lines and transformers. It is not fully
correct for synchronous or induction motors. However, since the
inuence of rotating machines on the source impedance is usually
small, the positive- and negative-sequence source impedance are nearly
equal in reality.
In the theory of symmetrical components, three sequence networks,
positive-, negative-, and zero-sequence networks are dened under the
unbalanced situation, as shown in Figure 3.2.
Chapter 3: Classication of Three-phase Voltage Dips
18
where V
a1
, I
a1
, and Z
s1
represent the positive-sequence voltage,
current, and source impedance; V
a2
, I
a2
, Z
s2
represent the negative-
sequence voltage, current, source impedance; V
a0
, I
a0
, and Z
s0
represent the zero-sequence voltage, current, and source impedance. F
is the fault point and N is the zero-potential bus. V
F
is the pre-fault
voltage of phase a at F.
The relations among them can be written in matrix form
(3.5)
Under the assumption
(3.6)
Equation (3.5) can be re-written, by adding and subtracting the second
and third rows, as
(3.7)
+
-
V
F
N
0
Z
s1
F
1
N
2
ZERO
POSITIVE NEGATIVE
Figure 3.2 Sequence networks with defined sequence quantities.
F
0 F
2
Z
s2
Z
s0
N
1
+
+
-
-
-
+
-
V
a0
V
a1
V
a2
I
a0
I
a1
I
a2
V
a0
V
a1
V
a2
0
V
F
0
Z
s0
0 0
0 Z
s1
0
0 0 Z
s2
I
a0
I
a1
I
a2
=
Z
s1
Z
s2
=
V
a0
V
a1
V
a2
+
V
a1
V
a2

0
V
F
V
F
Z
0
0 0
0 Z
s1
0
0 0 Z
s1
I
a0
I
a1
I
a2
+
I
a1
I
a2

=
Chapter 3: Classication of Three-phase Voltage Dips
19
These Equations are interesting because the last two rows are both
positive-sequence Equations.
For convenience we dene the sum and difference quantities
(3.8)
The analysis equation and synthesis equation for two-component
method can also be derived.
From the analysis equation of the three-component symmetrical
component method,
(3.9)
where
the new analysis equation is derived by adding and subtracting the
second and third rows of Equation (3.9)
(3.10)
V
a0
V
a
V
a
V
a0
V
a1
V
a2
+
V
a1
V
a2

V
a0
V
a1
V
a2
1 1 1
1 a a
2
1 a
2
a
V
a
V
b
V
c
=
a
1
2
--- j
3
2
------- + =
V
a0
V
a
V
a
1
3
---
1 1 1
2 1 1
0 j 3 j 3
V
a
V
b
V
c
=
Chapter 3: Classication of Three-phase Voltage Dips
20
accordingly the synthesis equation
(3.11)
can be also written as
(3.12)
The derivation of Equations for currents follow the same procedure as
voltages.
3.2.2 Unbalanced faults analysis by sequence networks
Under two-component symmetrical components, sequence networks
can be constructed for different kinds of shunt faults in power systems.
Detailed construction procedure can be found in [12].
The point for analysis is chosen at PCC in Figure 3.1, that is to say, we
analyze V
a0
, V
a
, and V
a
at PCC.
a. The single-line-to-ground fault (SLGF)
V
a
V
b
V
c
1 1 1
1 a
2
a
1 a a
2
V
a0
V
a1
V
a2
=
V
a
V
b
V
c
1 1 0
1
1
2
--- j
3
2
-------
1
1
2
--- j
3
2
-------
V
a0
V
a
V
a
=
Chapter 3: Classication of Three-phase Voltage Dips
21
b. The line-to-line fault (LLF)
c. The double-line-to-ground fault (2LGF)
Figure 3.3 Sequence network connections at PCC for a single-line-to-
ground fault on phase a.
+
-
-
+
-
P
1
V
a
V
F
Z
s1
N
1
I
a
0 =
P
1
Z
s1
V
F
I
a
2I
a0
=
V
a
V
a0
Z
s0
2
------------
N
0
N
1
2I
a0
I
a
P
0
+
-
+
+
-
Z
f 1
Z
f 0
2
--------- +
Figure 3.4 Sequence network connections for a line-to-line fault on phase
b and c.
+
-
-
+
P
1
V
a
V
F
Z
s1
N
1
I
a
0 =
+
-
-
+
P
1
V
a
V
F
Z
s1
N
1
I
a
Z
f 1
Chapter 3: Classication of Three-phase Voltage Dips
22
3.3 Definition of dip types
In Section 3.2, Equation (3.12) shows that three-phase unbalanced
voltages can be characterized through three sequence values V
a0
, V
a
,
and V
a
. Under the assumption of (3.6), both V
a
and V
a
are positive-
sequence quantities. That is, both the sigma and delta quantities are
dened as voltages or currents associated with the same positive-
sequence network. The number of networks is reduced from three to
two.
The zero-sequence voltage does not need to be considered in three-
phase voltage dips. Two reasons for this can be given: 1) The zero-
sequence voltage usually equals zero at the equipment terminals, since
it does not pass a delta-star, delta-delta, or ungrounded star-star
connection transformer. 2) Three-phase equipment is normally delta-
connected or ungrounded star-connected, so that the zero-sequence
voltage over the equipment terminals is zero.
3.3.1 The single-line-to-ground fault (SLGF)
The sequence network for the SLGF as shown in Figure 3.3 shows an
interesting result: where V
a
is always equal to the pre-fault voltage V
F
,
independent of the position of the fault. Thus V
a
is the only quantity
Figure 3.5 Sequence network connections for a double-line-to-ground fault
on phase b and c.
+
-
-
+
P
1
V
a
V
F
Z
s1
N
1
I
a
+
-
-
+
P
1
V
a
V
F
Z
s1
N
1
I
a
Z
f 1
+
-
2V
a0
2Z
s0
N
0
I
a0
P
0
Z
f1
+2Z
f0
Chapter 3: Classication of Three-phase Voltage Dips
23
subject to change. In other words, the dip can be characterized fully by
V
a
only. Of course, two-component symmetrical components are
based on the assumption Z
s1
= Z
s2
. These two values are never exactly
equal in reality. We therefore introduce the so-called positive-negative
factor (PN-factor) which is equal to V
a
for a SLGF. The PN-Factor
is equal to the pre-fault voltage if the positive- and negative- sequence
source impedances are the same.
Voltage dips due to single-line-to-ground faults were earlier dened as
type D[6]. The three voltage phasors for a voltage dip of type D with
characteristic voltage V and PN-factor F, are given as follows:
(3.13)
These expressions are obtained by lling in, V
a0
= 0, V
a
= F, V
a
= V,
in (3.12).
For Z
s1
= Z
s2
, the earlier analysis leads to F = 1, after which (3.13)
becomes identical to the expression for a type D dip in [6]. Thus the
earlier classication implicitly assumed equal positive- and negative-
sequence impedances. V is the only variable if PN-factor F equals to 1.
V is called the characteristic voltage of type D.
A single-line-to-ground fault generates a dip of type D at PCC, where
the characteristic voltage has the following expression:
(3.14)
Note that the characteristic voltage for a dip due to a three-phase fault,
would be:
(3.15)
V
a
V =
V
b
1
2
---V
1
2
--- jF 3 =
V
c
1
2
---V
1
2
--- jF 3 + =
V
Z
f 1
Z
s0
Z
f 0
+
2
----------------------- +
Z
s1
Z
f 1
Z
s0
Z
f 0
+
2
----------------------- +
( ,
j \
+
----------------------------------------------------------- V
F
=
V
3
Z
f 1
Z
s1
Z
f 1
+
----------------------- V
F
=
Chapter 3: Classication of Three-phase Voltage Dips
24
A SLGF fault gives a higher value for characteristic magnitude than a
3 fault. The difference can be thought as due to an additional
impedance (Z
s0
+ Z
f0
)/2 between the PCC and the fault.
3.3.2 The line-to-line fault (LLF)
The sequence network for the LLF as shown in Figure 3.4 also shows
that V
a
is equal to the pre-fault voltage. These dips can be fully
characterized by V
a
only. For these dips we dene the characteristic
voltage V as V
a
and the PN-factor F as V
a.
Voltage dips due to LLF
faults were earlier dened as type C [6]. The three voltage phasors for a
voltage dip of type C with characteristic magnitude V and PN-factor F,
are given as follows:
(3.16)
These expressions are obtained by lling in, V
ao
= 0, V
a
= V, V
a
= F,
in (3.12).
For Z
s1
= Z
s2
, the earlier analysis leads to F = 1, after which (3.16)
becomes identical to the expression for a type C dip in [6]. Thus the
earlier classication implicitly assumed equal positive- and negative-
sequence impedances. V is the only variable if PN-factor F equals to 1.
V is called the characteristic voltage of type C.
A line-to-line fault generate a dip of type C at the PCC, where the
characteristic voltage V is given by the following expression:
(3.17)
Note that the characteristic voltage for a dip due to a LLF fault is equal
to the characteristic voltage for a dip due to a three-phase fault. This
doesnt imply that the severitie of the events are the same: different
fault types lead to different dip types.
V
a
F =
V
b
1
2
---F
1
2
--- jV 3 =
V
c
1
2
---F
1
2
--- jV 3 + =
V
Z
f 1
Z
s1
Z
f 1
+
----------------------- V
F
=
Chapter 3: Classication of Three-phase Voltage Dips
25
3.3.3 The double-line-to-ground fault (2LGF)
The sequence network of a 2LGF, as shown in Figure 3.5, shows that
both V
a
and V
a
are variables and subject to change when the dip
propagates through the network. However, V
a
is much bigger than
V
a
, and V
a
is close to one when the systems zero sequence is large.
The resulting voltage dip is therefore considered as a type C dip,
according to the following expressions:
(3.18)
These expressions are obtained by lling in, V
ao
= 0, V
a
= V, V
a
= F,
in (3.12).
A 2LGF fault generates a dip of type C at the PCC, where the
characteristic voltage V and the PN-factor F are found from the
following expressions
(3.19)
and
(3.20)
The characteristic voltage is again the same as for a three-phase fault.
But contrary to dips due to LLF faults, the PN-factor is less than unity.
3.3.4 The three-phase fault (3F)
Dips due to balanced faults are also dened in a similar way. Since the
negative-sequence voltage V
2
is equal to zero for balanced dips, V
a
=
V
a
F =
V
b
1
2
---F
1
2
--- jV 3 =
V
c
1
2
---F
1
2
--- jV 3 + =
V
Z
f 1
Z
s1
Z
f 1
+
----------------------- V
F
=
F
Z
f 1
2 Z
s0
Z
f 0
+ ( ) +
Z
s1
Z
f 1
2Z
s0
2Z
f 0
+ + ( ) +
----------------------------------------------------------------- V
F
=
Chapter 3: Classication of Three-phase Voltage Dips
26
V
a
= V
1
. These kinds of dips are dened as type A. The phasor
voltages for dips of type A are given by the following expressions:
(3.21)
Equation (3.21) is also obtained from equation (3.12) by substituting
V
ao
= 0, V
a
=V
a
= V. V is the only variable in equation (3.21). V is
called the characteristic voltage of type A.
A 3F fault generates a dip of type A at the PCC, where the
characteristic voltage V is found from the following expressions
(3.22)
Figure 3.6 shows the phasor diagram of three different types of dips,
given a PN-factor equals to 1 for unbalanced dips, and a characteristic
voltage equal to 0.50
0
for each type.
The denitions of dip types were rst introduced by [6], where four
types of dips, type A, B, C, D were dened. The dip types as dened in
this section are a generalisation based on symmetrical components. An
additional PN-factor F is introduced in type C and type D to include
V
a
V =
V
b
1
2
---V
1
2
--- j 3V =
V
c
1
2
---V
1
2
--- j 3V + =
V
Z
f 1
Z
s1
Z
f 1
+
----------------------- V
F
=
Figure 3.6 Phasor diagram of the three types of voltage dips. PN-factor F
= 1.0, zero-sequence voltage V
a0
= 0.0, characteristic voltage V
= 0.50
0
. Dashed line: the pre-fault phase voltages; solid line:
phase voltages during the dip.
Type A Type C Type D
Chapter 3: Classication of Three-phase Voltage Dips
27
the situation where positive- and negative-sequence source impedances
are not equal. Type B is a special case of type D which holds for zero-
sequence impedance equal to positive-sequence impedance. This
assumption generally doesnt hold in power systems. In the proposed
classication, this type is considered as type D without special
consideration.
3.3.5 Overview of the classification
In the previous sections, a classication of three-phase unbalanced dips
into three types, is introduced. The classication is based on the
voltage components V
a
and V
a
(being sum and difference,
respectively of positive- and negative-sequence voltage). For a dip of
type A, V
a
and V
a
are equal; for a dip of type C, V
a
is equal to the
pre-fault voltage and V
a
is dependent on the distance to the fault. For a
dip of type D the situation is the other way around. In the ideal case,
the three dip types fall on the straight lines in Figure 3.7(a). Due to the
various assumptions, they fall in the three different areas in Figure
3.7(b). In Figure 3.7 (b), the normal operation and interruption are also
included as voltage dips in general. The determined value is
characteristic voltage V (V
a
in type D and V
a
in type C). The normal
operation state corresponds to voltage dips (Type A, C, D) where
characteristic magnitude |V| is bigger than 90%. The interruption
corresponds to voltage dips (Type A, C, D) where characteristic
magnitude |V| is smaller than 10%.
The accuracy of the proposed method of dip classication depends on
the correctness of the following three assumptions:
type C
type D
type A
type C
type D
type A
V
a
V
a
V
a
V
a
Figure 3.7 Definition of dip types (a) Dip types in ideal case. (b) Dip types
in general.
(a)
(b)
Normal Operation
Interruption
90%
10%
Chapter 3: Classication of Three-phase Voltage Dips
28
1) Zero-sequence voltages do not affect equipment operation.
2) Positive- and negative- sequence source impedances dont differ
much.
3) 2LGF faults are rare.
Field measurements, later shown in Chapter 5, show these assumptions
are quite acceptable in reality. Under these assumptions, any three-
phase voltage dips can be characterized by one phasor, namely the
characteristic voltage V. This conclusion greatly simplies analysis of
three-phase unbalanced dips. In Chapter 5, the accuracy of the
classication will be further discussed and assessed by eld measured
data.
From expressions of the characteristic voltage V for four different
types of faults, we notice that, if the fault places are the same, dips
from LLF, 3F, 2LGF have the same characteristic voltage V at PCC,
dips from SLGF have a larger characteristic voltage because of the
presence of the zero-sequence impedance in the sequence network of
the characteristic voltage V (V
a
). The presence of the zero-sequence
impedance also inuences the sequence network of the PN-factor F
(V
a
) for the dips due to 2LGF.
3.3.6 Phase-angle shift in unbalanced dips
In Section 3.1, we discussed the phase-angle shift associated with a
balanced voltage dip. We concluded that the phase-angle shift in
balanced dips is caused by the X/R ratio difference between the source
and the faulted feeder. Because the X/R ratio of the feeder is generally
smaller than the X/R ratio of the source in distribution systems, we
expect negative phase-angle shifts during a balanced dip. This will be
conrmed by eld measurements in Chapter 5.
The X/R ratio difference has the same effect on the characteristic
voltage for unbalanced dips. The argument of the characteristic voltage
is non-zero due to a difference of X/R ratio between the source and the
feeder. Dips caused by SLGF show a slightly different behavior since
the zero-sequence source impedance becomes part of the faulted
feeder. This effect often results in a smaller argument of the
characteristic voltage.
Besides the effect of X/R ratio difference, there is another effect
causing phase-angle shift in unbalanced dips. As Figure 3.6 shows,
while two faulted phases of type C dip tend to come closer to each
Chapter 3: Classication of Three-phase Voltage Dips
29
other, two unfaulted phases of type D tend to go further from each
other. The resulted phase-angle shift of phase voltages in unbalanced
dips is an aggregation of these two effects. Substituting the magnitude
and argument of characteristic voltage as a complex number in
expression (3.13) and (3.16), gives the nal phase-angle shift on phase
voltages.
3.3.7 Symmetrical phase for unbalanced dips
In previous sections, we only dened unbalanced dips with phase a as
the symmetrical phase, i.e. the fault at phase a for SLGF and a fault
between phases b and c for LLF and 2LGF. With the consideration of
phase b and phase c as symmetrical phases for Dips of type C and D,
the classication results in six different types of dips namely C
a
, C
b
,
C
c
, D
a
, D
b
, D
c
with the subscript indicating the symmetrical phase.
Table 3.1 gives the mathematical expressions and Figure 3.8 gives their
phasor diagrams. The mathematical expressions for unbalanced dips
with symmetrical phase b and c are derived by rotating the three-
phases of unbalanced dips by 240
0
and 120
0
respectively. In Table 3.1
and Figure 3.8, we assume the PN-factor F of type C and D equals to 1
to simplify the expressions.
Chapter 3: Classication of Three-phase Voltage Dips
30
a
b
c
a
b
c
a
b
c
a
b
c
a
b
c
a
b
c
Type C
a
Type C
b
Type C
c
Type D
a
Type D
b
Type D
c
Figure 3.8 Phasor diagram of unbalanced dips with the consideration of
symmetrical phases. PN-factor F = 1.0, zero-sequence voltage
V
a0
= 0.0, characteristic voltage V = 0.50
0
. Dashed line: the
pre-fault phase voltage; solid line: phase voltages during dip.
Chapter 3: Classication of Three-phase Voltage Dips
31
Table 3.1: Phase voltages for unbalanced dips
Table 3.2 gives the symmetrical component expressions for the dips in
Table 3.1 by using equation (3.9).
Type C
a
Type D
a
Type C
b
Type D
b
Type C
c
Type D
c
V
a
1 =
V
b
1
2
---
1
2
--- j 3V =
V
c
1
2
---
1
2
--- j 3V + =
V
a
V =
V
b
1
2
---V
1
2
--- j 3 =
V
c
1
2
---V
1
2
--- j 3 + =
V
a
1
4
---
3
4
---V
1
4
--- j 3
1
4
--- jV 3 + + =
V
b
1
2
---
1
2
--- j 3 =
V
c
1
4
---
3
4
---V
1
4
--- j 3
1
4
--- jV 3 + + =
V
a
1
4
---V
3
4
---
1
4
--- j 3
1
4
--- jV 3 + + =
V
b
1
2
---V
1
2
--- j 3V =
V
c
1
4
---V
3
4
---
1
4
--- j 3
1
4
--- jV 3 + + =
V
a
1
4
---V
3
4
---
1
4
--- j 3
1
4
--- jV 3 + + =
V
b
1
4
---
3
4
---V
1
4
--- j 3
1
4
--- jV 3 =
V
c
1
2
---
1
2
--- j 3 + =
V
a
1
4
---
3
4
---V
1
4
--- j 3
1
4
--- jV 3 + + =
V
b
1
4
---
3
4
---V
1
4
--- j 3
1
4
--- jV 3 =
V
c
1
2
---V
1
2
--- jV 3 + =
Chapter 3: Classication of Three-phase Voltage Dips
32
Table 3.2: Symmetrical components for the six types of dips
From Table 3.2 it follows that the positive-sequence voltage is always
along the reference phase axis. The direction of the negative-sequence
voltage depends on the type of dip. We also notice that the negative-
sequence voltages have opposite direction for type C and type D dips
with the same symmetrical phase. This is consistent with the denition
of the characteristic voltage, which is dened as the subtraction of
positive- and negative-sequence voltages for type C
a
but the addition of
positive- and negative-sequence voltages for type D
a
.
Under the generalised denition, these six types of dips have the same
characteristic voltage. By rotating the negative-sequence voltage over
an integer multiple of 60
0
all dips can be obtained from one prototype
dip; dip type C
a
has been chosen as the prototype dip. Thus the
characteristic voltage is obtained by the subtraction of positive and
negative sequence voltage of C
a
. Due to the same reason, these six
types of dips have the same PN-factor, where it is considered. The PN-
factor is obtained by the addition of positive- and negative- sequence
voltages of C
a
.
Figure 3.9 shows the phasor diagram for positive- and negative-
sequence voltages of the six types of unbalanced dips. While the
positive-sequence voltage is the same for the six types of dips, the
argument of the negative sequence determines which type the dip is.
Type C
a
Type D
a
Type C
b
Type D
b
Type C
c
Type D
c
V
1
1 V +
2
------------- =
V
2
1 V
2
------------- =
V
1
1 V +
2
------------- =
V
2
1 V
2
------------- =
V
1
1 V +
2
------------- =
V
2
a
1 V
2
------------- =
V
1
1 V +
2
------------- =
V
2
a
1 V
2
------------- =
V
1
1 V +
2
------------- =
V
2
a
2 1 V
2
------------- =
V
1
1 V +
2
------------- =
V
2
a
2 1 V
2
------------- =
Chapter 3: Classication of Three-phase Voltage Dips
33
The algorithm of recognizing dip type from eld measurements will be
studied in detail in Chapter 5.
3.4 Dip transformation through transformers
Transformers come with many different winding connections. As a
voltage dip passes through a transformer, the phasors relation of the
voltage dip at the secondary side of the transformer will becomes
different compared to the voltage dip at the primary side. In the
concept of dip classication, the type of dip could change. In this
section, we intend to build the mathematic models, in matrix form, for
various types of transformers in this section for modelling the change,
that is, the phasors of the voltage dip at the secondary side shall be
obtained by multiplying the phasors of the voltage dip at the primary
side with the matrix of the transformer. The load current is ignored in
the analysis.
The modelling procedure is taken in two steps:
C
b
C
c
D
a
C
a
D
b
D
c
V
a1
a
b
c
V
a2
Figure 3.9 Phasor Diagram of unbalanced dips with symmetrical
components, PN-factor F = 1, characteristic voltage V = 0.5
30
0
. Dashed line: pre-fault phase voltages.
Chapter 3: Classication of Three-phase Voltage Dips
34
1) The matrix model of the transformer changes voltage phasors, while
retaining the symmetrical phase of the voltage dip. These are called
basic transformer models.
2) Only the symmetrical phase of the dip is changed. These are called
symmetry changers.
In Section 3.4.4, We relate the basic transformer model and symmetry
changer to the physical transformers.
3.4.1 Basic transformer models
Three basic transformer models can be distinguished, based on [6]:
1. Transformers where each of the secondary voltages is the difference
between two primary voltages. These kinds of transformers include Dy,
Yd, and Yz connected transformers.
These transformers can be dened mathematically in matrix form, as
follows:
(3.23)
Each phase at the secondary of the transformer is a subtraction of two
phases at the primary side of this kind of transformer. The factor is
aimed at changing the base of the pu. values. The j is introduced so that
the symmetrical phase of the dip is maintained.
In symmetrical components, the positive-sequence voltage doesnt
change through such a transformer model, but the transformer model
reverses the direction of the negative-sequence voltage. Thus a type C
dip will change into a type D dip, and vice versa. This can be illustrated
by the following calculations:
Given a set of positive-sequence voltage V
1
(3.24)
T
1
j
3
-------
0 1 1
1 0 1
1 1 0
=
3
V
1
V
a
2
V
aV
=
Chapter 3: Classication of Three-phase Voltage Dips
35
and a set of negative-sequence voltage V
2
(3.25)
It follows that T
1
*V
1
= V
1
, but T
1
*V
2
= -V
2
.
The zero-sequence voltage is removed, given
(3.26)
it follows that T
1
*V
0
= 0.
2. Transformers that only remove the zero-sequence voltage. Examples
of this type are the star-star connected transformer with one or both star
points not grounded (Yny, Yyn), and the delta-delta connected
transformer (Dd). Also the delta-zigzag (Dz) transformer ts in this
category.
This type can be dened mathematically in matrix form, as follows:
(3.27)
Each phase at the secondary side of the transformer is obtained by
subtracting the primary side voltage by the zero-sequence voltage, e.g.
(3.28)
In symmetrical components, such a transformer model changes neither
the positive-sequence voltage nor the negative-sequence voltage, that
is, T
2
*V
1
= V
1
, and T
2
*V
2
= V
2
. But the zero-sequence voltage is
removed, that is, T
2
*V
0
= 0.
V
2
V
aV
a
2
V
=
V
0
V
V
V
=
T
2
1
3
---
2 1 1
1 2 1
1 1 2
=
V
a

