You are on page 1of 188

Electric Power Systems

Electric Power Transmission and Distribution


227-0122-00L
Goran Andersson
EEH Power Systems Laboratory
ETH Z
urich
September 2009

Contents
Preface

vii

Recommended Literature

viii

1 Introduction
1.1 Electrical Energy . . . . . . . . . . . . . . . . . . . . . . . . .
1.2 Production of Electrical Energy . . . . . . . . . . . . . . . . .
1.2.1 Power Stations . . . . . . . . . . . . . . . . . . . . . .
1.2.2 Electric Power Production in Different Countries . . .
1.3 Transmission and Distribution of Electrical Energy . . . . . .
1.3.1 Voltage Levels and Net Types . . . . . . . . . . . . . .
1.3.2 International Interconnected Grids and Synchronous
Zones . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.3.3 Grid Structures . . . . . . . . . . . . . . . . . . . . . .
1.3.4 Power Substations . . . . . . . . . . . . . . . . . . . .
1.3.5 Grid Operation and Supervision . . . . . . . . . . . .
1.4 Consumption of Electrical Energy . . . . . . . . . . . . . . .

1
1
1
2
6
7
9
10
13
14
16
17

2 Power in Alternating Current Systems


21
2.1 Power in Single Phase Networks . . . . . . . . . . . . . . . . . 21
2.2 Conservation of Apparent Power . . . . . . . . . . . . . . . . 27
2.3 Power in Symmetric Three-Phase Networks . . . . . . . . . . 29
2.3.1 Symmetric Three-Phase Systems . . . . . . . . . . . . 29
2.3.2 Power in Symmetric Three-phase Systems . . . . . . . 32
2.3.3 Wye-Delta-Transformation . . . . . . . . . . . . . . . 34
2.3.4 Single Phase Calculation of Symmetric Three-phase
Systems . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3 Transformers
3.1 Single Phase Transformer . . . . . . .
3.1.1 Coupled Windings . . . . . . .
3.1.2 Ideal Transformer Model . . . .
3.1.3 Realistic Transformation Model
iii

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

39
39
39
40
43

iv

Contents
3.2

Three-Phase Transformer . . . . . . . . . . . . . . . . . . . .
3.2.1 Design of Three-Phase Transformers . . . . . . . . . .
3.2.2 Three-Phase Transformer Model . . . . . . . . . . . .

4 Per
4.1
4.2
4.3

Unit System
Reason for the Calculation with Per Unit Values . . . .
Example of Use of P.U. Values . . . . . . . . . . . . . .
How to Choose Base Voltages for Interconnected Circuit
ements . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.4 Conversion between Different p.u. Systems . . . . . . . .

47
48
49

53
. . . 53
. . . 54
El. . . 56
. . . 58

5 Lines
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.1 Characteristics of Power Lines . . . . . . . . . . . . .
5.1.2 Reasons for High Transmission Voltages . . . . . . . .
5.2 Line Parameters . . . . . . . . . . . . . . . . . . . . . . . . .
5.2.1 Power Line Conductors . . . . . . . . . . . . . . . . .
5.2.2 Inductance of a Power Line . . . . . . . . . . . . . . .
5.2.3 Capacitance of a Power Line . . . . . . . . . . . . . .
5.2.4 Ohmic Losses . . . . . . . . . . . . . . . . . . . . . . .
5.2.5 Conductor Parameters for Cables and Overhead Lines
5.3 Electromagnetic Fields of Overhead Lines . . . . . . . . . . .
5.3.1 Electric Field . . . . . . . . . . . . . . . . . . . . . . .
5.3.2 Magnetic Field . . . . . . . . . . . . . . . . . . . . . .
5.4 Line Model and Solution of the Wave Equation . . . . . . . .
5.4.1 Equivalent Circuit Diagram of a Line Element . . . .
5.4.2 Telegraph Equation . . . . . . . . . . . . . . . . . . .
5.4.3 Wave Equation . . . . . . . . . . . . . . . . . . . . . .
5.4.4 Interpretation of the Wave Propagation . . . . . . . .
5.4.5 Inclusion of Boundary Conditions . . . . . . . . . . . .
5.4.6 Wave Equation of the Lossless Line . . . . . . . . . . .
5.5 Line Models . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.5.1 General Two-Port Model of a Line . . . . . . . . . . .
5.5.2 Lossless Line . . . . . . . . . . . . . . . . . . . . . . .
5.5.3 Additional Simplifications . . . . . . . . . . . . . . . .
5.5.4 Comparison of Different Line Models . . . . . . . . . .

61
61
61
62
63
64
65
75
78
80
81
81
84
88
89
91
92
94
95
97
99
99
101
101
102

6 Power Transmission Fundamentals


6.1 Decoupled quantities . . . . . . . . . . . . . . . . . . . .
6.2 Surge impedance loading . . . . . . . . . . . . . . . . . .
6.2.1 Surge impedance loading of a lossless power line
6.2.2 Surge impedance loading of a lossy power line . .
6.2.3 Typical values for overhead lines and cables . . .
6.3 Open circuit and short circuit . . . . . . . . . . . . . . .

105
105
107
107
109
110
110

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

Contents
6.3.1 Open circuit . . . . . . . . . . . . .
6.3.2 Short circuit . . . . . . . . . . . . .
6.4 Reactive power demand of a line . . . . . .
6.5 Voltage drop along a line . . . . . . . . . .
6.6 Efficiency of high voltage transmission lines
6.7 P -U -Diagram . . . . . . . . . . . . . . . . .
6.8 P --Diagram . . . . . . . . . . . . . . . . .
6.9 Load diagram and load limits . . . . . . . .
6.10 Power cables . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

7 Symmetrical components in three-phase systems


7.1 Unbalanced operating conditions . . . . . . . . . . . . .
7.2 Symmetrical components . . . . . . . . . . . . . . . . .
7.3 Powers in symmetrical component system . . . . . . . .
7.4 120 equivalent circuit . . . . . . . . . . . . . . . . . . . .
7.4.1 The basic power system . . . . . . . . . . . . . .
7.4.2 Neutral point grounding . . . . . . . . . . . . . .
7.4.3 Wye connected load . . . . . . . . . . . . . . . .
7.4.4 Transformers . . . . . . . . . . . . . . . . . . . .
7.5 Fault analysis using symmetrical components . . . . . .
7.5.1 Single-phase to ground fault . . . . . . . . . . . .
7.5.2 Two-phase short-circuit without ground contact
7.5.3 Three-phase short-circuit . . . . . . . . . . . . .
7.6 Resonant Grounding . . . . . . . . . . . . . . . . . . . .
A Zero sequence impedance of transformers

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

110
112
113
116
118
119
121
123
125

.
.
.
.
.
.
.
.
.
.
.
.
.