V
a
1
3
--- V
a
V
b
V
c
+ + ( ) =
Chapter 3: Classication of Three-phase Voltage Dips
36
3. Transformers that do not change anything to the voltage. For this
type of transformer the secondary-side voltages (in pu.) are equal to the
primary-side voltages (in pu.). The only type of transformers for which
this holds is the star-star connected one with both star points grounded
(Ynyn).
The mathematical expression for this kind of transformer is simple, it
is written as follows:
(3.29)
In symmetrical components, this type of transformer also doesnt
change anything, that is, T
2
*V
1
= V
1
, T
2
*V
2
= V
2
, and T
2
*V
0
= V
0
.
The effect of three transformer types on sequence voltages can be
summarized in Table 3.3.
Table 3.3: The effect of different types of transformer on sequence voltages
3.4.2 Effect of the basic transformer models on the basic dip types
From the analysis of the previous section, T
2
and T
3
do not affect
positive- and negative-sequence voltages. Voltages of the basic dip
types do only contain positive- and negative-sequence voltages. Thus
T
2
and T
3
will not affect the voltages of the basic dip types. T
1
changes
the direction of the complex negative-sequence voltage. From Table
3.2, we concluded that the difference between C and D is the sign of
the negative-sequence voltage. The effect of T
1
is thus that type C and
D change into each other.
This can be mathematically expressed as follows:
Transformer type Effect on sequence voltages
V
0
V
1
V
2
T
1
0 V
1
-V
2
T
2
0 V
1
V
2
T
3
V
0
V
1
V
2
T
3
1 0 0
0 1 0
0 0 1
=
Chapter 3: Classication of Three-phase Voltage Dips
37
Given
(3.30)
and
(3.31)
and
(3.32)
Table 3.4 gives the transformation of dips through three kinds of
transformers by multiplying a dip with the matrix model of
transformers, e.g. T
1
*D
a
, T
2
*C
a
, etc. Note that the symmetrical phase
is not affected by the basic transformer models.
Table 3.4: Transformation of dip types through transformer
Transformer
Connection
Dip type on primary side
A C
a
C
b
C
c
D
a
D
b
D
c
T
1
: Yd, Dy, Yz A D
a
D
b
D
c
C
a
C
b
C
c
T
2
: Yy, Dd, Dz A C
a
C
b
C
c
D
a
D
b
D
c
T
3
: YNyn A C
a
C
b
C
c
D
a
D
b
D
c
A
V
1
2
---V
1
2
--- jV 3
1
2
---V
1
2
--- jV 3 +
=
D
a
V
1
2
---V
1
2
--- j 3
1
2
---V
1
2
--- j 3 +
=
C
a
1
1
2
---
1
2
--- jV 3
1
2
---
1
2
--- jV 3 +
=
Chapter 3: Classication of Three-phase Voltage Dips
38
Any contents of zero-sequence voltages in type D
a
or C
a
will be
removed by the type T
2
transformer but not be affected by T
3
. A type
T
1
transformer changes a type C dip to a type D dip and vice versa
since it reverses the direction of the negative-sequence voltage.
3.4.3 Change of the symmetrical phase
The basic transformer models developed in Section 3.4.2 make sure
that the transformer doesnt affect the symmetrical phase of the voltage
dip through the transformer. But in reality, a dip could change its
symmetrical phase because of the labelling of the phases at the
secondary side of the transformer (often indicated by the clock
number). e.g. phase A at primary side may corresponds phase b at
secondary side in Wye-Wye connection, or phase A at primary side
may correspond to phase a-b at secondary side in Wye-delta. We will
use a so-called symmetry changer to represent the change of
symmetrical phase due to a transformer. Symmetry changers can be
classied into three types:
1. Those for which the symmetrical phase rotates clockwise. Phase a,
b, c change to phase b, c, a, respectively.
2. Those for which the symmetrical phase rotates in counter-clockwise.
Phase a, b, c change to phase c, a, b, respectively.
3. Those that dont change the symmetrical phase.
They can be written in matrix form, as follows:
(3.33)
(3.34)
S
1
a
2
0 0 1
1 0 0
0 1 0
=
S
2
a
0 1 0
0 0 1
1 0 0
=
Chapter 3: Classication of Three-phase Voltage Dips
39
(3.35)
where a = e
j120
. The operator a and a
2
in S
1
and S
2
are to make sure
that the pre-fault voltage, at primary and secondary side of transformer,
shows no phase shift. It has a similar function as the j in the expression
for T
1
. Table 3.5 shows the change of symmetrical phase through a
symmetry changer for the different types of dip.
Table 3.5: Symmetry change of dip types through symmetry changer
In symmetrical components, these symmetry changers dont affect the
positive-sequence voltage, but change the direction of negative-
sequence voltage differently. S
1
rotates the negative-sequence by 120
0
,
and S
2
rotates the negative-sequence by 240
0
. S
3
doesnt rotate it at all.
The zero-sequence is rotated 240
0
by S
1
and 120
0
by S
2
respectively,
and it isnt changed by S
3
.
Mathematically, the effect of a symmetry changer on the symmetrical
components can be put into similar expressions as for the transformer
models, as shown in Table 3.6.
Table 3.6: The effect of different types of transformer on sequence voltages
Symmetry
changer type
Dip type at primary side
A C
a
C
b
C
c
D
a
D
b
D
c
S
1
A C
b
C
c
C
a
D
b
D
c
D
a
S
2
A C
c
C
a
C
b
D
c
D
a
D
b
S
3
A C
a
C
b
C
c
D
a
D
b
D
c
Transformer type Effect on sequence voltages
V
0
V
1
V
2
S
1
a
2
V0 V
1
aV
2
2
S
2
aV0 V
1
a
2
V
2
S
3
V
0
V
1
V
2
S
3
1 0 0
0 1 0
0 0 1
=
Chapter 3: Classication of Three-phase Voltage Dips
40
Any physical transformer can be represented by a combination of a
basic transformer model from Table 3.4 and a symmetry changer
model from Table 3.5. From Table 3.4 and Table 3.5, we conclude dip
type A is not affected by any transformer connection. This is quite
straightforward and easy to understand, because a balanced dip keeps
the balance after a delta-star connection and it doesnt have phase
symmetry. While unbalanced dips may be changed from one type into
another depending on the transformers connection, no new type of
dips is generated.
The concept of transformer model and symmetry changer also applies
on the load connection and monitor connection. A delta-connected load
experiences a type C dip if a star-connected load experiences a type D
dip and vice versa.
3.4.4 Physical transformers to mathematic models
Setting up a mathematical model for a Yy or Dd transformer presents
no special problems, since the voltages are not affected apart from a
possible change in symmetrical phase. Transformer T
2
and symmetry
changer S
1
, S
2
, S
3
are the mathematical model. For a Ynyn transformer
where both sides are grounded, transformer T
3
and symmetry changer
S
1
, S
2
, S
3
are used. We examine a Ynyn transformer as an example.
A
B
C
c
a
Figure 3.10 A Ynyn-connected transformer bank. The secondary line-to-
neutral voltage leads the primary line-to-neutral voltage by
120
0
.
A
B
C
b
b
c
a
Chapter 3: Classication of Three-phase Voltage Dips
41
Figure 3.10 shows an Yy-connected transformer bank, where the
phases are relabled at the secondary side. The voltage phasor diagrams
at the primary and secondary side are also shown in the gure. The
phasor diagram shows that the secondary line-to-neutral voltage leads
the primary line-to-neutral voltage by 120
0
.
If the primary voltage is used as the reference, the above transformer
can be written mathematically as
(3.36)
In voltage dip study, as we mentioned before, the pre-fault voltage,
instead of the primary voltage, is used as the reference. That also
means, the positive-sequence voltage will not rotate through the
transformer. In (3.36), there is obvious a 120
0
clockwise rotation of
positive-sequence voltage. Thus an operator a
2
, that is 120
0
counter-
clockwise rotation, has to be introduced. Compared to the symmetry
changer model described in Section 3.4.3, this is the same as a S
1
symmetry changer model.
The transformer shown in Figure 3.10 is also called Ynyn-8
transformer. The meaning of this notation is as follows: the secondary
voltage delays the primary voltage by 240
0
(or leads by 120
0
), like
eight oclock on a clock. Similarly, by changing the transformer
connection, other clock numbers yield, as shown in Figure 3.11.
V
a
V
b
V
c
T
V
A
V
B
V
C
with , =
T
0 0 1
1 0 0
0 1 0
=
Chapter 3: Classication of Three-phase Voltage Dips
42
Similar to the analysis of Ynyn-8, the other transformer clock number,
like Ynyn-0, Ynyn-2, Ynyn-4, Ynyn-6, Ynyn-10 can be also modelled
by a symmetry changer and basic transformer type T
3
.
Table 3.7 gives the representation physical transformer by
mathematical models in different cases. From dip transformation point
of view, Ynyn-0 and Ynyn-6, Ynyn-2 and Ynyn-8, Ynyn-4 and Ynyn-10
are the same respectively. The modelling of Yy, Dd, obviously follows
the same procedure.
Table 3.7: Mathematical model of Yy, Dd for dip transformation
For a Yd or Dy transformation, the analysis is less straightforward. We
examine as an example a Dy transformer. The treatment of a Yd-
transformer proceeds along similar lines.
clock-0 clock-2 clock-4 clock-6 clock-8 clock-10
Yy T
2
*S
3
T
2
*S
1
T
2
*S
2
T
2
*S
3
T
2
*S
1
T
2
*S
2
Dd T
2
*S
3
T
2
*S
1
T
2
*S
2
T
2
*S
3
T
2
*S
1
T
2
*S
2
Ynyn T
3
*S
3
T
3
*S
1
T
3
*S
2
T
3
*S
3
T
3
*S
1
T
3
*S
2
Ynyn-0
Ynyn-4 Ynyn-8
Figure 3.11 Ynyn transformer clock numbers.
Ynyn-2
Ynyn-6
Ynyn-10
Chapter 3: Classication of Three-phase Voltage Dips
43
Figure 3.12 shows an Dy-connected transformer bank and the phasor
diagrams at the primary and secondary side. The phasor diagram shows
that the secondary line-to-neutral voltage delays the primary line-to-
neutral voltage by 30
0
.
If the primary voltage is used as the reference, the above transformer
can be written mathematically as
(3.37)
In voltage dip study, as we mentioned before, the pre-fault voltage,
instead of the primary voltage, is used as the reference. That also
means, the positive-sequence voltage will not rotate through the
A
B
C
a
b
c
Figure 3.12 A Dy-connected transformer bank. The secondary line-to-
neutral voltage delays the primary line-to-neutral voltage by
30
0
.
A B
C
a
b
c
V
a
V
b
V
c
T
V
A
V
B
V
C
with , =
T
1
3
-------
1 0 1
1 1 0
0 1 1
=
Chapter 3: Classication of Three-phase Voltage Dips
44
transformer. Applying the transformer model T of (3.37) on the
positive-sequence voltage V
1
of (3.24), we get
T*V
1
= e
-j30
*V
1
(3.38)
The positive-sequence voltage rotates -30
0
, which is the same as shown
by the phasor diagram in Figure 3.12.
To keep the positive-sequence voltage unchanged, a factor e
j30
has to
be introduced in (3.37), resulting in (3.39)
(3.39)
Applying the modied transformer model in (3.39) on the negative-
sequence voltage V
2
of (3.25), we get
T*V
2
= e
-j60
*V
2
(3.40)
Comparing with Figure 3.9, we nd that, if we keep the positive-
sequence voltage and rotate the negative-sequence voltage over -60
0
, a
type C dip will change into a type D and vice versa. Besides, the
symmetrical phase a, b, c will change to c, a, b, respectively. This is
equivalent to a combination of transformer type T
1
and a symmetry
changer S
2
.
The transformer shown in Figure 3.12 is also called Dy-1 transformer.
The meaning of this notation is as follows: the secondary voltage
delays the primary voltage by 30
0
, like one oclock on a clock.
Similarly, by changing the transformer connection, other clock
numbers yield, as shown in Figure 3.13.
T
e
j30
3
----------
1 0 1
1 1 0
0 1 1
=
Chapter 3: Classication of Three-phase Voltage Dips
45
From symmetrical components point of view, these transformers rotate
the positive-sequence voltage by 30
0
, 90
0
, 150
0
, 210
0
, 270
0
, 330
0
clockwise and thus rotates the negative-sequence voltage over the same
angle but in the other direction. If we keep the positive-sequence
unchanged, the negative-sequence will rotate by double the angle. By
comparing with Figure 3.9, Table 3.8 lists the corresponding
mathematical model for voltage dip transformation through different
Dy transformers.
Table 3.8: Mathematical models of Dy transformers
From Table 3.8, we can nd that the Dy transformer exchange C and D
type dips, and the different clock number only affects the
symmetrical phase. From dip transformation point of view, Dy-1 and
Dy-7, Dy-3 and Dy-9, Dy-5 and Dy-11 are the same respectively.
Following the similar analysis, we list the results in Table 3.9 for Yd
transformer. It shows the same result as Dy transformers.
Table 3.9: Mathematical model of Yd for dip transformation
Dy-1 Dy-3 Dy-5 Dy-7 Dy-9 Dy-11
T
1
*S
2
T
1
*S
3
T
1
*S
1
T
1
*S
2
T
1
*S
3
T
1
*S
1
Yd-1 Yd-3 Yd-5 Yd-7 Yd-9 Yd-11
T
1
*S
2
T
1
*S
3
T
1
*S
1
T
1
*S
2
T
1
*S
3
T
1
*S
1
Dy-1
Dy-3
Dy-5
Dy-7
Dy-9
Dy-11
Figure 3.13 Dy transformer clock numbers.
Chapter 3: Classication of Three-phase Voltage Dips
46
3.5 Terminology: three-phase voltage dips
Three-phase voltage dips: A power quality event in three-phases of
which at least one phase shows a decrease in rms voltage outside the
normal operating range. An equal decrease in three-phases is also
called a three-phase balanced dip, otherwise, three-phase
unbalanced dip.
Phase voltage: The remaining complex voltage of each of the three
phases in a three-phase unbalanced dip. The absolute values are the dip
magnitudes and the arguments are the phase-angle shifts. This can be
either phase-to-ground or phase-to-phase voltages, depending on the
connection of the equipment.
Dip type: Classication of three-phase voltage dips based on the
relations among the three phase voltages. Type A represents a balanced
dip. Type D and C represent three-phase unbalanced dips in which one
or two phases drop, respectively. All three types can be quantied by
one phasor based on some assumptions. If the symmetrical phase is
considered, dips of type C and D can be further classied as six types,
namely C
a
, C
b
, C
c
, D
a
, D
b
, D
c
. Unbalanced dips can change into each
other through different kinds of transformer connections.
Characteristic voltage: The phasor which is used to quantify the
severity of a three-phase voltage dip. For balanced dips, type A, it
equals to the phase voltage. For unbalanced dips, type C and type D, it
denes as the subtraction of positive-sequence voltage and negative-
sequence voltage of type C
a
. By knowing the characteristic voltage and
dip type, one can reproduce the three-phase voltages of the dip with a
reasonable level of accuracy. The absolute value is called the
characteristic magnitude and the argument is called the characteristic
phase-angle shift.
Positive and negative factor (PN-Factor): An additional phasor to
quantify a three-phase unbalanced dip where the systems positive- and
negative-sequence impedance are not equal. It is close to unity if the
dynamic loads are less connected to the grid. The PN-Factor also
indicates the degree to which the phase voltages can be reproduced by
using the characteristic voltage alone.
Zero-sequence component voltage: The addition of three-phase
voltages of three-phase voltage dips. In most cases the zero-sequence
voltage component of a three-phase unbalanced dip is zero, but in some
cases this additional characteristic is needed.
Chapter 4: Voltage Dip Propagation in Power Systems
47
Chapter 4 Voltage Dip Propagation in Power
Systems
One reason that voltage dips have become a serious concern for the
industry is their propagation through the power system. Interruptions
occur only when a protective device actually interrupts the circuit
serving a particular customer. On the other hand, a short-circuit fault
can generate dips over a wide part of the power system.
There are three different types of dip propagation in power systems,
which are shown in Figure 4.1.
(I) Propagation at the same voltage level.
(II) Propagation to a higher voltage level (or: towards the source).
(III) Propagation to a lower voltage level (or: towards the load).
For faults on distribution feeders, propagation types (II) and (III) are
important. For transmission system faults, type (I) and (III) need to be
considered.
In Chapter 3, we analyzed the characteristics of three-phase voltage
dips resulting from different types of system faults, by using a simple
voltage divider model. Based on a method of two-component
symmetrical components and a few assumptions, we classied the dips
I
II
III
III
Figure 4.1 Various types of dip propagation in power systems.
Fault
Chapter 4: Voltage Dip Propagation in Power Systems
48
as different types at the PCC. An important result obtained from the
classication is that all three-phase voltage dips, balanced or
unbalanced, can be quantied by one phasor, namely the characteristic
voltage. The next question is: what will happen to the dip type and
characteristic voltage when the dip propagates from the PCC to the
equipment terminal?
The study of voltage dip propagation is the basic for stochastic
prediction of voltage dips in a certain area. International standards and
publications suggest several methods to predict voltage dips based on
historical fault statistics and the knowledge of dip propagation [3][5].
However, these methods mostly assume a balanced dip. More
information about unbalanced dip propagation is still needed.
In this chapter, we will look at the propagation of voltage dips in
distribution systems and transmission systems. Simulations are
performed on two hypothetical power networks, a radial system and a
loop system. Based on symmetrical component analysis, a method of
studying voltage dips by a single-phase scheme for unbalanced
situation is proposed. Finally, the inuence of dynamic loads, e.g.
induction motors, on the characteristics of dips is discussed.
4.1 Voltage dip propagation in distribution systems
The radial layout is most often used in distribution network design. The
major advantages of the radial layout are that it is simpler and more
economical to install than other types of layout. Figure 4.2 shows an
example radial system, which transports power from 132 kV level
down to the loads connected at different voltage levels. Table 4.1- 4.3
give the parameters of the power source, transformers and feeders. The
total load connected to each bus is approximately 6% of the short
circuit power for that bus. The loads are assumed pure impedance with
power factor 0.9. Such arrangement makes sure that the systems
positive-sequence and negative-sequence source impedances are equal.
The systems parameters are obtained from [38].
Chapter 4: Voltage Dip Propagation in Power Systems
49
Table 4.1: Source impedance for the supply shown in Figure 4.2
Table 4.2: Transformer connections for the supply shown in Figure 4.2
Table 4.3: Feeder data for the supply shown in Figure 4.2
Voltages (kV) X
1
() R
1
() X
0
() R
0
()
132 4.9833 0.082 4.792 0.157
Transformer S
rated
(MVA)
Voltages
(kV)
X (pu.) Transformer
winding
connection
Neutral
grounding at
the star side
T
a
90 132/33 0.1 Ynd-1 solidly
grounded
T
b
23 33/11 0.08 Dyn-11 solidly
grounded
Voltages (kV) X
1
(/km) R
1
(/km) X
0
(/km) R
0
(/km)
132 kV 0.4478 0.176 1.133 0.4
33 kV 0.338 0.156 1.661 0.304
11 kV 0.315 0.117 1.355 0.223
Figure 4.2 A hypothetical radial distribution network.
132 kV
33 kV
11 kV
T
a
T
b
YnD Dyn
Dy
400 V
Chapter 4: Voltage Dip Propagation in Power Systems
50
4.1.1 Voltage dip propagates upwards and downwards
Generally a dip originated from a higher voltage level propagates
downwards to a lower voltage level without main change in magnitude,
but the voltage drop is reduced greatly when it propagates from a lower
level to a higher level. This effect can be illustrated by the following
simple calculation:
Naming X
s132
the short-circuit impedance at the 132kV voltage busbar
in Figure 4.2, X
t
the reactance of the transformer T
a
. For three-phase
short circuit at 33 kV busbar, the characteristic voltage at 33kV equals
zero. The dip magnitude at 132 kV is:
(4.1)
Using the values in Table 4.1 through 4.3, the dip magnitude is
approximately 80%.
Simulations are performed for different types of faults on the feeders at
2km from the 132 kV, 33 kV, 11 kV busbar. Table 4.4 gives the result
for the dips measured at each bus. Figures 4.3 through 4.5 show the
characteristic magnitude and PN-factor at each bus for four different
fault types.
V
132
X
t
X
s132
X
t
+
-------------------------- =
Chapter 4: Voltage Dip Propagation in Power Systems
51
Table 4.4: Simulation Results
*

V
-
-
-

C
h
a
r
a
c
t
e
r
i
s
t
i
c

V
o
l
t
a
g
e
*
*

F

-
-

P
N
-
f
a
c
t
o
r
F a u l t L o c a t i o n
F
a
u
l
t

t
y
p
e
C
h
a
r
a
c
t
e
r
i
s
t
i
c
s

o
f

v
o
l
t
a
g
e

d
i
p
s

m
e
a
s
u
r
e
d

a
t

e
a
c
h

b
u
s
1
3
2
k
V
3
3

k
V
1
1

k
V
D
i
p

T
y
p
e
V
*
F
*
*
V
0
*
*
*
D
i
p

T
y
p
e
V
F
V
0
D
i
p

T
y
p
e
V
F
V
0
1 3 2 k V f e e d e r
S
L
G
F
D
a
0
.
4
1

4
0
1
.
0

0
.
0
0
0
.
2
3

1
7
5
0
C
c
0
.
4
1

4
0
1
.
0

0
.
0
0
0
.
0
D
a
0
.
4
1

4
0
1
.
0

0
.
0
0
0
.
0
L
L
F
C
a
0
.
1
7

1
5
0
1
.
0

0
.
0
0
0
.
0
D
c
0
.
1
7

1
5
0
1
.
0

0
.
0
0
0
.
0
C
a
0
.
1
7

1
5
0
1
.
0

0
.
0
0
0
.
0
2
L
G
F
C
a
0
.
1
7

1
5
0
0
.
6
8

0
0
0
.
2
5

4
0
D
c
0
.
1
7

1
5
0
0
.
6
8

0
.
0
0
0
.
0
C
a
0
.
1
7

1
5
0
0
.
6
8

0
.
0
0
0
.
0
3

F
A
0
.
1
7

1
5
0
-
-
0
.
0
A
0
.
1
7

1
5
0
-
-
0
.
0
A
0
.
1
7

1
5
0
-
-
0
.
0
3 3 k V f e e d e r
S
L
G
F
-
-
1
.
0
1
.
0

0
.
0
0
0
.
0
-
-
0
.
9
8

0
.
2
0
1
.
0

0
.
0
0
0
.
9
7

1
8
0
0
-
-
0
.
9
8

0
.
2
0
1
.
0

0
.
0
0
0
.
0
L
L
F
D
b
0
.
8
7

0
.
6
0
1
.
0

0
.
0
0
0
.
0
C
a
0
.
3
4

1
6
0
1
.
0

0
.
0
0
1
.
0

0
.
0
0
D
b
0
.
3
4

1
6
0
1
.
0

0
.
0
0
0
.
0
2
L
G
F
D
b
0
.
8
7

0
.
6
0
1
.
0

0
.
0
0
0
.
0
C
a
0
.
3
4

1
6
0
1
.
0

0
.
0
0
0
.
5

0
.
0
0
D
b
0
.
3
4

1
6
0
1
.
0

0
.
0
0
0
.
0
3

F
A
0
.
8
7

0
.
6
0
-
-
0
.
0
A
0
.
3
4

1
6
0
-
-
0
.
0
A
0
.
3
4

1
6
0
-
-
0
.
0
1 1 k V f e e d e r
S
L
G
F
D
a
0
.
9
8

0
.
1
0
1
.
0

0
.
0
0
0
.
0
C
c
0
.
9
1

0
.
7
0
1
.
0

0
.
0
0
0
.
0
D
a
0
.
6
9

5
0
1
.
0

0
.
0
0
0
.
1
2

1
6
9
0
L
L
F
C
a
0
.
9
7

0
.
1
0
1
.
0

0
.
0
0
0
.
0
D
c
0
.
8
7

1
.
0
0
1
.
0

0
.
0
0
0
.
0
C
a
0
.
5
5

8
0
1
.
0

0
.
0
0
0
.
0
2
L
G
F
C
a
0
.
9
7

0
.
1
0
0
.
9
9

0
.
0
0
0
.
0
D
c
0
.
8
7

1
.
0
0
0
.
9
6

0
.
4
0
0
.
0
C
a
0
.
5
5

8
0
0
.
8
4

2
0
0
.
1
1

1
2
0
3

F
A
0
.
9
7

0
.
1
0
-
-
0
.
0
A
0
.
8
7

1
.
0
0
-
-
-
-
A
0
.
5
5

8
0
-
-
0
.
0
Chapter 4: Voltage Dip Propagation in Power Systems
52
Figure 4.3 Voltage dips resulting from a fault on a 132kV feeder. (a)
characteristic magnitude vs. measuring site. (_____) SLGF,
(........) LLF, 2LGF, 3F. (b) PN-factor vs. measuring site.
(_____) LLF, SLGF. (........) 2LGF
132kV 33kV 11kV
1.0
0.8
0.6
0.4
0.2
132kV 33kV 11kV
1.0
0.8
0.6
0.4
0.2
C
h
a
r
a
c
t
e
r
i
s
t
i
c

m
a
g
n
i
t
u
d
e

(
p
u
)
Measuring site
P
N
-
f
a
c
t
o
r

(
p
u
)
Measuring site
SLGF
Others
2LGF
Others
(a)
(b)
Figure 4.4 Voltage dips resulting from a fault on a 33kV feeder. (a)
characteristic magnitude vs. measuring site. (_____) SLGF,
(........) LLF, 2LGF, 3F. (b) PN-factor vs. measuring site.
(_____) LLF, SLGF, 2LGF.
132kV 33kV 11kV
1.0
0.8
0.6
0.4
0.2
132kV 33kV 11kV
1.0
0.8
0.6
0.4
0.2
C
h
a
r
a
c
t
e
r
i
s
t
i
c

m
a
g
n
i
t
u
d
e

(
p
u
)
Measuring site
P
N
-
f
a
c
t
o
r

(
p
u
)
Measuring site
SLGF
Others
All
(a)
(b)
Chapter 4: Voltage Dip Propagation in Power Systems
53
Several things are observed from the simulation results:
1. The dip due to a fault at a higher voltage level propagates to lower
voltage level without any change, but it becomes signicantly more
shallow when propagating from lower voltage level to higher level.
2. Dips due to SLGF and LLF have a PN-factor equal to 1, and the PN-
factor doesnt change when the dips propagate in power systems.
2LGF has a non-unity PN-factor (<1). It becomes closer to 1 when the
dip propagates from lower level to higher level.
3. Dips due to LLF, 2LGF, 3F at the same fault position have the
same characteristic voltage, while dips due to SLGF have a higher
characteristic magnitude. This phenomenon has also been noted in
Chapter 3 by symmetrical component analysis resulting in equations
(3.14), (3.17), (3.19), (3.22).
4. The X/R ratio of feeders are smaller than of the system impedance at
each voltage level. The dips always show negative phase-angle shift of
characteristic voltage.
5. Because the transformers in the system are of Delta-Wye type. The
zero-sequence component is always removed when a dip propagates to
another voltage level.
Figure 4.5 Voltage dips resulting from a fault at on 11kV feeder. (a)
characteristic magnitude vs. measuring site. (______) SLGF,
(........) LLF, 2LGF, 3F. (b) PN-factor vs. measuring site.
(_____) LLF, SLGF, 2LGF. (........) 2LGF.
132kV 33kV 11kV
1.0
0.8
0.6
0.4
0.2
132kV 33kV 11kV
1.0
0.8
0.6
0.4
0.2
C
h
a
r
a
c
t
e
r
i
s
t
i
c

m
a
g
n
i
t
u
d
e

(
p
u
)
Measuring site
P
N
-
f
a
c
t
o
r

(
p
u
)
Measuring site
SLGF
Others
2LGF
Others
(a)
(b)
Chapter 4: Voltage Dip Propagation in Power Systems
54
6. The Delta-Wye transformer changes a type D dip to type C and vice
versa. According to the classication of transformer in Chapter 3, T
a
is
a combination of T
1
type transformer and S
1
type of symmetry
changer, and T
b
is a combination of T
1
type transformer and S
2
type of
symmetry changer.
4.1.2 SLGF at medium voltage level
Medium voltage distribution systems, in Europe are often ungrounded
or high-impedance grounded. Voltage dips due to SLGF in such a
system mainly have a zero-sequence component. The characteristic
voltage and PN-factor are both close to one. The zero-sequence
component is removed when the dip propagates through ungrounded
transformers, so that only a small disturbance remains. Figure 4.6
shows the voltage waveform measured at the PCC of 132kV, 33 kV and
11 kV due to SLGF at a 33 kV feeder. The waveform shows that this
kind of dip propagates neither upward nor downward.
Chapter 4: Voltage Dip Propagation in Power Systems
55
4.1.3 Local generation
In studying voltage dip propagation in distribution systems or
performing stochastic voltage dip prediction, we often assume that the
magnitude of the voltage dip will not change when it propagates
towards the load. However, the local generator could change the
situation. A generator near the plant bus not only increases the fault
level but it also gives an extra damping to remote dips[5]. We will
show how the local generation affects the dips characteristic values in
this section.
We study the following case: a generator with short circuit power 10
MVA (approximately 10% of the short circuit power at 11 kV bus) is
connected to the 11 kV bus in the example radial system (Figure 4.2).
An SLGF is simulated at a 132kV feeder (2 km from the PCC). Figure
V
o
l
t
a
g
e

[
p
u
.
]
0 0.05 0.1 0.15 0.2 0.25 0.3
-2.0
-1.0
0.0
1.0
Time [s]
2.0
0.0 0.05 0.1 0.15 0.2 0.25 0.3
0.0 0.05 0.1 0.15 0.2 0.25 0.3
-2.0
-1.0
0.0
1.0
2.0
-2.0
-1.0
0.0
1.0
2.0
132 kV
33 kV
11 kV
Figure 4.6 Voltage dips at each voltage levels due to a fault at medium
voltage level.
Chapter 4: Voltage Dip Propagation in Power Systems
56
4.7 shows the characteristic magnitude and PN-factor with a
comparison to the situation without the local generator.
The characteristic magnitude of the dip increases at each voltage level,
but the PN-factor is not affected by local generation.
4.2 Voltage dip propagation in transmission systems
Voltage dips in transmission systems are mainly due to system faults
on transmission lines. Dips originating in transmission systems
inuence all the connected distribution systems since they can
propagate downward without attenuation.
Transmission systems are normally loop-connected or mesh-
connected, so that the voltage divider model can not directly be
applied. To predict the propagation of voltage dips, computer
simulations are often performed.
Figure 4.8 shows a loop-connected system with ve sites supplied by
the generator at site A. Table 4.5 and Table 4.6 give the parameters of
the system. The source impedance corresponds to a short circuit power
of 8,000 MVA. The transmission lines between the stations are
assumed all of length 50 km.
Figure 4.7 Voltage dips resulted from SLGF fault at 2 km of 132kV feeder.
(a) characteristic magnitude vs. measuring site. (______) with
local generation at 11 kV, (.......) without local generation. (b)
PN-factor vs. measuring site. (_____) with and without local
generation.
132kV 33kV 11kV
1.0
0.8
0.6
0.4
0.2
132kV 33kV 11kV
1.0
0.8
0.6
0.4
0.2
C
h
a
r
a
c
t
e
r
i
s
t
i
c

m
a
g
n
i
t
u
d
e

(
p
u
.
)
Measuring site
P
N
-
f
a
c
t
o
r

(
p
u
.
)
Measuring site
(a) (b)
Chapter 4: Voltage Dip Propagation in Power Systems
57
Table 4.5:Source impedance of the generator
Table 4.6:Transmission lines parameters
Simulations are performed for different types of faults at stations A, B,
C, and D. Figure 4.9 through 4.12 show the characteristic magnitude
and PN-factor at each site for four different types of fault.
Voltage(kV) X
1
() R
1
() X
0
() R
0
()
400 4.912 0.614 19.956 1.332
Voltage (kV) X
1
(/km) R
1
(/km) X
0
(/km) R
0
(/km)
400 0.288 0.016 0.8 0.112
station A
station B station C
station D
station E station F
Figure 4.8 A hypothetical loop-connected transmission network.
50 km
50 km
50 km
50 km
50 km
50 km
Chapter 4: Voltage Dip Propagation in Power Systems
58
Figure 4.9 Voltage dips resulting from faults at site A. (a) Characteristic
magnitude vs. measuring site. (______) SLGF, (........) LLF,
2LGF, 3F(Fall on X-axis). (b) PN-factor vs. measuring site.
(_____) LLF, SLGF, (........) 2LGF.
1.0
0.8
0.6
0.4
0.2
C
h
a
r
a
c
t
e
r
i
s
t
i
c

m
a
g
n
i
t
u
d
e

(
p
u
.
)
Measuring site
1.0
0.8
0.6
0.4
0.2
P
N
-
f
a
c
t
o
r

(
p
u
.
)
Measuring site
A B C D E F A B C D E F
SLGF
Others
Others
2LGF
(a) (b)
Figure 4.10 Voltage dips resulting from faults at site B. (a) characteristic
magnitude vs. measuring site. (______) SLGF, (........) LLF,
2LGF, 3F. (b) PN-factor vs. measuring site. (_____) LLF,
SLGF, (........) 2LGF.
1.0
0.8
0.6
0.4
0.2
C
h
a
r
a
c
t
e
r
i
s
t
i
c

m
a
g
n
i
t
u
d
e

(
p
u
.
)
Measuring site
1.0
0.8
0.6
0.4
0.2
P
N
-
f
a
c
t
o
r
Measuring site
A B C D E F A B C D E F
SLGF
Others
Others
2LGF
(a) (b)
Chapter 4: Voltage Dip Propagation in Power Systems
59
The characteristics of voltage dips in transmission systems show a
similar behavior as dips in distribution systems:
Figure 4.11 Voltage dips resulting from faults at site C. (a) characteristic
magnitude vs. measuring site. (______) SLGF, (.......) LLF,
2LGF, 3F. (b) PN-factor vs. measuring site. (_____) LLF,
SLGF. (........) 2LGF.
1.0
0.8
0.6
0.4
0.2
C
h
a
r
a
c
t
e
r
i
s
t
i
c

m
a
g
n
i
t
u
d
e

(
p
u
.
)
Measuring Site
1.0
0.8
0.6
0.4
0.2
P
N
-
f
a
c
t
o
r
Measuring Site
A B C D E F A B C D E F
SLGF
Others
Others
2LGF
(a) (b)
Figure 4.12 Voltage dips resulting from faults at site D. (a) characteristic
magnitude vs. measuring site. (______) SLGF, (........) LLF,
2LGF, 3F. (b) PN-factor vs. measuring site. (_____) LLF,
SLGF. (........) 2LGF.
1.0
0.8
0.6
0.4
0.2
C
h
a
r
a
c
t
e
r
i
s
t
i
c