129
129
130
140
142
142
144
146
146
149
149
155
158
160
169

B Zero sequence impedance of overhead lines


173
B.1 Impedance of a single line with return current through ground 173
B.2 Zero sequence impedance of three-phase overhead lines . . . . 175
Bibliography

179

vi

Contents

Preface
The part Electric Power Transmission and Distribution of the lecture Electric Power Systems in the fifth semester of the Department Information
Technology and Electrical Engineering at ETH Z
urich is concerned primarily with the power transmission over lines or cables in electric power systems. The aim is to describe the transportation of the electric power from
the generator till the final consumers by means of models and equations in
symmetric as well as in asymmetric operating states.
In contrast with the second part High Voltage Technology, where single
components of the high voltage grid (e.g. circuit breakers) are considered, we
are here more interested in the interactions between the devices in the grid,
and not primarily in the specific design of the individual grid components. As
an introduction, systems consisting of single lines from a source (generator)
to a load are considered. Based on that, the lecture Modelling and Analysis of
Electric Power Systems addresses the electric energy transport in stationary
as well as in transient state of a meshed system1 . In the lecture Optimization
of Liberalized Electric Power Systems in the eighth semester, the basics of
this lecture are again partly reused.
Because of the importance of the transformers in the electric power transmission the basics of transformers are reviewed at first in this lecture. Following a two-port model of a transformer is derived. Then we address the
basic elements of this lecture, the power line and power cable.
After the introduction we derive a mathematical two pole model (Telegraph equation of Maxwell) for power lines. At the solution of the Telegraph
equation we restrict ourselves to stationary, i.e. sinusoidal quantities and
thus obtain the equations, which describe the behaviour of current and voltage along a high voltage line. This part is closed with the interpretation of
the line equations and with the analysis of the interaction between active
and reactive power transport over a high voltage line.
The second part gives an introduction to the calculation of short circuit
currents in three-phase high voltage lines. Thereby, we will use a transformation which enables a simplified description of asymmetric three-phase
systems (positive, negative and zero sequence components).

A net is considered as meshed if not only one but multiple lines exist between some (or
also all) nodes which serve for the energy transport. With the intermeshing a redundancy
of the transmission paths is achieved, which has a significant importance for the security of
supply. This redundancy implies that the supply of the energy consumer is still warranted
for an outage of a line/power station.

vii

viii

Preface

Special thanks to the assistants of the ETH Z


urich Power Systems Laboratory, who were very helpful and skilful in preparing this compendium.

Z
urich, September 2009

G
oran Andersson

Recommended Literature
For further studies and clarifications we recommend the following books
available at the ETH library for students of ETH Z
urich:
[1] Bergen, A. R. ; Vittal, V.: Power Systems Analysis. 2nd edition.
Prentice-Hall, 2000.
[2] Oeding, D. ; Oswald, B. R.: Elektrische Kraftwerke und Netze.
6. Auflage. Springer, 2004.
[3] Crastan, V.: Elektrische Energieversorgung 1. Band 1. Springer, 2000.

1
Introduction
This chapter presents a general overview of electric power systems. The fundamental features and functions of electric power systems and the chain
production-transmission-distribution-consumption are discussed.

1.1

Electrical Energy

Electrical energy is produced or generated by conversion of primary energy


sources in electric power generators. Electrical energy is often called a final
energy, as it is used by the end-consumers for different purposes. In comparison with other final energies as for example gas, oil or coal, electrical
energy is characterized by
that it is easy to measure and control
that it could be used for a variety of purposes
that it can be converted efficiently from different primary sources
it has high thermodynamic quality, i.e. it can be converted to mechanical or thermal energy with a high efficiency
that it can be transmitted and distributed with reasonably low losses
However, electrical energy has also disadvantages. The most important
disadvantage is probably the danger of high voltages and currents for human
beings. Another one is that it is difficult to store electrical energy in an efficient way. Therefore, the production of electric power needs to be controlled
at each instant to match the consumption. Nevertheless, electrical energy
nowadays is one of the most important energy forms and constitutes the
basis of the modern society.
Figure 1.1 shows the worldwide overall energy consumption in various
energy forms. Oil and gas are still the energy forms most frequently utilized,
followed by electricity with about 16% of the overall consumption.

1.2

Production of Electrical Energy

In general we know that energy can neither be produced nor consumed.


Nevertheless one is using the expression of producing electrical energy but is
actually meaning the conversion of a primary energy into electrical energy.
This conversion takes place in power stations.
1

1. Introduction

Figure 1.1. Worldwide overall energy consumption of different energy


forms in 2005; Other includes geothermal, solar, wind, etc. [4].

1.2.1

Power Stations

Electricity is produced from diverse primary energy sources. Traditionally,


one is utilizing mostly the potential and kinetic energy of water by running it
through a turbine or utilizing the thermal energy of steam in steam turbines.
For steam generation, one usually uses fossil (coal, oil) or nuclear (enriched
uranium) fuels. A further fossil energy carrier is gas which is converted to
mechanical energy in gas turbines.
Due to worldwide efforts to reduce the greenhouse gas emissions and the
liberalization of the energy markets a new trend is looming since a couple
of years, namely the upcoming decentralized, distributed generation or embedded generation. Small power stations, using renewable1 primary energies
are evolving more and more and are directly connected to the consumer instead of to the transmission grid. By this, new technical, economical, and
regulatory challenges regarding the energy supply arise.
Since several years more power stations with wind and solar energy as
primary sources have been installed. Besides that, conventional primary energy sources are more frequently used in other (mostly more efficient) forms,
as for example fossil fuels in fuel cells or micro turbines. Also, it is sometimes
possible to operate existing technologies with novel fuels. An example is the
usage of biogas in internal combustion engines, which deliver heat as well
as electrical energy. In this context, these machines are often referred to as
Combined Heat and Power (CHP).
Various factors influence the location of power stations. Coal power stations, for example, are built close to large coal mines or nearby ports. Normally, nuclear power stations are not built in the surroundings of large agglomerations. In addition the location needs to fulfill certain requirements as
for example security in case of earth quakes and flooding. Generally, thermal
power stations need cooling water, thus they are often built at large rivers.
1

The term renewable is somewhat misleading, as truly renewable sources do not exist
strictly speaking.