M
a
g
n
i
t
u
d
e

(
p
u
)
Measuring Site
1.0
0.8
0.6
0.4
0.2
P
N
-
f
a
c
t
o
r
Measuring Site
A B C D E F A B C D E F
SLGF
2LGF
Others
Others
(a) (b)
Chapter 4: Voltage Dip Propagation in Power Systems
60
1. Dips due to SLGF and LLF have a PN-factor equal to 1, and the PN-
factor doesnt change when the dips propagate through the power
system. 2LGF has a non-unity PN-factor.
2. Dips due to LLF, 2LGF, 3F at the same fault place have the same
characteristic voltage, while dips due to SLGF have a higher
characteristic magnitude.
It shows that the classication method proposed in Chapter 3 is not
affected by the systems structure.
However, in a mesh-connected network, it is hard to predict the
characteristic magnitude when a dip propagates to different stations.
Generally we can observe: The dip characteristic magnitude decreases
when the fault moves nearer to the station busbar.
4.3 A single-phase scheme to study voltage dip
propagation
In power system analysis, a single-phase diagram is often used for
calculation if the three-phase system is balanced. Besides simplifying
the calculations, a single-phase scheme enables visualizing the
analytical results.
To study the propagation of balanced voltage dips, namely dips of type
A, a single-phase scheme can be used since the system remains
balanced. For unbalanced dips, a single-phase scheme can not be used
directly. However, based on the dip classication in Chapter 3, there is
a possibility to use a single-phase scheme for the study of unbalanced
dips.
4.3.1 Characteristic voltage
As we mentioned before, the characteristic voltage is the main
characteristic of three-phase dips. Knowing dip type and characteristic
voltage, the three phase voltages of a dip can be reproduced in most
cases.
From Equation (3.17), (3.19) and (3.22), we concluded that the
characteristic voltage of dips from LLF, 2LGF and 3F are the same if
the fault locations are the same. That also means, that the characteristic
voltages of dips from LLF and 2LGF can be obtained from a balanced
Chapter 4: Voltage Dip Propagation in Power Systems
61
fault at the same fault location. This result is also shown by the
simulations performed in Section 4.1 and 4.2.
For SLGF, the characteristic voltage is different because of the
presence of the zero-sequence impedance in the sequence network.
Figure 4.13 shows the sequence network for the characteristic voltage
for both 3F and SLGF.
Comparing the circuit for 3F and SLGF shown in Figure 4.13, we can
nd out that, by adding a half zero-sequence source impedance (Z
0
/2)
at the fault point, a balanced fault gives the characteristic voltage of
dips from SLGF. The added impedance (Z
0
/2) is the addition of half
the source impedance at PCC and half the zero-sequence impedance of
the feeder ((Z
s0
+Z
f0
)/2).
The zero-sequence impedance at the fault point may be obtained from
the utility. In case it is unknown, a method applicable in radial
networks is suggested to estimate it from historical fault data. The
detailed procedure is described in Appendix A.
-
+
-
P
1
Z
s1
V
F
I
a
2I
a0
=
V
a
V
a0
Z
s0
2
------------
N
0
N
1
2I
a0
I
a
P
0
+
+
-
Z
f 1
Z
f 0
2
--------- +
+
-
-
+
P
1
V
V
F
Z
s1
N
1
I
Z
f 1
Figure 4.13 Comparision of the characteristic voltage circuit. (a) 3F. V is
defined as the characteristic voltage. (b) SLGF. V
a
=V
1
+V
2
is defined as the characteristic voltage.
(a) (b)
Chapter 4: Voltage Dip Propagation in Power Systems
62
As an example, simulations of faults at an 11 kV feeder bus are
performed to verify the theoretical analysis. Figure 4.14 shows the
comparison of a dip from SLGF and a dip from 3F, where a half zero-
sequence source impedance is added at the fault point in Figure
4.14(b). The zero-sequence source impedance is calculated by the
method suggested in Appendix A. In Figure 4.14(b) the two curves fall
over each other, as both methods give the same result.
4.3.2 PN-factor
As another characteristic of three-phase unbalanced dips, the PN-factor
is normally close to unity and doesnt change when the dip propagates
in power systems. An exception is the dip from 2LGF, the PN-factor of
which is less than unity and changes with the propagation. This has
been shown by the simulations performed in Section 4.1 and 4.2.
Although 2LGFs are rare in power systems, we still consider them
when we study unbalanced dips by a single-phase scheme. Figure 4.15
shows the sequence network of characteristic voltage for 3F and the
sequence network of PN-factor for 2LGF.
Figure 4.14 Voltage dips resulting from an 11 kV fault. (a) Characteristic
magnitude vs. measuring site. (______) SLGF, (........) 3F. (b)
Characteristic magnitude vs. measuring site. (______) SLGF,
(.......) 3F with added impedance (Z
0
/2) at fault location (two
curves merged).
132kV 33kV 11kV
1.0
0.8
0.6
0.4
0.2
132kV 33kV 11kV
1.0
0.8
0.6
0.4
0.2
C
h
a
r
a
c
t
e
r
i
s
t
i
c

m
a
g
n
i
t
u
d
e

(
p
u
.
)
Measuring site
Measuring site
C
h
a
r
a
c
t
e
r
i
s
t
i
c

m
a
g
n
i
t
u
d
e

(
p
u
.
)
SLGF
3F
SLGF
3F
(a) (b)
Chapter 4: Voltage Dip Propagation in Power Systems
63
Comparing the circuits for 3F and 2LGF shown in Figure 4.15, we
can nd out that a balanced fault gives the PN-factor of dips from
2LGF at the same fault location, by adding the double zero-sequence
impedance (2Z
0
) at the fault point. This procedure is quite similar to
what we have used to get the characteristic voltage of the dips from
SLGF faults. The double zero-sequence impedance (2Z
0
) is the
addition of double source impedance at PCC and double zero-sequence
impedance of the feeder (2(Z
s0
+Z
f0
)).
Figure 4.16 shows the comparison of PN-factor of a dip from 2LGF
and characteristic magnitude of a dip from 3F, where a double zero-
sequence source impedance(2Z
0
) is added at the fault point for Figure
4.16(b). The simulations are performed at 2 km of 11 kV feeder. In
Figure 4.16(b) the two curves fall over each other, as both cases give
the same result.
+
-
-
+
P
1
V
V
F
Z
s1
N
1
I
Z
f 1
Figure 4.15 Comparision of the characteristic voltage circuit. (a) 3F. V is
defined as the characteristic voltage. (b) 2LGF. V
a
(=V
1
+V
2
)
is defined as the PN-factor of 2LGF.
(a) (b)
+
-
-
+
P
1
V
a
V
F
Z
s1
N
1
I
a
+
-
2V
a0
2Z
s0
N
0
I
a0
P
0
Z
f1
+2Z
f0
Chapter 4: Voltage Dip Propagation in Power Systems
64
4.3.3 Dip type
Besides the characteristic voltage and the PN-factor, the dip type is
another parameter that needs to be considered when studying the
propagation of three-phase unbalanced dips. As we have analyzed in
Chapter 3, the type of dip could change when a dip goes through a
transformer. A physical transformer can be represented by a
combination of a transformer model and a symmetry changer model,
both introduced in Section 3.4.
4.4 Load s influence
The proposed method of three-phase dips classication is based on the
assumption that the positive-sequence source impedance and the
negative-sequence source impedance are equal. In such a situation, the
PN-factor of the dips from SLGF and LLF, which are the sources of
most unbalanced dips, is equal to unity and dont change with dip
propagation. Further neglecting of the zero-sequence voltage, a dip is
quantied by one variable, namely characteristic voltage. This
assumption greatly simplies the analysis of unbalanced dips, which
was shown in Chapter 3 and in the previous sections in this chapter. In
Figure 4.16 Voltage dips resulting from a fault on an 11kv feeder. (a)
(______) PN-factor of dips from 2LGF, (.........) characteristic
magnitude of dips from 3F. (b) (______) PN-factor of 2LGF,
(........) characteristic magnitude of dips from 3F with added
impedance (2Z
0
) at fault location (two curves merged).
132kV 33kV 11kV
1.0
0.8
0.6
0.4
0.2
132kV 33kV 11kV
1.0
0.8
0.6
0.4
0.2
C
h
a
r
a
c
t
e
r
i
s
t
i
c

m
a
g
n
i
t
u
d
e

(
p
u
.
)
Measuring Site
Measuring Site
&
P
N
-
f
a
c
t
o
r
C
h
a
r
a
c
t
e
r
i
s
t
i
c

m
a
g
n
i
t
u
d
e

(
p
u
.
)
&
P
N
-
f
a
c
t
o
r
2LGF
3F
2LGF
3F
Chapter 4: Voltage Dip Propagation in Power Systems
65
a system with large induction motor loads connected, however, the
error of this assumption increases. In the simulations performed in
Section 4.1 and 4.2, we assumed static loads for all the situations. In
reality, dynamic loads, such as induction motors, can not be
represented as a constant impedance. The dynamic loads inuence the
dip characteristics in two ways. The rst effect is that a dip becomes
non-rectangular. That is, a dip takes one or two cycles to reach the
bottom and recovers gradually after the system fault is cleared, which
is known as the post-fault dip. The second effect is for unbalanced dips,
where a non-unity PN-factor often results.
4.4.1 Motor re-acceleration
In a static system with only constant-impedance load, a rectangular dip
can be assumed. That is, the calculation of magnitude and duration of a
dip can be based on two simple assumptions:
1) Due to the short circuit, the voltage drops to a low value
immediately.
2) When the fault is cleared, the voltage recovers immediately.
These assumptions, however, do not hold in the case of a substantial
part of the load consisting of electrical motors like in many industrial
power systems. The motors decelerate during the short circuit. After
fault-clearing they will accelerate again, drawing high current from the
supply, causing a post-fault voltage dip. A measured dip with a long
post-fault component is shown in Figure 4.17. Such a post-fault dip
prolongs the dip duration. Certain equipment which survived at the
during-fault dip may trip at this stage. In dening the voltage dip
shape, the assumption of a rectangular dip no longer holds in case of
large induction motor loads. Taking motor behavior into account will
make it no longer possible to dene a dip simply by its duration and
magnitude.
Chapter 4: Voltage Dip Propagation in Power Systems
66
A method of characterisation of non-rectangular voltage dips is
presented in [36]. The concept is to split the dip into two parts, the
during-fault dip and the post-fault dip. For the during-fault part, a
rectangular dip is assumed, where the deepest point is considered as
the magnitude of during-fault dip and the fault clearing time is
considered as the duration of the during-fault dip. For the post-fault
part, the point where the voltage recovers is considered as the post-
fault dip magnitude, and the period during which the voltage is below a
certain level (e.g. 90%) is considered as the post-fault duration. The
corresponding contour chart for non-rectangular dip is developed in
[39].
4.4.2 PN-factor
In the case of positive- and negative-sequence source impedance are
different, the two-component method for symmetrical components
cant be used. The PN- factor as dened in Chapter 3 will no longer be
equal to unity for SLGF and LLF. The following equations give
expressions for the PN-factor based on a three-component symmetrical
component analysis.
SLGF
(4.2)
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
-1.5
-1
-0.5
0
0.5
1
1.5
Time [s]
V
o
l
t
a
g
e

[
p
u
.
]
Figure 4.17 Measured dip with a clear post-fault component, data obtained
from scottish power.
F 1
Z
s2
Z
s1

Z
s0
Z
s1
Z
s2
+ + ( ) Z
f 0
Z
f 1
Z
f 2
+ + ( ) +
---------------------------------------------------------------------------------------------- + =
Chapter 4: Voltage Dip Propagation in Power Systems
67
LLF
(4.3)
Equations (4.2) and (4.3) show that in case of unequal positive- and
negative-sequence source impedances, the PN-factors of dips due to
SLGF and LLF are no longer equal to unity.
Induction machines are the main reason that positive- and negative-
sequence source impedances are not equal. The induction motor is
usually represented by a transformer equivalent, i.e., a T equivalent
where separate series branches represent the stator (primary) circuit
and the rotor (secondary) circuit, with a shunt branch to represent iron
losses and excitation. At standstill or locked rotor the induction
machine is indeed a transformer. With the motor turning, however, the
equivalent impedance of the secondary or rotor circuit is seen to be a
function of the slip s as noted in Figure 4.18.
If negative-sequence voltages are applied to the induction motor, a
revolving ux wave is established in the machine air gap which is
rotating backwards. Then the slip of the rotor with respect to the
negative-sequence eld is 2-s. The equivalent circuit for negative-
sequence is the same as that of Figure 4.18 with s replaced by 2-s as
shown in Figure 4.19.
F 1
Z
s2
Z
s1

Z
s1
Z
s2
+ ( ) Z
f 1
Z
f 2
+ ( ) +
--------------------------------------------------------------- + =
R
s
E
r1
X
s
R
r
s
------
X
r
X
m
I
s1
I
r1
V
a1
Figure 4.18 Positive-sequence induction motor equivalent circuit
Chapter 4: Voltage Dip Propagation in Power Systems
68
Table 4.7 lists typical parameter values for three-phase induction
motors and Table 4.8 lists the calculated sequence values based on the
equivalent circuit in Figure 4.18 and 4.19. Equation (4.4) and (4.5) are
used for the calculation.
(4.4)
(4.5)
It shows that the ratio of positive-sequence impedance and negative-
sequence impedance for most of the power ratings ranges from 7-9. If
the induction motor loads are large connected to the point-of-
common-coupling (PCC), the positive- and negative-sequence source
impedance will be different. Since the negative-sequence impedance is
always smaller than the positive-sequence impedance, it follows from
(4.2) and (4.3) that the PN-factor will be smaller than one when large
induction motor loads are connected.
R
s
E
r2
X
s
R
r
2 s
-----------
X
r
X
m
I
s2
I
r2
V
a2
Figure 4.19 Negative-sequence induction motor equivalent circuit.
Z
1
jX
m
jX
r
R
r
s
----- +
( ,
j \
jX
m
jX
r
R
r
s
----- + +
-------------------------------------- R
s
jX
s
+ + =
Z
2
jX
m
jX
r
R
r
2 s
----------- +
( ,
j \
jX
m
jX
r
R
r
2 s
----------- + +
-------------------------------------------- R
s
jX
s
+ + =
Chapter 4: Voltage Dip Propagation in Power Systems
69
Table 4.7: Approximate Constants for Three-Phase Induction Motors
source: P.Andersson [12].
Table 4.8: Calculated Sequence Values for Three-Phase Induction Motors
To look at the inuence of induction motor loads on dip characteristics
quantitatively, we simulated SLGF in the simple network shown in
Figure 4.20.
Rating
(HP)
Full load
slip (%)
R and X in per Unit*
X
s
X
r
X
m
R
s
R
r
Up to 5 3.0-5.0 0.05-0.07 0.05-0.07 1.6-2.2 0.04-0.06 0.04-0.06
5-25 2.5-4.0 0.06-0.08 0.06-0.08 2.0-2.8 0.035-0.05 0.035-0.05
25-200 2.0-3.0 0.075-0.085 0.075-0.085 2.2-3.2 0.03-0.04 0.030-0.04
200-1000 1.5-2.5 0.075-0.085 0.075-0.085 2.4-3.6 0.025-0.03 0.020-0.03
over
1000
1.0 0.075-0.085 0.075-0.085 2.6-4.0 0.015-0.02 0.015-0.025
Rating
(HP)
R and X in per Unit* |Z
1
/Z
2
|
R
1
X
1
R
2
X
2
Up to 5 0.80-0.94 0.71-0.60 0.06-0.09 0.10-0.14 9.3-6.8
5-25 0.94-1.04 0.73-0.59 0.05-0.07 0.12-0.16 9.2-6.9
25-200 1.0-1.13 0.79-0.61 0.04-0.06 0.15-0.17 8.3-7.2
200-1000 1.0-1.07 0.67-0.51 0.03-0.04 0.15-0.17 7.9-6.8
over 1000 1.09-1.76 0.75-1.24 0.02-0.03 0.15-0.17 8.9-12.6
Induction
Motor
Impedance
Load
2 km
SLGF
13.8 kV
Power
Source
Figure 4.20 A simple 13.8 kV network.
E
Z
f
Chapter 4: Voltage Dip Propagation in Power Systems
70
The system has a 600 MVA short circuit level at the 13.8 kV bus; The
total load connected at 13.8 kV bus is 6% of the short circuit level, and
with power factor 0.9. SLGFs are applied on the feeder at 2 km from
13.8 kV bus. The total load is kept the same, however, the proportion of
induction motor load is varied. Figure 4.21 and Figure 4.22 show the
time-domain plotting of PN-factor and characteristic magnitude for the
dips at the 11 kV bus. Five different curves correspond to induction
motor loads of 0%, 30%, 50%, 70%, and 100%. The trend is quite
clear: the increase of induction motor loads, the PN-factor becomes
lower and the characteristic magnitude becomes more and more non-
rectangular.
Time [s]
V
o
l
t
a
g
e

[
p
u
.
]
Figure 4.21 Plotting of PN-factor in time domain for an unbalanced dip at
11 kV level due to a SLGF. The curves from top to bottom
correpond to induction motor load of 0%, 30%, 50%, 70%,
and 100%.
0.2
0.4
0.6
0.8
1.0
0.0 0.1 0.2 0.3 0.4 0.5
Chapter 4: Voltage Dip Propagation in Power Systems
71
The simulation is performed with constant load but various proportions
of induction motor load. But this doesnt mean PN-factor only depends
on the ratio between impedance load and induction load. An increase
of induction motor load always results in lower PN-factor.
Besides the inequality of positive- and negative- sequence source
impedances, the induction motor loads also have another effect on the
PN-factor. The positive-sequence impedance of the induction motor
decreases gradually during the dip [21]. This effect also contributes to
a lower PN-factor. The left column of Figure 4.24 shows the sequence
networks of the system in Figure 4.20, where Z
s1
, Z
s2
, Z
s0
are
sequence impedances of the power source; Z
l1
, Z
l2
, Z
l0
are sequence
impedances of the total load (impedance + induction motor); Z
f1
,
Z
f2
, Z
f0
are sequence impedances of the feeder between the 13.8 kV
bus and the fault location.
0.0 0.1 0.2 0.3 0.4 0.5
Time [s]
V
o
l
t
a
g
e

[
p
u
.
]
0.2
0.4
0.6
0.8
Figure 4.22 Plotting of characteristic magnitude in time domain for an
unbalanced dip at 11 kV level. The curve from rectangular to
non-rectangular correspond to induction motor load of 0%,
30%, 50%, 70%, and 100%.
1.0
Chapter 4: Voltage Dip Propagation in Power Systems
72
Note the difference with the voltage divider model shown in Section
3.1 in Chapter 3. The load is explicitly included in the circuit. The
standard voltage divider model show in the right column of Figure 4.24
is obtained by the following,
(4.6)
(4.7)
(4.8)
(4.9)
Previously, the analysis was based on E = 1. Load was implicitly
included, but the load was not supposed to change impedance during
the event.
E
Z
s1
Z
l1
Z
f1
Z
s2
Z
f2
Z
l2
Z
l0
Z
s0
Z
f0
PCC
PCC
PCC
Figure 4.24 The sequence network of the system in Figure 4.20.
E
Z
s1
Z
f1
Z
s2
Z
f2
Z
s0
Z
f0
PCC
PCC
PCC
Load model non-load model
E

E
Z
l1
Z
l1
Z
s1
+
----------------------- =
Z
s1

Z
s1
Z
l1
Z
s1
Z
l1
+
----------------------- =
Z
s2

Z
s2
Z
l2
Z
s2
Z
l2
+
----------------------- =
Z
s0

Z
s0
Z
l0
Z
s0
Z
l0
+
----------------------- =
Chapter 4: Voltage Dip Propagation in Power Systems
73
With the speed of the motor slows down during the dip, the slip s
increases. From the equivalent circuit of positive- and negative-
sequence impedance of induction motor loads shown in Figure 4.18
and 4.19, it can be concluded that the slip inuences the sequence
impedance. Thus, the positive-sequence impedance decreases but the
negative-sequence impedance remains constant. The zero-sequence
impedance of the induction motors is innite, thus Z
s0
is not affected
by the increased slip either. As Equation (4.6) shows, the decreased
positive-sequence impedance results in a decreased voltage E.
Physically, we can also understand this effect as another simultaneous
3F fault due to the rather high power consumption because of the
increased slip of the induction motor.
4.4.3 Limitations of the classification method
The proposed classication theory of three-phase voltage dips and the
following theory of their propagations are based on the assumption that
the positive- and negative- sequence source impedances are equal.
Loads are of constant-impedance type with equal positive- and
negative-sequence impedances. The benet of this simplication is in
that it offers an analytical single-phase method for voltage dip studies
even in unbalanced cases. Thus, the theory developed for three-phase
balanced dips can be extended to unbalanced dips with only slight
modication. However, in the case of large induction motor loads
present in the system, the error may increase to such a level that it is
not suitable to use this theory. In this case, we have to turn to
traditional symmetrical components or digital simulations.
Chapter 4: Voltage Dip Propagation in Power Systems
74
Chapter 5: Field Measurement Analysis
75
Chapter 5 Field Measurement Analysis
Power quality surveys play important roles in dening the electrical
environment. Efforts have been made in several nations resulting in
large quantities of survey data. IEEE std 1159 [8] gives practical
instructions for monitoring electric power quality and interpreting the
results. However, voltage dips are only generally described in the
document, where it is characterized by magnitude and duration only.
Some other aspects of voltage dip characteristics, such as phase-angle
shift, point-on-wave, and three-phase unbalance are not included. More
recent publications have introduced different ways in presenting
voltage dips and other voltage variations from power quality surveys
[30][31][32]. This chapter intends to apply the theory of dip
classication to present measured voltage dips. The measuring data are
obtained from different sources [29][33]. The measuring sites range
from low voltage level (230 V), to distribution systems (11kV, 33kV,
132kV), and transmission systems (220kV, 400 kV).
5.1 Obtaining dip characteristics
Based on the dip classication theory proposed in Chapter 3, a three-
phase voltage dip can be fully characterized by four values: dip type,
characteristic voltage, PN-Factor (only for unbalanced dips), and zero-
sequence (only for unbalanced dips). Knowing the four characteristics,
three phase voltages of a dip can be reproduced. Note that dip types
can reach three values (A, C, D), where the other three characteristics
are complex values (6 real values). To fully characterize the dip, we
thus need 7 real values. The advantage of using this characterization to
present voltage dip is that the resulting characteristics can be further
reduced, since PN-factor is normally close to one and zero-sequence
voltage can be assumed zero. Thus, a three-phase voltage dip can be
characterized by one variable, namely characteristic voltage, and its
type. Also, the argument of the characteristic voltage may be neglected
for some applications, so that only the dip type and one real value
remains.
5.1.1 Principle
To obtain the characteristics from recorded voltage waveshapes, the
equations listed in Table 3.2 are used as the mathematical base. From
Table 3.2, it follows that the positive-sequence voltage is always along
Chapter 5: Field Measurement Analysis
76
the reference phase axis. The direction of the negative-sequence
voltage depends on the type of dip. For type C
a
the negative-sequence
voltage is along the positive real axis; for type C
b
the negative-
sequence voltage is leading by 120
0
; for type D
c
by 60
0
, etc. By
rotating the negative-sequence voltage over an integer multiple of 60
0
all dip types can be obtained from one prototype dip; dip type C
a
has
been chosen as the prototype dip. From Table 3.2, the following
relationship between positive- and negative- sequence voltage is
obtained for the prototype dip:
(5.1)
The dip type may be obtained from the angle between the negative-
sequence voltage of the measured dip and the negative-sequence
voltage of the prototype. Due to various approximations made and
measurement errors, this angle is not exactly an integer multiple of 60
0
so that the following expression may be used to obtain the dip type:
(5.2)
k = 0 type C
a
k = 1 type D
c
k = 2 type C
b
k = 3 type D
a
k = 4 type C
c
k = 5 type D
b
Knowing the dip type, the negative-sequence voltage can be calculated
back to the corresponding value for the prototype dip:
(5.3)
where k is obtained according to (5.2) and V
2
the negative-sequence
voltage of the measured dip. Characteristic voltage V and PN-factor F
are obtained from the expressions for the prototype dip:
(5.4)
V
2 ref ,
1 V
1
=
k round
angle V
2
1 V
1
, ( )
60
0
---------------------------------------------
( ,
, (
j \
=
V
2

V
2
e
jk60
0

=
V V
1
V
2

=
F V
1
V
2

+ =
Chapter 5: Field Measurement Analysis
77
This method has been applied to data obtained from various sources.
From the recorded voltage waveforms, positive- and negative-sequence
voltages are calculated. To obtain the dip type, the angle is obtained
between V
2
and 1-V
1
. In Figure 5.1 the negative-sequence voltage V
2
is plotted against the drop in positive-sequence voltage 1- V
1
. The
latter is toward the right in the gure, so that the direction of each point
gives the angle between V
2
and 1-V
1
. The distance between each point
and the origin corresponds to the absolute value of the negative-
sequence voltage.
Figure 5.1 shows the scatter plot of the measured unbalanced dips from
various sources. Each dot represent the negative-sequence voltage V
2
of a measured dip compared to 1-V
1
. When the PN-factor F exactly
equals 1, the scatter dots will all fall on one of the six axes in the gure.
Due to non-unity PN-factor, the dots are spread around these axes, but
in almost all cases, the dots are close to one of the six axes. An
exception is some dots close to the origin of the diagram. These dots
represent shallow dips, where a small error may change the apparent
dip type. We assume that the dots distributed within -30
0
and +30
0
region from an axis have the same dip type as the dots falling on the
axis. Equation (5.2) expresses this principle in a mathematical way.
The measuring data presented in Figure 5.1 shows that,
Figure 5.1 Recognition of dip types, data obtained from Scottish Power,
SINTEF, and STRI.
C
b
D
c
D
a
D
b
C
c
C
a
(1-V
1
)
Chapter 5: Field Measurement Analysis
78
1. Most dots are close to the origin. Those dots represent shallow dips,
which have small negative-sequence voltages. The shallow dips are the
majority of the measurements.
2. The dots within each region appears shifted toward the right side
compared to the ideal case. This phenomenon is due to the
characteristics of the PN-factor. As will be shown later (Figure 5.9 and
Figure 5.11), PN-factor values measured are slightly less than one in
absolute value with a negative argument. In a certain system, it may be
possible to use an PN-factor other than 1.0 to obtain V
2
, so that the
dots will distribute around the average case.
5.1.2 Algorithms for dip characterization
The proposed algorithm for classication and characterization of three-
phase unbalanced dips consists of a number of steps. It is assumed that
time-domain sampled data are available for all the three phases
including at least two cycles of pre-event voltages.
1. Determine the voltage frequency from the pre-event voltage
samples.
2. Determine voltage phasors for the three phase voltages by using a
DFT (Discrete Fourier Transform) algorithm. The voltage frequency is
used to obtain the phase-angle shift between the during-event and the
pre-event voltage.
3. Calculate positive-, negative-, and zero-sequence voltages by using
expression (3.9).
4. Determine if the dip is balanced or unbalanced from the magnitude
of the negative-sequence voltage V
2
compared to the positive-sequence
voltage V
1
. If |V
2
| << |1- V
1
| the dip is balanced.
5. For balanced dips the dip type is A and the characteristic voltage
equals the positive-sequence voltage.
6. For unbalanced dips the dip type is determined from positive- and
negative-sequence voltages by using expression (5.2). The
characteristic voltage V and PN-factor F are obtained by using
expression (5.4).
7. A balanced dip is fully characterized through the characteristic
voltage V.
Chapter 5: Field Measurement Analysis
79
8. An unbalanced dip is fully characterized through the dip type,
characteristic voltage, PN-factor and zero-sequence voltage.
5.1.3 Examples
Two examples of three-phase unbalanced dips obtained from
measurements are shown in Figure 5.2 and Figure 5.3. Due to the dip
in Figure 5.2, the equipment will experience a drop in voltage for one
phase, where the dip in Figure 5.3 will lead to a drop in two phases.
The effect of these two events on the equipment is likely to be
different, but they would be characterized as identical events if only the
lowest phase magnitude and duration are used for characterization.
0.05 0.1 0.15 0.2 0.25 0.3 0.35
-1.5
-0.5
0.5
1.5
V
o
l
t
a
g
e

[
p
u
.
]
Time [s]
Figure 5.2 An example of a three-phase unbalanced dip with one phase
experiencing a severe drop in voltage (waveform I).
Chapter 5: Field Measurement Analysis
80
The proposed characterization method has been applied to the three-
phase unbalanced dips shown in Figure 5.2 and Figure 5.3. Following
the algorithm described in Section 5.1.2, we obtain the characterization
results:
A. Following step 1, the voltage frequency of each event is obtained
from the pre-event voltage samples. The results are: Waveform I:
49.994 Hz, waveform II: 50.073 Hz. These values are used as the
fundamental frequency in performing DFT analysis.
B. Following step 2 and step 3, the voltage phasors are obtained by
DFT, and the symmetrical components are calculated. The
characteristic values and phasor diagram for waveform I are shown in
Figure 5.4 and Table 5.1. The results for waveform II are shown in
Figure 5.5 and Table 5.2. The phasors of the dips are obtained on the
second cycle of the waveform after the dip initiation.
0.05 0.1 0.15 0.2 0.25 0.3 0.35
-1.5
-0.5
0.5
1.5
Time [s]
V
o
l
t
a
g
e

[
p
u
.
]
Figure 5.3 An example of a three-phase unbalanced dip with two phases
experiencing severe drops in voltage (waveform II).
Chapter 5: Field Measurement Analysis
81
Table 5.1: Characteristics of waveform I
Dip
Type
Characteristic
voltage (V)
PN-
factor (F)
Zero-
sequence
(V
0
)
Positive-
sequence
(V
1
)
Negative-
sequence
(V
2
)
Drop in
positive-
sequence
(1-V
1
)
D
c
0.76-1
0
0.98-0
0
0.06-55.8
0
0.87-1
0
0.1260
0
0.136.7
0
Figure 5.4 Waveform I (a) Three-phase phasor diagram; (b) Sequence
phasor diagram;
(a)
(b)
a
b
c
V
a
V
b
V
c
a
c
b
V
1
V
2

Figure 5.5 Waveform II (a) Three phase phasor diagram; (b) Sequence
phasor diagram;
a
b
c
V
a
V
c
V
b
a
b
c
V
1
V
2
(a)
(b)
Chapter 5: Field Measurement Analysis
82
Table 5.2:Characteristics of Waveform II
C. By using (5.1) and (5.2) (step 4 and step 5), the dip types are
determined.
For waveform I we obtained
V
2
= 0.1260
0
, V
1
= 0.87-1
0
, 1-V
1
= 0.136.7
0
The angle between V
2
and (1-V
1
) is 53.3
0
, which according to (5.2)
gives a dip of type D
c
. Note that type D
c
is a major drop in phase C
with minor drops in phases a and b. After rotating the negative-
sequence voltage over -60
0
, according to (5.3), we obtain
V
1
= 0.87-1
0
and V
2
= 0.120
0
Applying (5.4) gives the characteristic voltage
V = V
1
- V
2
= 0.76-1
0
and the PN-factor
F = V
1
+ V
2
= 0.98-0
0
For waveform II the angle between V
2
and 1-V
1
is equal to 111
0
,
resulting in dip type C
b
(a major drop in phases a and c), the PN-factor
F = 0.98-2
0
and the characteristic voltage V = 0.56-4
0
.
D. By using (5.4), the characteristic voltage V and PN-factor F are
calculated, as shown in Table 5.1 and 5.2. Note that the characteristic
magnitude of waveform II is much less (56%) than of waveform I
(76%) despite them having the same lowest phase voltage.
The characteristic magnitude and PN-factor can also be plotted as a
function of time, shown in Figure 5.6 and Figure 5.7. The
characteristics are obtained from a one-cycle moving window.
Several observations can be made from these gures:
Dip
Type
Characteristic
voltage (V)
PN-factor
(F)
Zero-
sequence
(V
0
)
Positive-
sequence
(V
1
)
Negative-
sequence
(V
2
)
Drop in
positive-
sequence
(1-V
1
)
C
b
0.56-4
0
0.98-2
0
0 0.77-3
0
0.21121
0
0.239.9
0
Chapter 5: Field Measurement Analysis
83
1. Neither the characteristic voltage V nor the PN-factor F is
completely constant during the dip, both show small and slow
decreases. This shows the contribution from dynamic loads. The
inuence from dynamic loads on dip characteristics is discussed in
Section 4.4.
2. After fault-clearing (i.e. after the main recovery in voltage), both
characteristic voltage V and PN-factor F still show a small drop with
slow recovery to their pre-event values (1.0 pu.). This is most clearly
visible in Figure 5.7. The post-fault dip is balanced so that
characteristic voltage V and PN-factor F are equal. An explanation for
this phenomenon is given in Section 4.4.
0.05 0.1 0.15 0.2 0.25 0.3
0.2
0.4
0.6
0.8
1.0
Time [s]
V
o
l
t
a
g
e

[
p
u
.
]
Figure 5.6 Characteristic magnitude (lower curve) and PN-factor (upper
curve) as a function of time, for waveform I.
0.35
Chapter 5: Field Measurement Analysis
84
5.2 Characteristics obtained from measurements
As we have mentioned in Section 3.3.5, the accuracy of the theory of
dip classication depends on the correctness of several assumptions.
The assumption that the positive- and negative-sequence source
impedance are equal is most critical. From this assumption, it follows
that the PN-factor of an unbalanced dip is equal to unity and constant
as the dip propagates in the power system. To verify the theory, we
have collected measurement data from several data sources. Following
the algorithm introduced in Section 5.1, the characteristic values are
obtained. The phasor values are obtained by DFT phasor detection on
the second cycle of each dip after the dip initiation.
5.2.1 Transmission system: Sweden
The Swedish Transmission Research Institute (STRI) performed power
quality measurements in the swedish transmission system at four
400/220 kV substations from March 1996 to November 1996. From
the measured dip data, the characteristic voltages and PN-factors of the
dips are obtained and presented as scatter plots in the magnitude-
phase-angle plane. Figure 5.9 shows the characteristic voltage and the
PN-factor of unbalanced dips, namely type C and type D, and Figure
5.10 shows the characteristic voltage of balanced dips, namely type A.
0.05 0.1 0.15 0.2 0.25 0.3
0.2
0.4
0.6
0.8
1
Time [s]
V
o
l
t
a
g
e

[
p
u
.
]
Figure 5.7 Characteristic magnitude (lower curve) and PN-factor (upper
curve) as a function of time, for waveform II.
0.35
Chapter 5: Field Measurement Analysis
85
Figure 5.9 shows that the PN-factors of unbalanced dips in
transmission systems are close to 1. The reason is that the dynamic
0.2 0.4 0.6 0.8 1.0
-40
-30
-20
-10
0
10
20
30
40
Magnitude (pu.)
A
r
g
u
m
e
n
t

(
d
e
g
r
e
e
)
Figure 5.9 Scatter plots of PN-factors and characteristics voltage of
unbalanced dips in argument - magnitude plane, data obtained
from STRI (Dot -- PN-factor, Star -- characteristic voltage).
0.2 0.4 0.6 0.8 1.0
-40
-30
-20
-10
0
10
20
30
40
Magnitude (pu.)
A
r
g
u
m
e
n
t

(
d
e
g
r
e
e
)
Figure 5.10 Scatter plots of characteristic voltages of balanced dips in
argument - magnitude plane, data obtained from STRI.
Chapter 5: Field Measurement Analysis
86
loads are far from the transmission system, thus the positive- and
negative- sequence source impedance are equal. Two dots with
somewhat smaller PN-factors may indicate dips resulted from 2LGF. It
also shows that the arguments of the characteristic voltage for both
balanced and unbalanced dips are quite small. As explained in Chapter
3 and Chapter 4, the argument of the characteristic voltage is due to the
X/R ratio difference of the source and feeder. Since there is no clear
separation of source and feeder in transmission systems, it is consistent
with the theory that the argument is small.
5.2.2 Distribution system: Scotland
Scottish Power implemented an extensive monitoring program to help
identify quality of supply characteristics at a number of sites on its
system. The utility has used the collected data to identify where power
system performance can be improved and where customers can
improve the immunity of their processes to voltage dips. We obtained
from Scottish Power some measured voltage dips to apply the
proposed classication. The data are mostly from distribution systems
(132 kV, 33kV, 11kV). Figure 5.11 shows the characteristic voltages
and PN-factors for unbalanced dips. Figure 5.12 shows the
characteristic voltages for balanced voltage dips.
0.2 0.4 0.6 0.8 1.0
-40
-30
-20
-10
0
10
20
30
40
A
r
g
u
m
e
n
t

(
d
e
g
r
e
e
)
Magnitude (pu.)
Figure 5.11 Scatter plots of PN-factors and characteristics voltages of
unbalanced dips in argument - magnitude plane, data
obtained from Scottish Power (Dot -- PN-factor, Star --
Characteristic voltage).
Chapter 5: Field Measurement Analysis
87
The following observations are made from the measurements:
1. The PN-factors of most measured dips are close to 1 in distribution
systems with a few exceptions. The waveforms of those dips with
lower PN-factors also show long post-fault dips, which indicates the
presence of induction motor loads. Voltage dip propagation inuenced
by induction motor loads, will be shown in Section 5.3.
2.The arguments of the characteristic voltage of both balanced and
unbalanced dips are mostly negative. This is due to the source having a
bigger X/R ratio than the feeders in distribution systems.
3. Several bigger argument of the characteristic voltage of balanced
dips are observed. This could be due to two reasons: 1) Most
unbalanced dips are caused by SLGF, where the argument of
characteristic voltage is smaller. 2)Balanced dips are often caused by
cable fault. The X/R ratio of cable is much smaller than the X/R ratio
of the source.
5.2.3 Distribution system: Norway
The Norwegian Electric Power Research Institute (SINTEF) has
measured voltage dips and other voltage disturbances at over 400 sites
in Norway during November 1992 to June 1996 [29]. Most of these
sites are at 230 V level with a few at 6 kV and 11 kV. The measuring
0.2 0.4 0.6 0.8 1.0
-40
-30
-20
-10
0
10
20
30
40
A
r
g
u
m
e
n
t

(
d
e
g
r
e
e
)
Magnitude (pu.)
Figure 5.12 Scatter plot of characteristic voltages of balanced dips in
argument - magnitude plane, data obtained from Scottish
Power (Star -- Characteristic voltage).
Chapter 5: Field Measurement Analysis
88
period for each site varies from several weeks to a few months. Figure
5.13 shows the characteristic voltages and PN-factors for unbalanced
dips. Figure 5.14 shows the characteristic voltages for balanced voltage
dips.
0.2 0.4 0.6 0.8 1.0
-40
-30
-20
-10
0
10
20
30
40
Figure 5.13 Scatter plots of PN-factor and characteristic voltage of
unbalanced dips in argument - magnitude plane, data
obtained from SINTEF (Dot -- PN-factor, Star -- Characteristic
voltage).
A
r
g
u
m
e
n
t

(
d
e
g
r
e
e
)
Magnitude (pu.)
Chapter 5: Field Measurement Analysis
89
The PN-factors of most measured unbalanced dips from the
measurements are also quite close to 1 (98% in average) in distribution
systems. However, the arguments of characteristic voltages are rather
small and distributed around zero, contrary to the results shown in
Figure 5.11 and 5.12. Two possible explanations are:
1. The measured dips are rather shallow;
2. The difference of X/R ratio between the source and feeders is
small.
Analysis of measured dips, both in transmission and distribution
systems, shows that the PN-factor is close to unity. This indicates that
the assumption, namely positive- and negative- source impedance are
equal, generally holds in distribution systems and transmission
systems.
0.2 0.4 0.6 0.8 1.0
-40
-30
-20
-10
0
10
20
30
40
A
r
g
u
m
e
n
t

(
d
e
g
r
e
e
)
Magnitude (pu.)
Figure 5.14 Scatter plots of characteristic voltage of balanced dips in
argument - magnitude plane, data obtained from SINTEF
(Star -- Characteristic voltage).
Chapter 5: Field Measurement Analysis
90
5.3 Further application examples
5.3.1 A propagating dip
The rst case is a dip caused by a system fault at 132 kV propagating
down to the lower voltage levels. The method is used to analyse the dip
characteristics individually.
Figure 5.15 shows the structure of the system in which the
measurement has been performed. Monitors were located in the system
at 132 kV, 33 kV, and 11 kV, as shown in Figure 5.15.
The recorded results are shown for a dip due to a transmission line fault
at 132 kV and propagates down to 33 kV and 11 kV. The monitors are
installed at three different voltage levels, as shown in Figure 5.15. The
recorded voltage waveforms at the three voltage levels are shown in
Figure 5.16, Figure 5.17, and Figure 5.18. Note that the measured
voltage at 132 kV is phase-to-ground voltage, while the measured
voltage at 33 kV and 11 kV are phase-to-phase connected.
Fault
132 kV
33 kV
11 kV
PCC
Figure 5.15 The system structure. G--the place where the monitors are
installed.
Yd
Dd
132 kV
Chapter 5: Field Measurement Analysis
91
The waveforms shows that a SLGF occurs on phase a of the 132 kV
transmission line. This results in a type D
a
dip on the 132 kV bus. The
0.0 0.05 0.1 0.15 0.2 0.25 0.3
-1.5
-0.5
0.5
1.5
Time (s)
V
o
l
t
a
g
e

(
p
u
)
Figure 5.16 Voltage waveform at 132 kV.
0.0 0.05 0.1 0.15 0.2 0.25 0.3
-1.5
-0.5
0.5
1.5
Time (s)
V
o
l
t
a
g
e

(
p
u
)
Figure 5.17 Voltage waveform at 33 kV.
0.0 0.05 0.1 0.15 0.2 0.25
-1.5
-0.5
0.5
1.5
V
o
l
t
a
g
e