1.2. Production of Electrical Energy

Wind power stations can only be operated efficiently at locations where the
wind conditions give enough so-called full load hours per year, e.g., in coastal
regions.
Hydro Power Stations
Hydro power stations use the kinetic and potential energy of water in rivers
and water supply dams. Depending on the head and the flow rate, the exploitation of either the kinetic or potential energy of the water dominates.
One distinguishes between high, middle or low pressure hydro power stations.
High pressure hydro power stations have very high heads in the range
between 500 and 2000 m. The potential energy of the stored water can be
disposed in hydro reservoirs and processed at peak load times. Accordingly,
these power stations are referred to as storage power stations. For energy
conversion so-called Pelton turbines are employed. At low load times the
power produced in other power stations can be used to fill the reservoirs.
These kind of power stations are referred to as pumped storage hydro stations.
The generators are then used as prime movers for the pumps. High pressure
hydro power stations are characterized by their fast response times. They
are able to adapt their power output within short time (minutes) as well as
to strong fluctuations.
One refers to middle pressure hydro power stations in the case of middle
heads until about 50 m and middle flow rates. Potential and kinetic energy
are used equally. Mostly, Francis turbines are employed.
Low pressure hydro power stations are also referred to as run of river
power stations. The heads are relatively low (a few meters), the flow rate
is instead correspondingly larger. Here Kaplan turbines are well suited as
prime movers for the generators. As the water level and the flow rate of
rivers are more or less constant for a short period of time (several hours
until a day), these power stations are predominantly used to cover the base
load. Low pressure hydro power stations can be used for power control too,
however with higher time constants and a smaller power range than the high
pressure power stations.
As a rule of thumb, the following relation holds for the generated active
power in a water power station
P 8QH

(1.1)

where Q is the flow rate in m3 /s, H the head in m, and P the power output of
the turbine in kW. In this equation the efficiency coefficients of the hydraulic
system h , the turbine t and the generators g are incorporated in the
overall efficiency coefficient:
= h t g 0.82

(1.2)

1. Introduction

where
h 0.90 , t 0.93 und g 0.98
The factor 8 in (1.1) results by multiplication of the overall efficiency coefficient by the earths gravitation constant, 9.81 m/s2 .
As seen, the power output of a hydro power station is linearly dependent
of the flow rate and the head.
In Europe and northern America the rivers carry considerably less water
at winter times or after long dry periods than at summer times or times
with frequent rain. In mountain regions melting water plays an important
role in spring. The power producing capacity of a hydro power station is
strongly depending on meteorological conditions. As an example, in winter
in continental Europe often only a quarter of the installed power rating can
be produced.
Thermal Power Stations
In thermal power stations hot combustion gases or steam are processed in
gas or steam turbines. Typical efficiency coefficients of conventional steam
power stations lie below 40 %. By utilizing the excess heat from the turbines
for heating the overall efficiency can be increased significantly . Basically,
one distinguishes between gas turbine and steam turbine power stations:
gas turbine power stations: Here, gas turbines are employed, in principal corresponding to air plane turbines.
steam turbine power stations: Here, steam turbines are employed. Multilevel turbines (high, middle or low pressure component) are often
used. Depending on the way of steam generation one distinguishes
between:
fossil fired power stations (fossil fuel)
nuclear power plants
Combined Cycle Gas Turbine (CCGT) where both a gas and steam
turbines are used
Thermal power stations are difficult to control compared with hydro power
stations. Regarding the thermal mechanic equipment, dedicated maximal
temperature gradients need to be maintained. This leads to high time constants regarding the changes of output power. Furthermore, the operation
of a thermal power station is economically reasonable mostly only in certain
operation stages. The operation as control power plant is mostly uneconomical. These reasons lead to the fact that steam power stations are employed
mostly for base load coverage. Gas turbines are able to follow varying power
demands within minutes and are employed to cover peak loads as well.

1.2. Production of Electrical Energy

Wind Power
Wind mills use the kinetic energy of wind as energy source. Modern constructions having rotor diameters of 120 m produce power in the range of 5
MW. Mainly in windy coastal regions and vast plains wind energy can be
employed very efficiently and economically.
The produced power of a wind converter P rises with the third power of
the wind speed v:
P v3
(1.3)

Due to natural fluctuations in the wind speed the delivered power varies
quite strongly with wind speed as seen in equation (1.3). Therefore, the
usage of wind energy requires normally fast responding power stations for
the compensation of these power variations.
Solar Power
Solar radiation (sunlight) is employable directly or indirectly for the production of electrical energy. The two dominant technologies are
photovoltaic (direct conversion of sunlight into direct current) and
solar thermic power stations (sunlight for steam production).
The produced power fluctuates mainly between day and night, but as well
between seasons. Besides that, the weather conditions and the cloudiness
influence the produced power.
Fuel Cells
Fuel cells are systems which produce thermal and electrical energy by direct
chemical conversion of fuels. Water is produced as a waste product. Possible
fuels are (bio)gas, kerosene products, hydrogen and alcohol.
Depending on the usage, different technologies are applied. The two most
important types are
Proton Exchange Membrane (PEM) Fuel Cells: Operating temperature 50 until 80 C, efficiency ca. 50 %, typical applications are in the
transportation sector.
Solid Oxide Fuel Cells (SOFC): Operating temperature 600 until 1000 C,
efficiency ca. 70 %, mostly stationary applications.
The maximal nominal power of fuel cell power stations are nowadays in
the range of several MW.
Fuel cells are employable as storage too by using the delivered electrical
energy for the production of hydrogen. The hydrogen can be stored and
be converted back to electrical energy. This technology is referred to as
reversible fuel cell.

1. Introduction

Geothermal Power
Geothermic power stations use the heat from the inner of the earth as primary energy. For this, bore holes to a depth of 5000 m are required. A temperature gradient of about 5 C per 100 m yields a temperature difference
of about 200 C. By the so-called injection bore holes cold water is pressed
in the depth and the heated water rises in the production bore holes. This
heated water is directed to a heat exchanger. Out of the secondary circuit of
the heat exchanger thermal and mechanical/electrical energy is extracted.
More than 99 % of the earth mass has a temperature of more than
1000 C. Correspondingly large or almost inexhaustible is the available potential. Nowadays, about 200 geothermal power stations are in operation.
Currently, a geothermic project named Deep Heat Mining Basel is running in Basel. According to the plans 3 MW of electrical energy and 20 MW
of thermal energy (district heat) will be produced by the five bore holes
(three with a depth of 5000 m in each case, two of 2700 m depth).
An advantage of geothermal energy compared to other alternative energy forms is the (almost) constant availability of the primary energy source.
Thereby the produced power can be adjusted to the needs of the loads.
Operation of Power Stations
As mentioned and discussed above, not only the consumption but also the
production of electrical power varies depending on weather, time of the day,
and season. The difference between the stochastic power producers, i.e. wind,
solar and to some extent hydro power, needs to be balanced by power stations
which can be controlled (dispatched) to the actual total load demand. Figure
1.2 shows the usage of different power station types for energy production
on a winter day in Switzerland. Systems with other mixes of power plants
will exhibit other production patterns.