(
p
u
)
Time (seconds)
0.3
Figure 5.18 Voltage waveform at 11 kV.
Chapter 5: Field Measurement Analysis
92
faulted line is disconnected after 4 cycles. The monitoring equipment is
installed at the line side of the circuit breaker, so it also captures the
decreasing voltage waveform after the faulted line is cleared. The rst
4 cycles of the dip propagate down to low voltage level, which results
in a D
b
type dip at 33 kV and a C
c
type dip at 11 kV.
Following the method introduced in Section 5.1, the characteristics of
the measured dips are obtained for the second cycle after dip initiation.
Figure 5.19, Figure 5.20, and Figure 5.21 show the three-phase phasor
diagram and the sequence phasor diagram. Table 5.3, Table 5.4, and
Table 5.5 show the calculated characteristic values of the measured
dips.
Table 5.3: Characteristics of dip at 132 kV
Type Characteristic
voltage (V)
PN-factor
(F)
Zero-
sequence
(V
0
)
Positive-
sequence
(V
1
)
Negative-
sequence
(V
2
)
D
a
0.31-11
0
0.98-2
0
0.14-167
0
0.64-4
0
0.34-178
0
a
c
a
c
b
b
Figure 5.19 132 kV (a) Three-phase phasor diagram; (b) Sequence phasor
diagram;
V
a
V
b
V
c
V
1
V
2
V
0
(a)
(b)
Chapter 5: Field Measurement Analysis
93
Table 5.4: Characteristics of dip at 33 kV
Table 5.5:Characteristics of dip at 11 kV
Type Characteristic
voltage (V)
PN-factor
(F)
Zero-
sequence
(V
0
)
Positive-
sequence
(V
1
)
Negative-
sequence
(V
2
)
D
b
0.40-11
0
0.95-3
0
0 0.68-6
0
0.28-58
0
Type Characteristic
voltage (V)
PN-factor
(F)
Zero-
sequence
(V
0
)
Positive-
sequence
(V
1
)
Negative-
sequence
(V
2
)
C
c
0.54-8
0
0.93-5
0
0 0.74-6
0
0.20-120
0
a a
b
b
c
c

(a) (b)
V
a
V
b
V
c
V
1
V
2
Figure 5.20 33 kV (a) Three phase phasor diagram; (b) Sequence phasor
diagram;
a a
b
b
c
c
(a)
(b)
V
a
V
b
V
c
V
1 V
2
Figure 5.21 11 kV (a) Three-phase phasor diagram; (b) Sequence phasor
diagram;
Chapter 5: Field Measurement Analysis
94
Several things can be observed from this propagating dip by using the
proposed characterization:
1.There is a type 1 transformer (Yd) with a type 1 symmetry changer (a
to b) between 132 kV and 33 kV, which changes the D
a
dip to a C
b
dip.
However, the phase-to-phase connected monitor measurde a dip of
type D
b
instead. A type 1 transformer (Dy) with a type 1 symmetry
changer (b to c) between 33 kV and 11 kV changes the C
b
dip to a D
c
dip, but the phase-to-phase connected monitor measured a C
c
dip
instead. This phenomena shows that, in recognizing dip type from
measurement from eld measurement, it is important to consider the
monitors connection.
2. The zero-sequence voltage is removed from the dip at 33 kV because
of the transformer.
3. An increased characteristic magnitude is observed when the dip
propagates from 132 kV down to 33 kV and 11 kV (31%, 40%, and
54%). This phenomena has also been observed with other measured
propagating dips. It is hard to give a clear explanation without detailed
investigation. However, several possible reasons could be:
1)Embedded generation; 2) Reduction of loads during dips, e.g. trip of
rectier loads; 3) Induction motor loads feeding to the fault due to the
remaining ux [21];
4. The lower PN-factor at 33 kV and 11 kV (0.95 and 0.93) indicates
that large induction motor loads are connected at 33 kV and 11 kV,
which increases the difference between positive- and negative-
sequence source impedance, and make the system no static. The loads
inuence on dip characteristics is studied in Section 4.4.
5. The long tail of the post-fault dip at 33 kV and 11 kV also indicates
the presence of large induction motor loads which results in a balanced
post-fault dip (33kV: V = 0.92-3
0
, 11 kV: 0.90-5
0
immediately after
the dip recovery). When the fault is cleared, motor inrush current
prevents the voltage from returning to normal instantaneously.
5.3.2 Statistics from a power quality survey
The Norwegian Electric Power Research Institute (SINTEF) has
measured voltage dips and other voltage disturbances at over 400 sites
in Norway during November 1992 to June 1996 [29]. From the
recordings, 55 voltage dips from 18 sites were chosen to apply the
proposed method. Most of these sites are at 230 V level with a few at 6
Chapter 5: Field Measurement Analysis
95
kV and 11 kV. The measuring period for each site varies from several
weeks to a few months.
Figure 5.22, Figure 5.23, and Figure 5.24 give the scatter diagram for
different types of dips. Each dip is classied by its type. It is further
characterized with its characteristic magnitude and its duration, and is
plotted as point in the magnitude-duration planes. The duration is
dened as the time during which the characteristic magnitude is lower
than 93% of the pre-fault voltage.
0.0 0.2 0.4 0.6 0.8 1.0
0.2
0.4
0.6
0.8
C
h
a
r
a
c
t
e
r
i
s
t
i
c

m
a
g
n
i
t
u
d
e

(
p
u
.
)
Time (s)
Figure 5.22 Scatter plots of dips of type A (characteristic magnitude vs.
duration).
1.0
Chapter 5: Field Measurement Analysis
96
Compared to the traditional magnitude-duration scatter plots, this
method introduces some new features:
1. The dips are classied by their types. Studies have shown that
certain types of three-phase equipment, such as Adjustable Speed
Drives (ASD), behave quite differently for different types of dip [28].
0.2 0.4 0.6 0.8 1.0
0.2
0.4
0.6
0.8
C
h
a
r
a
c
t
e
r
i
s
t
i
c

m
a
g
n
i
t
u
d
e

(
p
u
.
)
Time (s)
Figure 5.23 Scatter plots of dips of type C (characteristic magnitude vs.
duration).
0.0
1.0
0.0 0.2 0.4 0.6 0.8 1.0
0.2
0.4
0.6
0.8
1.0
C
h
a
r
a
c
t
e
r
i
s
t
i
c

m
a
g
n
i
t
u
d
e

(
p
u
.
)
Time ()
Figure 5.24 Scatter plots of dips of type D (characteristic magnitude vs.
duration).
Chapter 5: Field Measurement Analysis
97
The classication of voltage dips is important to apply the dip
equipment co-ordination chart [11] to three-phase equipment.
2. The characteristic magnitude, instead of lowest phase magnitude, is
used. Using lowest magnitude is based on the assumption that only the
lowest of the three phase voltages has inuence on equipment
behaviour. This is not true for many types of equipment during type C
and type D dips. By using the characteristic magnitude, all three phase
voltages of type C and type D dips can be estimated with reasonable
accuracy.
Although dip type, characteristic magnitude and duration give the
major information of a measured dip, other values, such as PN-factor,
symmetrical phase, zero-sequence voltage are also of importance to
obtain a detailed description of a measured dip. Table 5.6 gives an
example record of voltage dips measured on one site.
Table 5.6: Records of measured voltage dips of a site
Voltage Level: 11 kV
Measuring Period: 01/01/93 - 01/31/93
It should be noticed that the characteristic voltage and the PN-factor
are complex numbers by their denitions; Zero-sequence voltage
doesnt exist in dips of type A and is very rare in dips of type C; Zero-
sequence voltage might be present in dips of type D if the monitor is
phase-to-ground connected. All the dips of type C and D have zero
zero-sequence voltage at this site due to the phase-to-phase connection
of the monitors.
Figure 5.25 shows the frequency of type A, type C and type D during
the measured period of the 18 sites. The result shows that unbalanced
dips, namely type C and type D form the majority of the events. The
No. Date Time Type Duration
(second)
Characteristic
voltage (V)
PN-factor
(F)
Zero-
sequence
(V
0
)
1 01/05/93 19:35:26 C
a
0.10 0.851.5
0
1.0-1
0
0.0
2 01/13/93 05:42:03 D
c
0.08 0.871.0
0
0.99-1
0
0.0
3 01/20/93 23:12:04 A 0.06 0.912.0
0
-- 0.0
4 01/22/93 03:10:17 D
a
0.64 0.881.0
0
0.960.7
0
0.0
5 01/27/93 14:34:59 C
a
0.06 0.892.0
0
1.0-0.7
0
0.0
Chapter 5: Field Measurement Analysis
98
symmetrical phase of them is rather equally spread over the three
phases.
0
2
4
6
8
10
12
14
16
18
20
15
11
5
8
4
4
8
A C
a
C
b
C
c
D
a
D
b
D
c
Dip type
N
u
m
b
e
r
Figure 5.25 Frequency of different types of dip of the survey.
Chapter 6: Equipment Immunity Tests
99
Chapter 6 Equipment Immunity Tests
Process industry equipment can be particularly susceptible to problems
with voltage dips because the equipment is interconnected and a trip of
any component in the process can cause the whole plant to shut down.
Different types of electrical equipment have different voltage
immunity. There is no complete reference for this, but Thomas Keys
CBEMA (Computer Business Equipment Manufacturers Association)
curve is widely quoted as mainframe computers ride-through
ability[1] (shown in Figure 6.1). It is the combination of US Navy
electronic equipment immunity test results, computer power studies,
and unofcial advice of several computer manufacturers regarding
their design standards and test results. The interpretation of the curve is
that a voltage disturbance with a magnitude and a duration between the
two curves will not cause problems to computers. The area between the
two curves stands for the voltage immunity.
IEC 61000-4-11 [10] denes the immunity test methods and range of
preferred test levels for electrical and electronic equipment connected
to low-voltage power supply networks for voltage dips, short
interruptions, and voltage variations. This standard doesnt mention the
term voltage-immunity curve. Instead it denes a number of preferred
magnitudes and durations of sags for which the equipment has to be
0.0001 0.001 0.01 0.1 1 10 100 1000
-100
-50
0
50
100
150
200
250
TIME IN SECONDS
P
E
R
C
E
N
T

C
H
A
N
G
E

I
N

B
U
S

V
O
L
T
A
G
E
8
.
3
3

m
s
OVERVOLTAGE CONDITIONS
UNDERVOLTAGE CONDITIONS
0
.
5

C
Y
C
L
E
RATED
VOLTAGE
ACCEPTABLE
POWER
Figure 6.1 The CBEMA Curve, figure obtained from Jerry Heydt.
Chapter 6: Equipment Immunity Tests
100
tested. The North American Facilities Committee recently published a
standard test methods for voltage dip susceptibility of semiconductor
processing equipment[34]. This test method is intended for the
equipment used in semiconductor factories.
In this chapter, the general concern of dip immunity test on single-
phase equipment and three-phase equipment will be described. The
various characteristics of voltage dips are considered. In performing
three-phase equipment tests, a new concept is introduced based on the
method of three-phase dip classication proposed in Chapter 3.
6.1 Single-phase equipment test
The characteristics of voltage dips in single phase generally can be put
into two categories: energy-related characteristics and non-energy-
related characteristics. A voltage dip characterized by magnitude and
duration implies that the equipment is only sensitive to the energy-
related characteristics. The non-energy-related characteristics include
phase-angle shift, waveform distortion, point-on-wave of initiation and
point-on-wave of recovery. The importance of the different
characteristics is load-dependent. The power supply of a computer, and
of most consumer-electronics equipment normally consists of a diode
rectier along with an electronic voltage regulator. These kinds of
equipment are sensitive to the magnitude and duration of the dip. More
complicated power supplies, such as thyristor-based power supplies
used in dc drives, need accurate zero-crossing information, so phase-
angle shift and waveshapes are important [35]. AC motor contactors
are not as sensitive to phase-angle shifts and waveshape distortion,
however voltage magnitude and point-on-wave of the initiation of the
event plays a key role in tripping the contactor [23]. The point-on-wave
of recovery of the dip might be critical to the power supply with a
capacitor connected at the dc bus. The in-rush current at the recovery
of the voltage depend on the point-on-wave. When the recovery is at
90
0
the instantaneous supply voltage is at maximum and a large current
ows into the capacitor. This might damage the equipment or trip the
current limiting protection.
6.1.1 Test items
A. Magnitude and Duration
Chapter 6: Equipment Immunity Tests
101
Existing international standard documents still only consider
magnitude and duration of voltage dips for equipment testing. Table
6.1 shows the test range of magnitude and duration suggested by [10].
Table 6.1:Preferred Magnitude and Duration
In Table 6.1, the nominal voltage is used as the reference and the rst
column is for testing of interruption. The X is an open duration. This
duration can be given in the product specication.
The magnitude and duration suggested in Table 6.1 are only general
instructions. In obtaining the tolerance curve of a certain piece of
equipment, it is important to nd out the critical point of malfunction.
Computers and most consumer electronics which has a front-end
rectier power supply have an undervoltage protection at the dc bus.
The voltage-tolerance curve is often rectangular [4], as shown in
Figure 6.2. For such equipment, it is important to nd the knee of the
curve.
The voltage tolerance in such a case can also be represented by a set of
two values, like (50ms, 70%), which means V
min
= 70% and t
max
=
50ms in Figure 6.2.
B. Phase-angle shift
Voltage dip and short
interruptions Un%
0 40 70
Duration (in period) 0.5 1 5 10 25 50 X
Figure 6.2 Rectangular voltage-tolerance curve.
100%
V
min
Minimum steady-state voltage
Maximum duration
of zero voltage
Duration
M
a
g
n
i
t
u
d
e
t
max
Chapter 6: Equipment Immunity Tests
102
The inuence of phase-angle shift on equipment varies among
different types of equipment. Those using the phase-angle or zero-
crossings of source voltage as control information may be very
sensitive, for example, some controlled rectiers and voltage source
inverters. To stabilize their operation, these devices usually employ a
PLL (Phase Locked Loop) to synchronize their valve triggering.
Usually, PLLs have time constants, of 500ms or more, while phase-
angle shift caused by faults have a very sharp rate of rise. This makes
PLLs unable to respond during voltage dips.
The cause of phase-angle shift in a single-phase dip has been discussed
in Section 3.1 and Section 3.3.6 for both balanced and unbalanced
dips.
The direction of phase-angle shift is dened such that for positive
phase-angle shift the during-dip waveform leads the prefault
waveform; the negative phase-angle shift indicates a lagging during-
dip waveform, as shown in Figure 6.3. The phase angle should shift
back at the recovery point. A positive phase-angle shift at the dip
initiation corresponds to a negative phase-angle shift at the dip
recovery and vice versa.
The range of phase-angle shift can be theoretically estimated [37]. But
also eld measurements, can give information on the phase angle
values that can be expected. Figure 6.4 shows a scatter plot of
magnitude versus phase-angle shift obtained from eld measurements.
Figure 6.3 Measured voltage dips with phase-angle shift. a) a dip with
positive phase-angle shift, magnitude 50%, phase-angle shift
+28
0
.b) a dip with negative phase-angle shift, magnitude 62%,
phase-angle shift -38
0
, data obtained from Scottish Power.
(b)
(a)
Chapter 6: Equipment Immunity Tests
103
Equipment immunity tests for single-phase equipment should include
voltage dips with positive and negative phase-angle shifts. From the
data presented in Figure 6.4, a range from -40
0
and +20
0
may include
the majority cases.
C. Point-on-wave
The point-on-wave of initiation is where the voltage suddenly drops in
value. The point indicates the starting instant of the fault. The point is
often represented in degrees. As most faults are associated with a
ashover, they are more likely to occur near voltage maximum than
near voltage zero. Electrical contactors were found to be an example of
a device that is extremely sensitive to point-on-wave of initiation. Tests
showed that contactors can withstand a deeper dip when initiated at 90
0
degree than when initiated at 0
0
[13].
An impression of the equipment sensitivity to point-on-wave can be
obtained by performing tests at 0
0
and 90
0
. For equipment highly
sensitive to point-on-wave, the test at intermediate angles should be
performed.
Figure 6.5 gives the denition of point-on-wave, as the angle in
degrees since the last upward zero crossing of the voltage. It is not
clear if equipment immunity will be different for 0
0
and 180
0
(or for
Figure 6.4 Voltage dips characterized by phase magnitude and phase-
angle shift information, data obtained from Scottish Power,
SINTEF, and STRI.
0.0 0.2 0.4 0.6 0.8 1.0
-60
-40
-20
0
20
40
60
Phase magnitude (pu)
P
h
a
s
e
-
a
n
g
l
e

s
h
i
f
t

(
d
e
g
r
e
e
)
Chapter 6: Equipment Immunity Tests
104
90
0
and 270
0
). In case of doubt tests need to be performed for at least
the four angles shown in Figure 6.5. There are four different cases for
the test of 0
0
and 90
0
degree of point-on-wave of initiation, which are
indicated as 0
0
, 90
0
, 180
0
, 270
0
.
Point-on-wave of recovery is where the voltage suddenly rises. It
indicates the instant when the circuit breaker clears the fault. Circuit
breakers often clear a fault at the zero-crossing of the current. The
point-on-wave of recovery thus be somewhat less than 90
0
, related to
the X/R ratio of the source impedance at the fault location. Like the test
of point-on-wave of initiation, the test of point-on-wave of recovery
should at least be performed for 0
0
and 90
0
. A rectier circuit could be
sensitive to point-on-wave of recovery. The dip recovered at 90
0
causes
bigger in-rush current for recharging the capacitor than a dip recovered
at 0
0
. The big in-rush current sometimes trips a piece of equipment
because of the overcurrent protection.
Figure 6.6 and 6.7 show two measured voltage dips. The point-on-
wave of initiation and recovery are indicated on the waveforms.
Figure 6.5 a) point-on-wave at 0
0
; b) point-on-wave at 90
0
; c) point-on-
wave at 180
0
; d) point-on-wave at 270
0
.
(c)
(a)
(b)
(d)
Chapter 6: Equipment Immunity Tests
105
6.1.2 Test setup
IEC 61000-4-11 in its current form only gives requirements for voltage
dip generator. Two examples of dip generators are given in an
informative appendix:
1. Generate the dip by using a waveform generator in cascade with a
power amplier.
2. Use a transformer with two output voltages. Make one output
voltage equal to 100% and the other to the required during-dip
magnitude value. Switch very fast between the two outputs, e.g. by
using thyristor switches.
The two test setups are shown in Figure 6.8 and 6.9. The test setup by
waveform generator needs a power amplier to perform the test. The
Figure 6.6 A measured dip with point-on-wave of initiation at 90
0
and
point-on-wave of recovery at 90
0
, data obtained from Scottish
Power.
Figure 6.7 A measured dip with point-on-wave of initiation at 210
0
and
point-on-wave of recovery at 200
0
, data obtained from Scottish
Power.
Chapter 6: Equipment Immunity Tests
106
advantage is that it can easily generate different kinds of waveform. It
also has the potential to recreate a measured dip. The problem is that
the power amplier becomes very expensive when the rated power of
the Equipment Under Test (EUT) is high.
The transformer-type dip generator is a more practical choice. The
prefault voltage is supplied by the phase voltage as switch 1 is closed
and switch 2 is opened. Closing switch 2 and opening switch 1
simultaneously give the during-dip voltage. The dip magnitude is
determined by the ratio of the transformer. However, with the scheme
shown in Figure 6.9, it is only possible to test magnitude and duration.
Figure 6.8 Waveform-generator-type dip generator suggested by IEC
61000-4-11.
EUT
Measuring
instrument
Controller
Waveform
generator
Power
amplifier
Phase
Neutral
Chapter 6: Equipment Immunity Tests
107
An improved transformer-type dip generator which is also capable of
performing phase-angle shift and point-on-wave test is shown in Figure
6.10. The phase-angle shift is achieved by a three-phase power supply.
The pre-fault voltage is supplied by phase a when switch 1 is closed
and switch 2, and 3 are open. Opening switch 1 and closing switch 2 or
switch 3 gives the during-dip voltage with phase-angle shift. The
during-dip voltage is the sum of two phase voltages. Depending on the
direction and quantity of phase-angle shift, phase a-c or phase a-b
combinations are used. As the phasor diagram in Figure 6.11 shows,
the addition of phase a and phase c gives a dip with positive phase-
angle shift, while the addition of phase a and phase b gives a dip with
negative phase-angle shift. The resulting magnitude and phase-angle
shift are determined by the ratios of the transformers. Compared to the
scheme in Figure 6.9, where two switches are used, three switches are
used in the improved scheme, one for positive phase-angle shift and
one for negative phase-angle shift.
Figure 6.9 Transformer-type dip generator suggested by IEC 61000-4-11.
EUT
Phase
Neutral
Measuring
instrument
switch 1
switch 2
Chapter 6: Equipment Immunity Tests
108
Figure 6.10 An improved transformer-type dip generator which is capable
of performing magnitude, duration and phase-angle shift test.
Phase a
Phase b
Phase c
Neutral
EUT
Measuring
instrument
Z
e
r
o
-
c
r
o
s
s
i
n
g

d
e
t
e
c
t
o
r
I
n
t
e
r
f
a
c
e
C
o
n
t
r
o
l
l
e
r
switch 1
switch 2
switch 3
Figure 6.11 Phasor diagram of generating dips with phase-angle shift.
a
b
c
prefault voltage
Dip with positive
phase-angle shift
Dip with negative
phase-angle shift
Chapter 6: Equipment Immunity Tests
109
6.1.3 Test example
A dip generator set-up as shown in Figure 6.10 was used to perform
tests on a number of low voltage devices[59]. As an example, the test
results for a matrix printer are presented here. Figure 6.12 shows a
simplied conguration of the power supply of the printer. The supply
consisted of a transformer, a rectier and a capacitor.
A single-phase dip with the following characteristics was applied to the
printer:
Magnitude: 70%
Duration: 20 cycles
Point-on-wave of initiation: 0
0
Point-on-wave of recovery:0
0
Phase-angle shift: 0
0
The results are shown in Figure 6.13.
Figure 6.12 Power supply of the printer
230 V ac
rectifier circuit
printer
Chapter 6: Equipment Immunity Tests
110
The ac current consists of two components: the magnetizing current of
the transformer (the sinusoidal part of the pre-dip current) and the load
current owing through the transformer (spikes in the pre-dip current).
During the dip the magnetizing current keeps on owing, though less
than normal. The load current ceases while the capacitor supplies the
printer. When the capacitor voltage has dropped below a certain level,
load current starts to ow through the transformer again. When the
voltage recovers the capacitor is recharged to its normal voltage. This
causes the peaks in the load component of the ac current.
A voltage tolerance curve of the printer under test is produced. The
printer is tripped below the curve. The dip is applied as the printer is
printing. A trip is dened as the stop of printing. The applied dip
characteristics are as following:
Point-on-wave of initiation: 0
0
Point-on-wave of recovery:0
0
Phase-angle shift: 0
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
-2.0
-1.0
0.0
1.0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
22
26
30
34
38
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
-400
-200
0
200
400
V
o
l
t
a
g
e

(
V
)
V
o
l
t
a
g
e

(
V
)
C
u
r
r
e
n
t

(
A
)
AC voltage
AC current
DC voltage
Time (second)
Figure 6.13 The test result of a printer during a 70% voltage dip.
Chapter 6: Equipment Immunity Tests
111
6.2 Three-phase equipment test
Three-phase equipment is widely used in many different industries.
Adjustable speed drives (ASD), and voltage source converters (VSC)
are common three-phase equipment. Besides the dip characteristics
described in Section 6.1 for single-phase equipment, for three-phase
equipment, the characteristics introduced in Chapter 3 should be used:
characteristic voltage, PN-factor, zero-sequence voltage, next to
duration, point-on-wave of initiation and recovery.
Currently used test methods simply combine three test setups for
single-phase equipment to perform the test for three-phase equipment.
However, as the previous chapters have shown, the three phase
voltages during a dip are following certain regularities rather than
taking random values, depending on the type of dip. Randomly
choosing test values has the risk of overestimating or underestimating
the ride-through ability of the equipment under test (EUT).
The method of three-phase voltage dip classication proposed in
Chapter 3 facilitates the understanding of the dip characteristics during
its occurrence and propagation. Compared to the phase voltages, the
characteristics, such as characteristic voltage, PN-factor, and zero-
sequence, are much easier to set a range for equipment test. This range,
200 400 600 800 1000 1200
0
20
40
60
80
100
Time (ms)
M
a
g
n
i
t
u
d
e

(
%
)
Figure 6.14 Voltage tolerance curve of the printer
Chapter 6: Equipment Immunity Tests
112
which can be obtained either from theoretical study or from eld
measurement, provides a test environment closer to the reality.
6.2.1 Test items
A. Dip types
Chapter 3 classies voltage dips experienced by three-phase equipment
as three fundamental types. Type A corresponds to the balanced dip.
Type C and type D correspond to the unbalanced dips. If the
symmetrical phase is considered, type C and type D dips can be further
classied as six types: C
a
, C
b
, C
c
, D
a
, D
b
, D
c
.
The equipment should be tested for the three fundamental types. The
symmetrical phase is not considered if the three phases of the
equipment have equal status. In case additional power is taken from
one or two phases, like the eld winding circuit of a dc ASD, the
equipment should be further tested for the unbalanced dips with
different symmetrical phases.
The three-phase connection of the equipment under test (EUT) should
be indicated in the test result. For unbalanced dip test, a dip of type D
applied to delta-connected equipment has the same effect as a dip of
type C applied to the same equipment with star-connection, and vice
versa.
B. Characteristic voltage and duration
By classifying as different types, a three-phase voltage dip can be
quantied by the characteristic voltage with an acceptable accuracy
level in most cases. The characteristic voltage is a complex number by
its denition. The argument generally reects the X/R ratio difference
between the source and the faulted feeder. For a dip due to a fault in a
transmission system, the argument is often around zero, thus the
characteristic voltage is represented by a real number; for a dip due to a
fault in a distribution system, the argument is mostly negative.
The test should be performed rst for zero argument of the
characteristic voltage. In choosing the characteristic magnitude and
duration, a table similar to Table 6.1 can be used. For unbalanced dips,
a characteristic magnitude of 33% is suggested to be further tested. The
reason is that, for an unbalanced dip due to SLGF, where most
unbalanced dips are generated, the characteristic magnitude is not
likely to be smaller than 33%. This can be explained by (3.14). Since
Chapter 6: Equipment Immunity Tests
113
the zero-sequence source impedance (Z
s0
) generally has a bigger value
than the positive source impedance (Z
s1
), the characteristic magnitude
is bigger than 1/3 of pre-fault voltage even for a SLGF at the PCC (Z
f0
=0, Z
f1
=0).
Table 6.2:Preferred Characteristic voltage and Duration
The argument of the characteristic voltage, namely the characteristic
phase-angle shift, can be added as a further test. The range of the
argument depends on the system parameters, i.e. the X/R ratio of the
source and the feeders. Field measurements like shown in Figure 5.11
and 5.12 can be used as a reference. The general trend is that a bigger
argument is associated with a lower characteristic magnitude and that
only negative argument values need to be considered. The preferred
characteristic voltages and durations are shown in Table 6.2.
After setting down the test range of characteristic voltage, the phase
voltages of different types of dip are calculated from (6.1), (6.2), (6.3),
which were derived in Chapter 3:
Type A:
(6.1)
Type C &
Type D
Characteristic
magnitude |V|
0 33% 40% 70%
Characteristic phase-
angle shift arg(V)
-- -20
0
-15
0
-10
0
Type A Characteristic
magnitude |V|
0 -- 40% 70%
Characteristic phase-
angle shift arg(V)
-- -- -20
0
-15
0
Duration (in period) 0.5 1 5 10 25 50 X
V
a
V =
V
b
1
2
---V
1
2
--- jV 3 =
V
c
1
2
---V
1
2
--- jV 3 + =
Chapter 6: Equipment Immunity Tests
114
Type C:
(6.2)
Type D:
(6.3)
Note: V, V
a
, V
b
, and V
c
are complex values.
C. PN-factor
Field measurements were shown in Chapter 5, resulting in, among
others, values for the PN-factor of unbalanced dips. The conclusion
from this was that the actual value of the PN-factor is close to unity.
Only when large induction motor loads are connected to the system,
the PN-factor deviates from unity. It is suggested to use a relation
between PN-factor F and characteristic magnitude V as follows:
(6.4)
with f
0
equal to 1, 0.95, and 0.9. The derivation of (6.4) can be found in
Appendix B.
The phase voltages of type C and type D can be calculated by Equation
(6.5) and (6.6) after the PN-factor is introduced.
V
a
1 =
V
b
1
2
---
1
2
--- jV 3 =
V
c
1
2
---
1
2
--- jV 3 + =
V
a
V =
V
b
1
2
---V
1
2
--- j 3 =
V
c
1
2
---V
1
2
--- j 3 + =
F f
0
1 f
0
( )V + =
Chapter 6: Equipment Immunity Tests
115
Type C:
(6.5)
Type D:
(6.6)
D. Zero-sequence
As mentioned before, the zero-sequence component rarely appears at
the equipment terminals. Only for SLGF and 2LGF zero-sequence
voltage occurs at the fault location. This component is blocked by most
transformers; the exception is the star-star connected transformer
grounded on both sides.
The zero-sequence component is normally only present in dips of type
D due to SLGFs. Theoretically, zero-sequence component could also
be present in dips of type C due to 2LGF, but this is rather rare, and
thus will be not considered in the test.
From the sequence network of SLGF shown in Figure 3.3, the
following relation between the zero-sequence voltage V
0
and the
characteristic voltage V is derived:
(6.7)
V
a
F =
V
b
1
2
---F
1
2
--- jV 3 =
V
c
1
2
---F
1
2
--- jV 3 + =
V
a
V =
V
b
1
2
---V
1
2
--- jF 3 =
V
c
1
2
---V
1
2
--- jF 3 + =
V
0
Z
s0
2
--------
1 V
Z
s1
------------- =
Chapter 6: Equipment Immunity Tests
116
In a solidly-grounded system, the positive- and zero-sequence source
impedance can be assumed equal. The zero-sequence voltage can be
chosen as
(6.8)
In case the source impedance at the point-of-common coupling is
dominated by overhead line impedance, the zero-sequence ground
impedance may be twice the positive-sequence value. This leads to the
following relation:
(6.9)
In a high-impedance grounded system, a single-phase fault leads to a
characteristic voltage and a zero-sequence voltage both close to 1 pu.
As the use of grounded equipment is extremely unlikely in high-
impedance grounded systems, this case is not relevant for equipment
test.
The zero-sequence voltage is opposite in direction to the characteristic
voltage. The presence of zero-sequence will even decrease the voltage
in the symmetrical phase of type D, and swells may occur in the other
two phases.
The phase voltages of type D dip can be calculated by Equation (6.10)
after both PN-factor and zero-sequence voltage are introduced:
Type D:
(6.10)
E. Other test items
In case long post-fault dips are possible, it is recommended to test the
equipment against it. The post-fault test can be performed as an
V
0
1
2
--- 1 V ( ) =
V
0
1 V =
V
a
V V
0
+ =
V
b
1
2
---V
1
2
--- jF 3 V
0
+ =
V
c
1
2
---V
1
2
--- jF 3 V
0
+ + =
Chapter 6: Equipment Immunity Tests
117
additional type A dip immediately after the during-fault dip. The depth
and duration should be obtained from a system study.
The point-on-wave of initiation is generally not considered in three-
phase equipment tests, since the three phase voltages drop at different
points on the waveform. In case the point-on-wave is considered, the
phase where the point-on-wave is specied should be indicated. The
point-on-waves of the other two phases have -120
0
and +120
0
difference.
The point-on-wave of recovery on three phases could be different
depending on the coordination of the protection relay. The type of dip
could change during the dip recovery if the recovery instant is different
for three phases. If the immunity of the equipment under test (EUT) is
very dip-type dependent, the test setup should simulate such a
procedure.
Source impedance: If the cause of equipment mal-operation is due to
overcurrent or current unbalance, the source impedance is likely to
affect the immunity. Tripping on overcurrent may occur at voltage
recovery. Tripping on current unbalance may occur for three-phase
ASD.
6.2.2 Test setup
The dip generator of three-phase equipment can again be put into two
categories: the amplier-type and transformer-type.
The amplier-type dip generator should be able to generate three
independent waveforms to supply three phases of the equipment under
test. Three sets of the improved transformer-type dip generator shown
in Figure 6.10 should be used, as phase-angle shifts always accompany
unbalanced dips. For both kinds of dip generator, the three phases
should be coordinated to generate the intended type of dip.
Chapter 6: Equipment Immunity Tests
118
Chapter 7: Conclusions and Future Research
119
Chapter 7 Conclusions and Future Research
7.1 Conclusions
In this thesis, a method for characterizing three-phase voltage dips is
proposed and investigated. Under this characterization, three-phase
voltage dips are classied as three main types, denoted by A, C, and D.
Type A corresponds to balanced dips with an equal change in the three
phases. Each balanced dip is quantied by its duration and a
characteristic voltage. The characteristic voltage is dened as the
positive-sequence voltage for balanced dips. The absolute value of the
characteristic voltage (the characteristic magnitude) corresponds to the
dip magnitude of each phase, while the argument of the characteristic
voltage (the characteristic phase-angle shift) corresponds to the phase-
angle shift of each phase.
Type C and D correspond to unbalanced dips due to unbalanced faults.
If the symmetrical phase is considered, unbalanced dips can be further
classied as six types, which are denoted by C
a
, C
b
, C
c
, D
a
, D
b
, D
c
.
Each unbalanced dip is fully dened by its duration plus three
characteristics:
The characteristic voltage is the main quantier for the severity of the
event. It is dened as the subtraction of positive- and negative-
sequence voltage of type C
a
, which is called a prototype dip. For many
applications the characteristic magnitude (absolute value of
characteristic voltage) may be sufcient to characterize the event.
The so-called PN-factor is a measure for the effect of the system load
on voltages at the equipment terminals during the fault. It is dened as
the summation of positive- and negative-sequence voltage of type C
a
.
Neglecting system load gives a PN-factor equal to one. In many cases
the PN-factor is close to one and can thus be neglected. The PN-factor
needs to be considered for specic applications in systems with a large
amount of induction motor loads.
The zero-sequence voltage due to an earth fault rarely transfers down
to the equipment terminals. Also equipment behaviour is rarely
affected by the zero-sequence voltage. This characteristic is therefore
of minor importance.
Chapter 7: Conclusions and Future Research
120
A big advantage of the proposed classication method is that
characteristic magnitude plus dip duration are sufcient to characterize
a three-phase voltage dip, either balanced or unbalanced.
The two-component symmetrical component method is used for the
theoretical analysis of dip classication. Positive- and negative-
sequence source impedance are assumed to be equal in using this
method. Various faults are analyzed by this method. The dip type and
the characteristics of voltage dips are dened based on the analysis.
Mathematical models of transformers for voltage dip transformation
are dened and studied. It is concluded that the transformers do not
generate new types of dips but could change dip types. Most
importantly, a type C and a type D dip change into each other through a
Delta-Wye transformer. This reects the internal relationship between
single-line-to-ground fault (SLGF) and line-to-line fault(LLF). The
phase-angle shift phenomena associated with voltage dips are also
examined. The phase-angle shifts in a balanced dip are caused by the
X/R ratio difference between the source and the faulted feeder, while
phase-angle shifts in an unbalanced dip are the aggregation effect of
X/R ratio difference and three-phase unbalance.
The propagation of voltage dips requires different analysis tools in
distribution systems and transmission systems. A voltage divider
model can be used for studying both balanced and unbalanced dips in a
radial distribution network. In transmission systems, however, the
situation is more complicated. Digital simulation has to be applied for
a numerical solution. The proposed dip classication method greatly
simplies the study of unbalanced dip propagation in power systems. If
the fault locations are the same, the characteristic voltage of dips from
three-phase fault (3), line-to-line fault(LLF), and two-line-to-ground
fault (2LGF) are the same with the dip propagation; the characteristic
voltage of a dip from SLGF is the same as the characteristic voltage of
a 3 with an extra half zero-sequence impedance (Z
0
/2) connected at
the fault location; the PN-factor of the dip from 2LGF is the same as
the characteristic voltage of a 3 with an extra double zero-sequence
(2Z
0
) at the fault place. Thus, a single-phase scheme is even capable of
studying unbalanced dip propagation in power systems.
Dynamic loads, e.g. induction motor loads, have a strong inuence on
dips characteristics during their propagation. In a system where large
induction motor loads are connected, a non-unity PN-factor of the
unbalanced dip results due to two reasons:
Chapter 7: Conclusions and Future Research
121
Positive- and negative-sequence impedances are different for the
dynamic loads; The consequence of this is that, at any position in the
system, positive- and negative-sequence impedance of the equivalent
source, are not exactly equal. This difference causes a non-unity PN-
factor. The effect of induction motor load is a reduction of the PN-
factor.
The positive-sequence impedance of motor load decreases with
increasing motor slip. As motors slow down during a voltage dip, the
load impedance will decrease with time during the dip. Because of this,
the source voltage of the equivalent source will decrease with time,
leading to a drop in the PN-factor.
Field measurements in both transmission systems and distribution
systems shows that the recorded dips can be classied as the proposed
types. The PN-factors of the measured dips in transmission systems are
very close to unity; the PN-factors of the measured dips in distribution
systems are slightly less than unity. The characteristic voltage of both
balanced and unbalanced dips show similar values. The characteristic
phase-angle shift is close to zero in transmission systems, but might
have a big negative value in distribution systems. The negative phase-
angle shift in distribution systems is often larger for balanced dips than
unbalanced dips. The analysis of a propagating voltage dip from 132
kV down to 33 kV and 11 kV shows increased characteristic
magnitude and decreased PN-factor from higher voltage level to lower
voltage level. The application of the classication method on a power
quality survey shows that this method has advantages in interpreting
voltage dip measurement. The waveform of the recordings can be
reproduced with a reasonable level of accuracy. It is also shown that the
shallow and unbalanced dips constitute the majority of the recordings.
Immunity testing of single-phase equipment against voltage dips is part
of international standards. The dip magnitude and duration are the
main concern for equipment immunity. However, other dip
characteristics, e.g. phase-angle shift and point-on-wave could also
contribute in tripping certain kinds of equipment. Thus, these non-
energy-related characteristics should be also considered in the future
standards. The dip classication method makes the immunity test of
three-phase equipment against voltage dips more systematic. Since
both balanced and unbalanced dips can be mainly characterized by the
characteristic voltage, the method developed for presenting single-
phase immunity tests is easily adapted to three-phase tests by the
separation of different types of dip. In a more detailed test, the
Chapter 7: Conclusions and Future Research
122
additional characteristics, i.e. PN-factor, zero-sequence voltage, might
be added.
7.2 Future Research
Dynamic loads have a strong inuence on dip characteristics. Induction
motor loads have been modelled and investigated. However, the
systems loads are complicated. Other dynamic loads, e.g. rectier
loads, could also affect the dips characteristics and the correctness of
the classication method. The loads effects on characteristic voltage
and PN-factor needs to be further studied.
The phase-angle shift phenomena associated with both balanced and
unbalanced dips are well explained by the fault analysis and the
mathematic dip model. However, the point-on-wave issue, which is
related to the system fault characteristic and protection relay behavior,
still lack information. The studies of the point-on-wave issue should
combine theoretical analysis with eld measurements.
The dip classication method simplies the dip propagation study for
unbalanced dips, where a single-phase scheme is used. The conclusion
could be directly used by stochastic prediction of voltage dips, where
the existing methods only concern balanced dips. The dip classication
method could adapt the existing method to unbalanced situations.
The dip classication method can have several applications for various
power quality standards. The dip classication is quite promising in
presenting voltage dip measurements from power quality surveys. This
characterization can assist in further development of standards for
monitoring voltage dips and for exchange of information between
utilities, customers, and equipment manufacturers. Three-phase
equipment immunity tests against voltage dips is another application of
the dip classication, in which it becomes critical to understand the
magnitude and phase-angle shift relation of the unbalanced voltage
dips. In development of equipment immunity test protocol, the test
range of characteristic voltage, PN-factor, zero-sequence, need to be
further studied with a consideration of statistics from eld
measurements.
Filed measurements have been obtained from various sources. Field
measurements show that the classication method is valid for both
transmission systems and distribution systems. Further statistics of dip
Chapter 7: Conclusions and Future Research
123
characteristics still need to be obtained from eld measurements to
compare with the theoretical study. Several monitors should be
installed at different voltage levels to measure the dip propagation.
This can also help understand the change of dip type through
transformers. The proposed classication method is based on several
assumptions. The limitation of the method should also be assessed
through eld measurements.
Chapter 7: Conclusions and Future Research
124
References
125
References
[1] R. C. Dugan, M.F. McGranaghan, H.W. Beaty, Electrical power systems
quality, New York: McGraw-hill,1996.
[2] M.F. McGranaghan, D.R. Mueller, and M.J. Samotyj, Voltage sags in
industry power systems, IEEE Transactions on Industry Applications,
vol.29, no.2, pp. 397-403, Mar/Apr 1993.
[3] L.E. Conrad (Chairman), IEEE Std 493 Chapter 9, Voltage sag analysis,
Power System Reliability Subcommittee Voltage Sag Working Group,
draft 6 1995.
[4] M.H.J. Bollen, Understanding and Solving Power Quality Problems:
Voltage Sags and Interruptions, New York, IEEE press, 1999.
[5] M.H.J. Bollen, Fast assessment methods for voltage sags in distribution
systems, IEEE Transactions on Industry Applications, vol.32, no.6,
pp.1414-1423, Nov/Dec 1996.
[6] M.H.J. Bollen, Characterization of voltage sags experienced by three-
phase adjustable-speed drives, IEEE Transactions on Power Delivery, vol.
12, no.12, pp.1666-1671, October 1997.
[7] IEEE Project Group 1159.2, Task force on characterization of a power
quality event given an adequately sampled set of digital data points, first
draft 1998.
[8] IEEE Std 1159-1995, IEEE recommended practice for monitoring electric
power quality, November 1995.
[9] CENELEC, Voltage characteristics of electricity supplier by public
distribution systems, European Standard - EN50160, November 1994.
[10] Electromagnetic Compatibility (EMC), part 4. Testing and measurement
protocols. Section 11. Voltage Dips, short interruptions and voltage
variations immunity tests. IEC Document 61000-4-11.
[11] L.E. Conrad, M.H.J. Bollen, Voltage sag coordination for reliable plant
operation, IEEE Transactions on Industry Applications, vol.33, no.6,
pp.1459 - 1464, Nov/Dec, 1997.
[12] P.M. Andersson, Analysis of faulted power systems, New York: IEEE
Press, 1995.
References
126
[13] R. Collins, Waveform characteristics of voltage sags: definition and
algorithm development, Clemson University, South Carolina, March
1998.
[14] L.D. Zhang, M.H.J. Bollen, A method for characterizing unbalanced
voltage dips (sags) with symmetrical components, IEEE Power
Engineering Letters, pp. 50-52, July 1998.
[15] L.D. Zhang, M.H.J. Bollen, Characteristics of voltage dips (sags) in power
systems, accepted by IEEE PES Transactions, in print.
[16] M.H.J Bollen, P. Wang, N. Jenkins, Analysis and consequences of the
phase jump associated with a voltage sag, Power Systems Computation
Conference, Dresden, Germany, August 1996.
[17] Measurement guide for voltage characteristics, UNIPEDE report 23003
ren 9531. UNIPEDE documents can be obtained from UNIPEDE 28, rue
Jacques Ibert, 75858 paris Cedex 17, France.
[18] C.A. Warren, T.A. Short, J.J. Burke, H.Morosini, C.W. Burns, J. Storms,
Power quality at champion paper - the myth and the reality, IEEE
Transactions on Power Delivery, vol.14, no.2, April 1999.
[19] A.V.Zyl, R.Spee, A.Favelike, S.Bhowmik, Voltage sag ride-through for
adjustable-speed drives with active rectifiers, IEEE Transactions on
Industry Applications, vol.34, no.6, pp. 1270-1277, Nov/Dec, 1998.
[20] G.T. Heydt, W.T. Jewell, Pitfalls of electric power quality indices, IEEE
Transactions on Power Delivery, vol.13, no.2, pp. 570-578, April 1998.
[21] G. Yalcinkaya, M.H.J. Bollen, and P.A.Crossley, Characterization of
voltage sags in industrial distribution systems, IEEE Transactions on
Industry Applications, vol.34, pp.682-688, 1998.
[22] IEEE P1346, Recommended practice for evaluating electric power system
compatibility with electronic process equipment, 4th draft, September
1996.
[23] IEEE guide for application of power electronics for power quality
improvement on distribution systems rated 1kV through 38 kV, February
1997.
[24] M.H.J. Bollen, The influence of motor reacceleration on voltage sags,
IEEE Transactions on Industry Applications, vol.31, no.4, Jul/Aug 1995.
References
127
[25] G.Desquilbet, C. Foucher, P. Fauquembergue, Statistical analysis of
voltage dips, 95NR00102, EDF, 1995.
[26] A.E. Turner, E.R. Collins, The performance of ac contactors during
voltage sags, 7th International Conference on Harmonics and Quality of
Power, Las vegas, October 1996.
[27] A. Mansoor, E.R. Collins, R.L. Morgan, Effects of unsymmetrical voltage
sags on adjustable speed drives, 7th International Conference on
Harmonics and Quality of Power, Lars vegas, October 1996.
[28] M.H.J. Bollen, L.D. Zhang, Analysis of voltage tolerance of adjustable-
speed drives for three phase balanced and unbalanced sags, IEEE
Industrial and Commercial Power Systems Technical Conference, Sparks,
Nevada, May 1999.
[29] H. Seljeseth, A. Pleym, Spenningskvalitetsmlinger 1992 til 1996 ( voltage
quality measurements, 1992 to 1996, in Norwegian), report EFI TR A4460
published by EFI, 7034 Trondheim, Norway, March 1997.
[30] D.S. Dorr, M.B. Hughes, T.M. Gruzs, Interpreting recent power quality
surveys to define the electrical environment, IEEE Transactions on
Industry Applications, vol.33, no.6. pp.1480-1487, Nov/Dec 1997.
[31] D.O. Koval, M.B. Hughes, Canadian national power quality survey:
frequency of industrial and commercial voltage sags, IEEE Transactions
on Industry Applications, vol. 33, no.3, pp.622-627, May/June 1997.
[32] D.O. Koval, R.A. Bocancea, K. Yao, M.B. Hughes, Canadian national
power quality survey: frequency and duration of voltage sags and surges at
industrial sites, IEEE Transactions on Industry Applications, vol.34, no.5,
pp.904-910, Sep/Oct 1998.
[33] A. Burden, A. MacGregor, power quality initiatives at scottish power,
Transmission & Distribution World, February 1998.
[34] SEMI F42-0699, Test method for voltage sag susceptibility of
semiconductor processing equipment, SEMI online January 1999,
http://dom.semi.org.
[35] P. Wang, The use of FACTS devices to mitigate voltage sags, PH.D thesis,
UMIST, UK, July 1997.
[36] G.Yalcinkaya, The influence of induction motors on voltage sags due to
short circuits, PH.D thesis, UMIST, UK, October 1997.
References
128
[37] M.H.J. Bollen, P.Wang, N.Jenkins, Magnitude and phase-angle jump of
voltage sags -- a theoretical analysis, submitted to IEEE Transactions on
Power Delivery.
[38] M.H.J. Bollen, T. Tayjasajant, G.Yalcinkaya, Assessment of the number of
voltage sags experienced by a large industrial customer, IEEE Transaction
on Industry Applications, Nov/Dec 1997.
[39] M.H.J. Bollen, A.Mansoor, E.R. Collins, Characteristics of voltage sags
experienced by single-phase and three-phase Equipment, PQA97 Europe,
Stockholm Sweden, June 1997.
[40] P.Rioual, J.L. Javerzac, F.de Chateauvieux, Characterization and
recognition of HV and EHV voltage dip signatures, PQA97 Europe,
Stockholm Sweden, June 1997.
[41] A.David, E. Lajoie-Mazenc, C.Sol, Ride-through capability of AC
adjustable speed drives in regards to voltage dips on the distribution
network, EPE93, the European Power Electronics Association, Brighton,
1993.
[42] V.E. Wagner, A.A. Andreshak, J.P. Staniak, Power Quality and Factory
Automation, IEEE Transactions on Industry Applications, vol.26, no.4, pp.
620-626, Jul/Aug 1990.
[43] D.O. Koval, J.C. Chang, J. Leonard, Rural power quality, IEEE
Transactions on Industry Applications, vol.28, no.4, pp.761-765, Jul/Aug
1992.
[44] M.R. Qader, M.H.J. Bollen, Stochastic voltage sag prediction technique,
31st University Power Engineering Conference, Heraklion, Greece,
pp.441-444, 1996.
[45] M. Morgan, M.Johans, Voltage sag mitigation through ride-through
coordination, IEEE annual Textile, fiber, film conference, Greenville, SC,
USA, pp.1-5, 1994.
[46] R.A. Epperly, F.L. Hoadley, R.W. Piefer, Considerations when applying
ASDs in continuous processes, IEEE Transactions on Industry
Application, vol.33, no.2, pp. 389-396, Mar/Apr 1997.
[47] C.J. Melhorn, A.B. Braz, P. Hofmann, R.J. Mauro, An evaluation of
energy storage techniques for improving ride-through capability for
sensitive customers on underground networks, IEEE Transactions on
Industry Applications, vol.33, no.4, pp. 1083-1095, Jul/Aug 1997.
References
129
[48] C.J. Melhorn, T.D.Davis, G.E. Beam, Voltage sags: their impact on the
utility and industrial customers, IEEE Transactions on Industry
Applications, vol. 34, no.3, pp. 549-558, May/Jun 1998.
[49] J.L. Zabala, S. Ruiz, G. Vargas, F. Soto, X. Prat, R. Barba,
Characterisation of voltage dips in electrical networks and their Impact on
customer installations, Cigre 21, rue dArtois, F-75008 Paris.
[50] P.P. Khera, K.C. Dickey, Analysis and mitigation of voltage disturbances
at an industrial customers corporate campus, IEEE Transactions on
Industry Application, vol.34, no.5, pp. 893-896, Sep/Oct 1998.
[51] D.L. Brooks, R.C. Dugan, M.Waclawiak, A. Sundaram, Indices for
assessing utility distribution system RMS variation performance, IEEE
Transactions on Power Delivery, vol.13, no.1, pp. 254-259, January 1998.
[52] A. Kara, P. Dahler, D. Amhof, H. Gruning, Power supply quality
improvement with a dynamic voltage restorer (DVR), APEC98, IEEE
Thirteenth Annual Applied Power Electronics Conference and Exposition,
vol.2, pp.986-993 New York, 1998.
[53] K. Hadda, G. Joos, A fast algorithm for voltage unbalance compensation
and regulation in faulted distribution systems, APEC98, IEEE Thirteenth
Annual Applied Power Electronics Conference and Exposition, vol.2,
pp.986-993, New York, 1998.
[54] Y.Kai, D. Koval, X. Wilsun, J. Salmon, An investigation of voltage sags
by phasor methodology, IEEE Canadian Conference on Electrical and
Computer Engineering, vol.2, pp.689-692, New York, 1998.
[55] A.E. Turner, E.R. Collins, The performance of AC contactors during
voltage sags, 7th International Conference on Harmonics and Quality of
Power, Las Vegas, October 1996.
[56] E.G. Strangas, V.E. Wagner, T.D. Unruh, Variable speed drives evaluation
test, IEEE Industry Applications Magazine, pp.53-57, Jan/Feb 1998.
[57] J. Holtz, W. Lotzkat, Controlled AC drives with ride-through capability at
power interruption, IEEE Transactions on Industry Applications, vol. 30,
no.5, pp.1275-1283, Sep/Oct 1994.
[58] Manitoba HVDC Research Centre, PSCAD/EMTDC Manual, Winnipeg,
Canada, 1992.
[59] J. Hilgers, Design and realization of a voltage sag generator, M.Sc thesis
References
130
No. 34 E, Department of Electric Power Engineering, Chalmers University
of Technology, 1999.
Appendix A:Determination of Zero-sequence Source Impedance
131
Appendix A Determination of Zero-sequence
Source Impedance
The power networks zero-sequence source impedance is not always
known. However, it can be determined from historical voltage dip
recordings, together with the positive-sequence impedance of the
source and the feeders. Two cases will be considered below.
1. Both the positive-sequence source impedance at PCC and the feeder
positive-sequence impedance are known. Only a dip from an SLGF at
the feeder is needed.
For a dip due to an SLGF at a distance, L from the PCC, the
characteristic voltage V
SLGF
is:
(A.1)
where L is the distance from the fault to the PCC, Z
s1
is the positive
source impedance at the PCC, Z
f1
is the feeders positive-sequence
impedance per kilometer, Z
0
is the zero-sequence impedance at fault
location.
Knowing the characteristic voltage V
SLGF
and the fault position L, we
can calculate Z
0
from (A.1),
(A.2)
2. Only the feeders positive sequence impedance is known. A dip from
a SLGF and a dip from a 3F , LLF or 2LGF are needed.
For a dip due to a 3F, LLF or 2LGF at a distance L from the PCC, the
characteristic voltage V
3
is
(A.3)
V
SLGF
LZ
f 1
Z
0
2
------ +
Z
s1
LZ
f 1
Z
0
2
------ + +
----------------------------------------- =
Z
0
2
V
SLGF
Z
s1
1 V
SLGF