1.2.2

Electric Power Production in Different Countries

The geographical location of a country and its topographical properties


highly influence the usage of primary energies for the production of electrical energy. Figure 1.3 compares the structures of energy production in
different countries. A few examples will be briefly discussed in the following.
Norway, for example, covers its electrical energy demand almost completely
with hydro power as a lot of hydro resources exist by geographical and meteorological reasons. Another extreme is France, which covers around three
quarters of its electrical energy demand by nuclear power. In 2007 Denmark,
not depicted in the figure, produced around 19 % of its electrical energy demand by wind power [6]. Since Denmark is located geographically close to
Sweden and Norway, both with significant hydro power resources, it is possible for Denmark to export its excessive power produced by wind power

1.3. Transmission and Distribution of Electrical Energy


12

Power in GW

10

8
Hydro Power With Storage
6

Run of River Hydro Power Plants

Nuclear Power Plants

Thermal Power Plants


12
18
Hour of the Day

24 h

Figure 1.2. Electric power production in Switzerland a typical winter day.

stations during high wind condition. This is possible since Norway and Sweden can easily regulate their hydro power plants as a response to the wind
power in Denmark. The energy stored in the large hydro reservoirs can be
used when best requested by the system. This so-called balancing power
will become more important in the future due to the increasing number
of stochastic, also called non-dispatchable, power sources and their limited
predictable production profiles.
In Switzerland more than 50 % of the overall electricity demand is covered by hydro power, which means that the existing hydro power potential
already is exploited relatively well. In a study published by the Swiss Federal
Office of Energy (BfE = Bundesamt f
ur Energie)) the possible extension of
hydro power in Switzerland is estimated to an additional 16 % until 2050 [7].
Next to hydro power 40 % of the electricity is produced by nuclear power
plants.

1.3

Transmission and Distribution of Electrical Energy

Due to economical, environmental, technological reasons and reasons of


availability it is not feasible to cover the overall electric power demand of
large load centers such as cities, agglomerations and industries locally.
Therefore, a major part of electric power is produced in remote power stations and transmitted to the consumers through electric power lines, which

1. Introduction

Figure 1.3. The structure of electric power production in different


European countries in 2006. [8].

constitute the power grid.


The main element of the electrical transmission and distribution grid is
the electric power line or circuit, which transmits power between two points
(buses) in the system. These lines are interconnected to meshed grids. A
three-phase AC line requires three conductors while single-phase AC lines
and DC lines usually only require two conductors. Depending on the insulation media the power circuits can be can be:
overhead lines,
cables, or
gas insulated lines (GIL)
For overhead lines the insulating medium is air, for cables a solid or solidliquid insulator and for gas insulated lines special gas, e.g. SF6 , with much
higher voltage insulation capacity than air is used. The latter two power
circuits can be made more compact and is therefore often used in densely
populated areas where land use is restricted. Furthermore, cables and GILs
are less exposed to influences from weather, lightning, and pollution, but on
the other side they are more expensive.

1.3. Transmission and Distribution of Electrical Energy

Figure 1.4. The structure of a power grid with different voltage levels
(Source: VSE).

1.3.1

Voltage Levels and Net Types

As illustrated in figure 1.4, an electric power grid consists of different voltage


levels.
Depending on the length of the power line and the required power capacity an economic optimization results in a certain voltage level. Higher
voltages lower the transmission losses but lead to increased costs of the
components. 2 Table 1.1 shows the usual denotation of the different supply
voltage levels.
Table 1.1. Levels of supply voltage

Level
Low Voltage
Medium Voltage
High Voltage

Abbreviation
LV
MV
HV

nominal voltage
below 1 kV
ca. 50 kV
above 110 kV

The structure of the electric grid can be compared to the road network for traffic. The 400-kV- and 220-kV- transmission lines serve for the
transport of bulk power over wide distances, e.g. from large hydro power
stations in the Alps or from nuclear power stations to the large substations
2

The relation between voltage and transmission losses is explained in section 5.1.2.

10

1. Introduction

Figure 1.5. Swiss high voltage net (source: ETRANS).

in the load centers, but also for the power exchanges between different countries and regions. (These latter type of lines are often called tie lines.) The
transmission lines constitute the so-called transmission grid and therefore
correspond to the highways in the road network. Figure 1.5 shows the high
voltage transmission network of Switzerland. The regional power distribution
(sub-transmission) in Switzerland is carried out by the 50-kV-distribution
network and allows the energy transport from the large substations to the
regional substations (in Switzerland denoted as Kantonswerke (cantonal
substations)). This net is comparable to the regional road network. The distribution from the regional substations to the single communities and loads
corresponds finally to the local roads and is carried out on a voltage level
of 16 kV and 10 kV, referred to as distribution network. As a comparison to
the Swiss grid figure 1.6 illustrates the high voltage net of Germany.
The major differences between the transmission and distribution net are
summarized in Table 1.2.

1.3.2

International Interconnected Grids and Synchronous Zones

Interconnection of two or more national transmission nets by tie lines results


in
a higher reliability of supply in the overall network and
the possibility to exchange electrical energy internationally, i.e. trad-

1.3. Transmission and Distribution of Electrical Energy

Figure 1.6. German high voltage net (source: VDN).

11

12

1. Introduction
Table 1.2. Comparison between transmission and distribution systems

Task
Voltage Level
Grid Topology
Covered area

Distribution System
Distribution to end-users
LV, MV
Radial or ring
community small region

Transmission System
Transmission of bulk power
HV
Meshed
country continent

ing.
The second point yields a fundamental basis to enable an electricity market
as no trading is possible without transportation possibility.
Basically, different systems can be coupled in two ways, namely
synchronously, i.e. through alternating /three-phase current connections or
asynchronously i.e. through dc current connections
If three-phase grids are connected directly, i.e. by a connection of all
three phases, these grids run synchronized and are accordingly referred to
as a synchronized zone. Additionally to the two above mentioned aspects,
these zones yield further advantages, especially concerning the operation,
which can be substantially more efficient.
The transmission grid of Switzerland is part of the European interconnected grid, namely of the Union for the Coordination of Transmission of
Electricity (UCTE). The transmission grids of 23 countries from Portugal to Poland are synchronized in this interconnection. About 450 million
people are supplied with energy by the UCTE net. The grids of Denmark,
Finland, Island, Norway and Sweden are part of the interconnected grid of
the Organization for Nordic Electrical Cooperation (NORDEL). The grid
of the UCTE and NORDEL are not synchronized. In order to still enable
an exchange of energy several dc current connections exist High Voltage
Direct Current (HVDC), e.g. between Germany and Sweden, realized with
submarine cables. Several states in Eastern Europe are interconnected with
states of the earlier Sovietunion, referred to as EES/VES -interconnection.
In northern Africa Morocco, Algeria, Tunisia etc. form the COMELEC interconnected grid.
In northern America, the gird of USA is organized in three synchronous
zones: Eastern Interconnection, Western Interconnection and ERCOT Interconnection (Texas).