----------------------- LZ
f 1

( ,
j \
=
V
3
LZ
f 1
Z
s1
LZ
f 1
+
--------------------------- =
Appendix A:Determination of Zero-sequence Source Impedance
132
The positive-sequence source impedance can be calculated from (A.3),
(A.4)
and substituted into (A.2) to calculate Z
0
/2.
In transmission systems, the voltage divider model can not be used. To
calculate the zero-sequence impedance Z
0
, the positive-sequence
source impedance at the fault point has to be known, and a dip from
SLGF has to be measured at the fault point. From (A.2) with L = 0, we
get
(A.5)
Basically, zero-sequence impedance can be calculated by (A.5) in any
systems. But the problem is that positive-sequence source impedance
at the fault point and a measured dip at the fault point are also not
available. More practical ways to obtain zero-sequence source
impedance in non-radial systems are still needed.
As an example, we use the dip data in Table 4.4 to calculate the zero-
sequence source impedance. Faults at a feeder away from the 11 kV
bus are chosen for calculations.
The rst step is to calculate the positive-sequence source impedance
from a 3F, LLF or 2LGF by (A.4). Given Z
f1
= 0.1174+j0.3146
ohm/km, L = 2 km, and V = 0.55-8
0
, we get Z
s1
= 0.03+j0.5628 .
The second step is to calculate the half zero-sequence (Z
0
/2) from a dip
caused by SLGF by (A.3). Given Z
f1
= 0.1174+j0.3146 /km, L = 2
km, and V = 0.69-5
0
, we get Z
0
= 0.2524 + j1.0532 .
Z
s1
1 V
3

V
3
--------------------LZ
f 1
=
Z
0
2
V
SLGF
Z
s1
1 V
SLGF

----------------------- =
Appendix B:PN-factor and Characteristic Voltage
133
Appendix B: PN-factor and Characteristic Voltage
In a system where positive-sequence source impedance (Z
s1
) and
negative-sequence impedance (Z
s2
) are not equal, the PN-factor (F)
will be not equal to unity. Besides, the PN-factor (F) also related the
characteristic voltage (V). We will study the dips from SLGF and LLF
separately to give their relationships.
The standard voltage divider model as shown in Figure B.1 gets the
form shown in Figure B.2 for SLGF and in Figure B.3 for LLF.
B.1 Single-line-to-ground Fault (SLGF)
Figure B.1 Voltage divider model
PCC
F
Load
Z
s
Z
f
+
-
V
F
V
2
V
0
V
1
N
1
N
2
N
0
Z
s1
Z
s2
Z
s0
Z
f1
+ Z
f2
+ Z
f0
Figure B.2 Sequence-network connection of SLGF
P
1
P
2
P
0
+
-
+ - + -
I
a1
= I
a2
= I
a0
Appendix B:PN-factor and Characteristic Voltage
134
From Figure B.2, the three sequence voltages at the PCC are obtained:
(B.1)
According to the denition of characteristic voltage (V) and PN-factor
(F) for SLGF in Chapter 3,
(B.2)
(B.3)
For Z
s1
= Z
s2
we obtain F = V
F.
It follows from (B.2) and (B.3)
(B.4)
Note that this value is no longer dependent on the source voltage V
F
.
All quantities in F/V are expected to be time-independent as well.
It also follows from (B.2)
(B.5)
V
1
V
F
Z
s2
Z
s0
+ ( ) Z
f 1
Z
f 2
Z
f 0
+ + ( ) +
Z
s1
Z
s2
Z
s0
+ + ( ) Z
f 1
Z
f 2
Z
f 0
+ + ( ) +
---------------------------------------------------------------------------------------------- =
V
2
V
F
Z
s2
Z
s1
Z
s2
Z
s0
+ + ( ) Z
f 1
Z
f 2
Z
f 0
+ + ( ) +
---------------------------------------------------------------------------------------------- =
V
0
V
F
Z
s0
Z
s1
Z
s2
Z
s0
+ + ( ) Z
f 1
Z
f 2
Z
f 0
+ + ( ) +
---------------------------------------------------------------------------------------------- =

V V
1
V
2
+ =
V
F
Z
s0
Z
f 1
Z
f 2
Z
f 0
+ + ( ) +
Z
s1
Z
s2
Z
s0
+ + ( ) Z
f 1
Z
f 2
Z
f 0
+ + ( ) +
---------------------------------------------------------------------------------------------- =
F V
1
V
2
=
V
F
Z
s0
2Z
s2
+ ( ) Z
f 1
Z
f 2
Z
f 0
+ + ( ) +
Z
s1
Z
s2
Z
s0
+ + ( ) Z
f 1
Z
f 2
Z
f 0
+ + ( ) +
---------------------------------------------------------------------------------------------- =
F
V
---- 1
2Z
s2
Z
s0
Z
f 1
Z
f 2
Z
f 0
+ + ( ) +
----------------------------------------------------------- + =
Z
f 1
Z
f 2
Z
f 0
+ + ( )
V Z
s0
Z
s1
Z
s0
+ + ( ) V
F
Z
s0

V
F
V ( )
----------------------------------------------------------------------- =
Appendix B:PN-factor and Characteristic Voltage
135
Substitute (B.5) into (B.4),
(B.6)
where
(B.7)
B.2 Line-to-line fault (LLF)
From Figure B.3, the three sequence voltages at PCC are obtained:
(B.8)
F
2Z
s2
Z
s1
Z
s2
+
-----------------------V
F
1
2Z
s2
Z
s1
Z
s2
+
-----------------------
( ,
, (
j \
V + =
f
0
V
F
1 f
0
( )V + ( ) =
f
0
2Z
s2
Z
s1
Z
s2
+
----------------------- =
-
+
P
1
I
a0
= 0
N
0
+
-
-
+
Z
s1
N
1
N
2
Z
s2
-
+
V
a1
V
a2
Z
f1
+ Z
f2
I
a1
I
a2
V
F Z
s1
P
1
P
2
V
a0
Figure B.3 Sequence-network connection of LLF
V
a1
V
F
Z
s2
Z
f 1
Z
f 2
+ ( ) +
Z
s1
Z
s2
+ ( ) Z
f 1
Z
f 2
+ ( ) +
--------------------------------------------------------------- =
V
a2
V
F
Z
s2
Z
s1
Z
s2
+ ( ) Z
f 1
Z
f 2
+ ( ) +
--------------------------------------------------------------- =
V
a0
0 =

Appendix B:PN-factor and Characteristic Voltage


136
According to the denition of characteristic voltage (V) and PN-factor
(F) for LLF in Chapter 3,
(B.9)
(B.10)
For Z
s1
= Z
s2
we obtain F = V
F
It follows from (B.9) and (B.10)
(B.11)
It also follows from (B.9)
(B.12)
Substitute (B.12) into (B.11),
(B.13)
where
(B.14)
V V
1
V
2
=
V
F
Z
f 1
Z
f 2
+ ( )
Z
s1
Z
s2
+ ( ) Z
f 1
Z
f 2
+ ( ) +
--------------------------------------------------------------- =
F V
1
V
2
+ =
V
F
2Z
s2
Z
f 1
Z
f 2
+ ( ) +
Z
s1
Z
s2
+ ( ) Z
f 1
Z
f 2
+ ( ) +
--------------------------------------------------------------- =
F
V
---- 1
2Z
s2
Z
f 1
Z
f 2
+
----------------------- + =
Z
f 1
Z
f 2
+ ( )
V Z
s1
Z
s2
+ ( )
V
F
V ( )
--------------------------------- =
F
2Z
s2
Z
s1
Z
s2
+
-----------------------V
F
1
2Z
s2
Z
s1
Z
s2
+
-----------------------
( ,
, (
j \
V + =
f
0
V
F
1 f
0
( )V + ( ) =
f
0
2Z
s2
Z
s1
Z
s2
+
----------------------- =
Appendix B:PN-factor and Characteristic Voltage
137
The characteristic voltage and PN-factor follow the same relationship
for both dips from SLGF and LLF. The constant f
0
, which is
independent of the fault reects the system parameters inuence on
the dip characteristics.
Appendix B:PN-factor and Characteristic Voltage
138

Potrebbero piacerti anche