1.3. Transmission and Distribution of Electrical Energy

13

Figure 1.7. Examples of grids with different security levels: left N 0


secure, right N 1 secure.

1.3.3

Grid Structures

A single point-to-point connection would lead to an interruption of supply


already at an outage of one single element in the power circuit; thus single
lines are not sufficient to enable a reliable supply of electricity. A grid of
power lines is needed by which the energy can be transported on parallel
paths in the case of a line outage. This is referred to as redundancy. To
guarantee a high availability, power grids should always be operated within
the following two system conditions:
under normal operation, if all components, as for example power stations, transformers or lines work properly, and
under outage operation, if a defined number of the operating components have failed.
If one arbitrary component is allowed to drop out without affecting the
power supply this is referred to as N 1 security. Analogous a N 2 secure
net is still operating with an outage of two arbitrary components. Under normal condition nets are mostly operated N 1 secure, special, very important
connections are N 2 secure.
Figure 1.7 shows two systems, which are supplied by the same bus. The
system on the left hand side is only interconnected through a transformer
and already the outage of this single component leads to an interruption of
supply. Thus, this system is not N 1 secure. The grid on the right hand
side however is still supplied even if one of the two supply lines drops out,
hence it is N 1 secure.
The consequent usage of the N 1/N 2 conditions results in a meshed
grid. Each branch can drop out without influencing the security of supply

14

1. Introduction

as enough parallel or redundant paths of supply exist which can compensate


the line outage. The transmission grid, which serves for the transmission of
large powers over long distances is built fulfilling these requirements.
However, a grid that is too meshed has also disadvantages: in case of a
fault, too many power circuits can lead to high short circuit currents which
are difficult to interrupt.3 Starting from a certain redundancy enabled by a
high number of circuits, the redundancy is improved only insignificantly by
adding more power lines.
Regarding the final distribution of electrical energy in smaller geographic
regions, such as a city, other aspects need to be considered. A strongly
meshed grid would imply certain disadvantages. Because of the short geographic distances the impedances between the points of delivery are low,
wherewith an unintentional high amount of power could flow through such
a grid. Because of economic reasons the lines are marginally designed on the
distribution level. Through additional superimposing powers the physical
rating limits could be exceeded. By this reason, the distribution nets are not
arranged meshed but as radial, namely formed as islands with radial distribution which most often are not N 1 secure. More and more so-called
open rings are established (see Figure 1.8). Under normal conditions, the
distribution bus bars are supplied according to Figure 1.8 (a). In the case of
a feeder outage, the ring is re-configured as shown in Figure 1.8 (b) and (c).
The system is quasi N 1 secure, though with a short interruption until the
new ring circuit is established.

1.3.4

Power Substations

Continuing the comparison of the power network and the road network the
power substations correspond to crossroads. They interconnect the different
voltage levels (e.g., 220 kV and 110 kV) and basically fulfill the following
tasks:
The power substations represent the nodes within the power grid. The
lines of the different voltage levels merge within the substations and
are coupled by transformers.
The power substations are switching stations, in which the in- and
outgoing lines depending on the operational demands can be
switched on and off; here, changes within the grid topology are possible.
In the power substations equipment for measurement, supervision, and
control, such as voltage- and current transformers, protection systems,
meters, compensation equipment, etc. are located.
3

A highly meshed network corresponds principally to a parallel connection of multiple


line impedances, resulting in a low effective impedance.

15

1.3. Transmission and Distribution of Electrical Energy

T1

T2

T1

T2

T1

T2

(a)

(b)

(c)

Figure 1.8. Open ring configuration (a), faulty situation (b), reconfigured ring after fault (c).

16

1. Introduction
In Switzerland, the big transmission utilities pass the electric energy
at the 50/16-kV and 110/16-kV substations to the cantonal utilities
which are responsible for the final distribution to end consumers.

1.3.5

Grid Operation and Supervision

The supervision of a grid is carried out via different regional grid control
centers which are subordinated to the larger control centers. These regional
grid control centers communicate their data with dedicated communication
channels, e.g. radio links, coax cables, and optic fibre cables, to the other
control centers and to the central control center. These communicated data
enable an overview of the operating state of the system at the control center
and allow the system operator to exercise a secure and optimal operation of
the system.
The electrical energy supply is not only characterized by energy (MWh)
and power (MW) but also by quality properties and demands. In order to
use electrical energy the preservation of the voltage form (sinusoidal) as well
as of a constant amplitude and a constant frequency are highly desirable and
important. By means of control and compensating equipment deviations of
the effective voltage from the nominal voltage are kept typically below 10%
at the customers side.
Another important aspect concerns the availability of electrical energy;
by economical reasons it is not realistic to expect a supply availability of
100%. On average, the availability of electrical energy supply amounts about
99.98% in Switzerland which corresponds to around 1.7 hour power outage
per year 4 This value is remarkable especially when considering that the
net planning and operation around-the-clock is optimized among others according to the following criteria: economic supply of power, reserves, fault
clearings, maintenance and a high utilization of equipment.
Furthermore, no energy can be stored in the power transmission system.
This fact complicates the electrical energy transmission compared with e.g.
a pipeline for the transport of natural gas. Accordingly, the actually needed
power must to be supplied at this actual time instant. If this is not the case,
the following will occur:
If the power demand of the loads is higher than the instantaneous
produced power, the net frequency drops below its nominal value (50
Hz).
4

It is questionable if such a high availability for private persons is needed when considering the increased usage of renewable energy producers and the thereby increased amount
of stochastic energy production. There are system operators with loads which indeed pay
lower current rates but could be disconnected by the system operator in critical situations
(depending on the agreement for example 10 hours a month).

1.4. Consumption of Electrical Energy

17

If the power stations produce more than the customers are currently
consuming, the net frequency rises above nominal frequency.
The frequency deviation within the UCTE net is normally controlled to be
on |f | < 0.05 Hz.

1.4

Consumption of Electrical Energy

As already mentioned, the percentage of electrical energy of the overall


worldwide final energy consumption amounts to 20%. The major part of
it is converted again at the end user into
mechanical energy,
chemical energy,
heat and
light.
The final consumption of electrical energy in industrialized countries is
divided almost equally into the areas of industry, public sector and households [1].
The per capita consumption of electrical energy varies significantly from
country to country, as shown in Figure 1.9. Iceland, Norway and Sweden are
the countries with the worldwide highest per capita electricity consumption.
The reason for this are the long cold winters with little daylight, but also
an extensive use of electric heating. Furthermore, in Iceland and Norway
there exist large industries that operate electrolysis processes consuming a
high amount of electric power. Japan, France and Austria together with
Switzerland lie on the OECD average.
The electrical energy consumption is subject to very strong temporal
fluctuations. The consumed power fluctuates
daily: In figure 1.10 mainly the peak at noon and at evening are salient.
The consumption is low at nights.
weekly: On working days the consumption is higher than on weekends.
Holidays are especially interesting, depending on the season they show
either high or low consumption.
seasonal and annual, respectively: In middle Europe more energy is
consumed during the winter than in summer. In more southern regions the situation is reversed. There one uses much more energy for
air conditioning in summer. In many regions where the peak load used
to be in the winter, the summer peaks has often increased significantly
due to an extensive installation of air-conditioning. This trend is believed to be accentuated in the coming years.

18

1. Introduction

Figure 1.9. The average per-capita consumption of electric energy in


different countries in the year 2008 (Based on data taken from the CIA
World Factbook).

In addition, the consumed power depends on many other, partly stochastic, factors such as the weather. The generation of electric power needs to
be adjusted to the consumption at every instance. The difficulty consists in
establishing an accurate load forecast for the proceeding day(s). If one does
not succeed in that, it may imply that one needs expensive regulating or
balancing power from the grid because of the unexpected load behavior and
the failing generation reserves.

19

1.4. Consumption of Electrical Energy

12

10

power in GW

12
Time of day

18

24 h

Figure 1.10. Consumption of electric energy on a winter day in Switzerland.

20

1. Introduction

2
Power in Alternating Current Systems
In this chapter the power of sinusoidal voltages and currents of single and
three-pase systems are analyzed. Basically, this serves as a repetition of already known fundamentals which form an important basis for the proceeding
chapters.

2.1

Power in Single Phase Networks

Consider an AC power grid in steady state. The power grid contains passive
elements(R, L, C) as well as pure sinusoidal current and voltage sources,
all with the same frequency. Thereby, all currents and voltages in this network are purely sinusoidal and could be described by their amplitude, their
phase angle based on an arbitrarily defined zero position, and by their frequency. The instantaneous values of voltage and current of an element can
be described by the following equations:
b cos (t)
u (t) = U
i (t) = Ib cos (t )

(2.1b)

= 2f

(2.2)

(2.1a)

b and Ib are the amplitudes whereas denotes the phase shift


Thereby U
between current and voltage. The angular frequency is defined as

where f corresponds to the nominal frequency of the network (for example


50 Hz in Europe, 60 Hz in northern America). Often, the root mean square
b and Ib are used; these
values 1 U and I instead of the amplitude values U
magnitudes are related as:
b
U
U=
2
Ib
I=
2
1

(2.3a)
(2.3b)

The root mean square value, or rms value, of an alternating current corresponds to
the value of a dc current, which produces the same heat in a resistance as the alternating
current does.

21

22

2. Power in Alternating Current Systems

In order to compute the instantaneous value of the power we multiply


the instantaneous values of the voltage and the current and obtain
b Ibcos (t) cos (t )
p(t) = u(t) i(t) = U


b Ibcos (t) cos (t) cos + sin (t) sin
=U


b Ib cos2 (t) cos + sin (t) cos (t) sin
=U

(2.4)

By applying the relations


1
1 + cos (2t)
2
1
sin (t) cos (t) = sin (2t)
2
cos2 (t) =

(2.5a)
(2.5b)

the instantaneous value of the power is given by



 1
1 bb
b Ib sin sin (2t)
p (t) = U
I cos 1 + cos (2t) + U
2
2
|
{z
} |
{z
}
1

(2.6)

These terms are depicted in figure 2.1. Considering the first part one can
b Ib cos
see that the power oscillates between zero and the maximal value U
because of the cosine term in brackets. The time averaged value of this term
can be expressed by the rms values from equations (2.3a) and (2.3b) as
P =

1 bb
U I cos = U I cos
2

(2.7)

This time averaged value P is called active or real power whereas the factor
cos is called the power factor or simply cos . The SI unit of active power
is Watt (W).
The time averaged value of the second term in equation (2.6) is zero. Due
b Ib sin .
to the sinus term with the frequency 2 it oscillates between 12 U
The amplitude of this oscillation is
Q=

1 bb
U I sin = U I sin
2

(2.8)

and is called reactive power. The unit of the reactive power is volt ampere
reactive (Var).
Sinusoidal quantities can also be described as phasors in the complex
plane. The time function x(t) corresponds to the rotating phasor X:
b
X
b cos (t + ) X =
x(t) = X
(cos + j sin ) = Xej
2

(2.9)

The instantaneous value of the time function x(t) at the time instant t
corresponds to the real part of the phasor which is rotating with the angular
velocity :

23

2.1. Power in Single Phase Networks


p(t)
p(t)

Zeit

2
p (t) =


 1
1 bb
b Ib sin sin (2t)
U I cos 1 + cos (2t) + U
|2
{z
} |2
{z
}
1

Figure 2.1. Plot of the terms in equation (2.6).

x(t) =



2 Xejt

(2.10)

Accordingly, the time functions of current and voltage can be expressed as


complex quantities:
U = U (cos u + j sin u ) = U eju
ji

(2.11a)

I = I (cos i + j sin i ) = Ie

(2.11b)

= u i

(2.11c)

The angles u and i denote the phases of current and voltage. Taking the
voltage as reference, i.e. u = 0 the currents phase results in relatively
to the voltage (similar to equation (2.1b)).
Forming the product of current and voltage, we get
U I = U Iej(u +i )

(2.12)

If one prefers to have the already as defined phase difference between


voltage and current in the exponent of this expression, the voltage needs
to be multiplied with the conjugate complex value of the current I . This
product is defined as complex apparent power S:
S = UI = U Iej = U I (cos + j sin )

(2.13)

24

2. Power in Alternating Current Systems

Figure 2.2. Power triangle relating the active, reactive and apparent powers.

The magnitude of this quantity is called apparent power S:


(2.14)

S = |S|

The (complex) apparent power is denoted as volt ampere (VA), correspondingly to voltage times current.
By use of the conjugate complex value of the current the active power
P and the reactive power Q correspond according to (2.7) and (2.8) to the
real and to the imaginary part of the complex apparent power, respectively:

P = {S} = {U I } = U I cos

Q = {S} = {U I } = U I sin
S = P + jQ
p
S = |S| = P 2 + Q2

(2.15a)
(2.15b)
(2.15c)
(2.15d)

In figure 2.2, these relations are represented as phasors in the complex plane,
also referred to as the power triangle.
The terms active, reactive and apparent power will now be illustrated
with two examples.
Example: ohmic inductive load Given is the network in figure 2.3. The input power at the parallel connection consisting of resistance and inductance
is to be calculated at steady state.
i(t)

iR (t)

b cos (t)
u(t) = U

iL (t)

Figure 2.3. Ohmic-inductive load

25

2.1. Power in Single Phase Networks

Using the Kirchhoffs Current Law (KCL), the overall current (in steady
state operation) is obtained as:
i(t) = iR (t) + iL (t)

b
b
U

U
cos t
i(t) = cos (t) +
R
L
2
The instantaneous power is thus

b2
b2
U
U

cos2 (t) +
cos (t) cos t
R
L
2
2
2
b
b
1 U
1 U
=
(1 + cos (2t)) +
sin (2t)
2 R
2 L
U2
U2
=
(1 + cos (2t)) +
sin (2t)
{z
} |L {z
}
|R

p(t) = u(t) i(t) =

pR (t)

pL (t)

The overall input power to the network p(t) is composed by two terms,
namely by the power at the resistance pR (t) and the power at the inductance
pL (t). Using the equations (2.6), (2.7) and (2.8) yields the active and reactive
power as
P =

U2
R

und Q =

U2
U2
=
L
XL

mit XL = L

The value of the reactive power is positive, i.e. the inductor absorbs reactive
power, it consumes reactive power. This power value corresponds to the
amplitude of the oscillating magnetic energy and its time averaged value
amounts to zero.
The power consumed by the inductance could also be computed via
the magnetic energy. Computing the energy wL (t) which is stored in the
inductance at the time instance t, we obtain
1
wL (t) = Li2L (t)
2
The current in the inductance iL (t) amounts for a sinusoidal voltage to

b
U

cos t
L
2
Inserting this relation into the energy equation we obtain

b2
1 U

wL (t) = L 2 2 cos2 t
=
2 L
2
b 2  cos(2t ) + 1 
1 U
=
=
2 2L
2
b2
1 U
= 2 (1 cos (2t))
4 L
iL (t) =

26

2. Power in Alternating Current Systems

The power consumed by the inductance is obtained by differentiating this


energy with respect to the time:
dwL (t)
=
dt
b2
U
sin (2t) =
=
2 2 L
U2
sin (2t)
=
L

pL (t) =

As seen we obtained the same result for the power in the inductance as by
multiplication of current and voltage. The consumption of reactive power in
the inductance corresponds to the amplitude of the time derivative of the
magnetic energy.
As a third possibility, we can compute the power using the complex
phasors. The overall current as complex phasor amounts


U
1
1
I=
=U
+
Z
R jL
Regarding equation (2.13) the overall complex apparent power amounts to
S = U I
Assuming the voltage with reference angle 0 (pure real) results in U =
U = U and
U2
U2
S=
+j
R
L
The real part of the apparent power is the active power which is consumed by
the resistance. The imaginary part of the apparent power equation yields the
reactive power in the inductor. Here, the sign is positive, i.e. the inductance
absorbs or consumes reactive power. This yields:
P = {S} =

U2
R

and

Q = {S} =

U2
L

Example: ohmic capacitive load Equivalent to the example above, the


power consumed by an ohmic capacitive load can be computed. We consider
the circuit in figure 2.4.
Analogous to the inductive load we obtain the active and reactive power
as
P =

U2
R

und Q = U 2 C

Here, the sign of the reactive power is negative, i.e. the capacity delivers

27

2.2. Conservation of Apparent Power


i(t)

iR (t)
b cos (t)
u(t) = U

iC (t)

Figure 2.4. ohmic-capacitive load

reactive power, it produces reactive power. Thus, capacitors can be used to


compensate the reactive power consumed by the inductances. Later, we will
consider the influence to the system voltage by compensation of reactive
power.
The capacitive reactive power can be calculated by the energy approach
as well. The basis is the equation for the energy in the electric field of a
capacitor
1
wC (t) = Cu2C (t)
2
Similarly, the same expressions are obtained as by multiplication of the
time functions of current and voltage.

2.2

Conservation of Apparent Power

When working with complex apparent power we often use the Theorem of
the conservation of apparent power. This indicates that the sum of apparent
power delivery of all sources is equal to the sum of the apparent power
consumption by all loads or sinks. This is valid in a network consisting of
multiple sources and consumers, each of them independent of each other.
Thereby one assumes all currents and voltages to be purely sinusoidal and
having the same frequency.
For a single source this theorem can be proved by the use of Kirchhoffs
laws. The proof is more complicated for the general case.
By usage of this theorem one often replaces networks by equivalent
sources (Thevenin equivalents). Figure 2.5 illustrates a net Nb which is supplied by two sources and a further net Na . The sum of the apparent powers
amounts
X
S ab + S 2 + S 3 =
S b,i
(2.16)
i

where S b,i represents the apparent power of element i in net Nb .

Example: series impedance The conservation of apparent power is investigated by means of a series element with an impedance of
Z = R + jX

28

2. Power in Alternating Current Systems

S2
Na

S ab

Nb

S3

Figure 2.5. Preservation of the apparent power.

Figure 2.6 depicts the corresponding circuit.


I1
S1
U1

I2

S2

U2

U1

I
U2

Figure 2.6. Circuit of a serial impedance and phasor diagram.

The complex apparent power at the input of a two-port (left hand side,
index 1) amounts
S 1 = P1 + jQ1
In the series circuit the three indicated currents need to be equal:
I1 = I2 = I
The current will yield a voltage drop U at the impedance by which the
output voltage differs from the input voltage (see phasor diagram in figure
2.6). Thus, the voltage at the output is
U 2 = U 1 U = U 1 ZI
The apparent power at the output of the two-port (right hand side, index
2) is obtained as a product of voltage times conjugate complex current:
S 2 = U 2 I 2 = U 2 I
Inserting the expression for U2 , we obtain for the output power
S 2 = (U 1 ZI) I = S 1 Z|I|2

2.3. Power in Symmetric Three-Phase Networks

29

We can now insert S 1 according to the first equation and obtain




S 2 = P1 R|I|2 +j Q1 X|I|2
|
|
{z
}
{z
}
P2

Q2

Input and output power of the two-port differ exactly by the power consumed
by the impedance Z.
The circuit shown in figure 2.6 can be regarded as a simple model of a
high voltage power line, as will be shown later.

2.3
2.3.1

Power in Symmetric Three-Phase Networks


Symmetric Three-Phase Systems

A symmetric three-phase system is supplied by three voltages with equal


amplitudes, having equal frequency, but each phase shifted by 120 (2/3)
with respect to the others. Figures 2.7 and 2.8 show the time variation of
these three voltages. The three phases are usually denoted by R, S, and
T, but sometimes also as A, B, and C. In this compendium we will use R,
S, and T. Voltages, currents, resistances, etc. are indicated similarly. The
voltage of phase R is assumed to be the reference phase, thus with phase 0 ,
as illustrated in figure 2.8. Accordingly, the voltages of phase S and T are
shifted by 120 and 240 , respectively. This voltage wye or voltage star
rotates anti-clockwise with the angular velocity such that the phase order
R-S-T evolves.2
As indicated in figure 2.8, two different voltages occur in an alternating
current system: the phase voltage and the phase to phase voltage.
Phase voltage Up : This voltage is measured between one phase and the
neutral. It corresponds to the voltage at an element occurring in wye
or Y-connection (see figure 2.9). The phase voltages are described by
UR , US and UT
Phase to phase voltage U : This voltage is measured between two phases.
According to the geometry
in the phasor diagram in figure 2.8 the value
of this voltage is higher by 3 than the value of the phase voltage. The
phase to phase voltage is the voltage at an element when having a delta
circuit (see figure 2.9). In the three-phase system three different phase
to phase voltages occur between the phases R-S (URS ), S-T (UST ) and
T-R (UTR ). These three voltages are phase shifted by 120 , analogous
to the phase voltages.
2

This convention is arbitrary. To receive the same phase order R-S-T at reverse rotating
direction the phases S and T need to be exchanged.

30

2. Power in Alternating Current Systems


u(t)

uR (t)

uS (t)

uT (t)

4 t

Figure 2.7. Time related devolution of the voltages of the phases R, S and T.

As mentioned, the root mean square values of the phase voltages and the
phase to phase voltages have the following relation:

U = 3 Up
(2.17)
At low level voltage, the values of the phase voltages and the phase to phase
voltages are 230 and 400 V, respectively.

31

2.3. Power in Symmetric Three-Phase Networks

UTR
UT

UR

UST

Up

120

US
Up

URS

Figure 2.8. Phasor representation of the three phase voltages(magnitude Up ) and of the phase to phase voltages(magnitude U = 3 Up ).

Up
U

U
U

Figure 2.9. Wye and delta configuration of a consumer: At wye configuration the phase voltage Up lies over the load resistances, at delta
configuration the phase to phase voltage U .

32

2.3.2

2. Power in Alternating Current Systems

Power in Symmetric Three-phase Systems

Now, the power in the symmetric alternating current systems will be investigated. Again, we start with the time representations of the voltages in the
corresponding phases (see figure 2.7):
bp cos t
uR (t) = U
bp cos(t 2/3)
uS (t) = U

(2.18a)
(2.18b)

bp cos(t 4/3)
uT (t) = U

(2.18c)

uR (t) + uS (t) + uT (t) = 0

(2.19)

Because of the assumed symmetry all three phase voltages have the same
bp . The sum of these phase voltages at each instant is
amplitude U
Equivalently, for the three currents
iR (t) = Ibp cos(t )
iS (t) = Ibp cos(t 2/3)

(2.20a)
(2.20b)

iT (t) = Ibp cos(t 4/3)

(2.20c)

iR (t) + iS (t) + iT (t) = 0

(2.21)

and again with the assumption that the current magnitudes are equal for
all phases

The part of the instantaneous power from phase R amounts according


to equation (2.6) to
pR (t) =

1 b b
Up Ip (cos (1 + cos (2t)) + sin sin (2t))
2

(2.22)

The angle corresponds to the phase shift between current and voltage,
that is why the sin - and cos - terms are time independent and equal for
all phases.
The sin (2t) and cos (2t) expressions occur in the other two terms as
well, but phase shifted by 2/3 and 4/3, respectively. By means of well
known trigonometrical theorems one can show, that these terms cancel out:
 

 

2
4
cos (2t) + cos 2 t
+ cos 2 t
=
| {z }
3
3
|
{z
} |
{z
}
Phase R
Phase S
Phase T
 
 
4
4
= cos (2t) + cos (2t) cos
+ sin (2t) sin
+
3
3
 
 
8
8
+ cos (2t) cos
+ sin (2t) sin
= 0 (2.23)
3
3

2.3. Power in Symmetric Three-Phase Networks

33

where



8
1
cos
= cos
=
3
2

 
 
4
8
3
= sin
=
sin
3
3
2
4
3

(2.24a)
(2.24b)

The instantaneous value of the overall three phases, i.e. the sum of the
instantaneous powers of all three phases, therefore results in a time constant
value and corresponds to the three times the active power of one phase:
p(t) = 3

bp Ibp
U
cos = 3 Up Ip cos = P
2

(2.25)

This means, the oscillating parts of the power in the three phases add to
zero at each time instant and not only in the time averaged value.3
The complex three-phase apparent power results from the sum of all
phase apparent powers using equation (2.13)
S = U R I R + U S I S + U T I T

(2.26)

In the symmetric state, all three phase currents and voltages of the same
size, the three-phase apparent power can therefore be calculated as three
times the apparent power of one single phase. For phase quantities, one
usually uses by definition the voltage and current of phase R, where the
angle of phase R is assumed to be 0 .
S = 3 U R I R = 3 S R = 3PR + j3QR

(2.27)

Now, the phase to phase quantities instead of the phase quantities of phase
R are inserted into the equation above. The phase to phase voltage results
according to equation (2.17) in

U = 3 UR
(2.28)
Actually, a phase to phase current in the
strict sense is not existing,
this means, there occurs no current of the size 3 times the phasor current
in a alternating current system physically. Nevertheless, such a current can
be defined as a virtual, phase to phase current in order to simplify the
calculation. We therefore define

I = 3 IR
(2.29)
With
this definition, we can reformulate (2.27). Thereby, the factors 3 and

3 cancel out and we obtain for the complex three-phase apparent power
3

This has the consequence of having a constant torque over one rotation at three-phase
machines. By contrast, the rotor of a single phase motor gets a ripple moment of double
nominal frequency

34

2. Power in Alternating Current Systems

with the conventions in (2.28) and (2.29) the same formula as for the single
phase apparent power in (2.13):
S = U I = U Iej = U I (cos + j sin )
Correspondingly, the known equations (2.15a)-(2.15c) apply for the active
and reactive power as well:
P = {S} = {U I } = U I cos

Q = {S} = {U I } = U I sin
S = P + jQ

These values correspond in the symmetric system exactly to the triple power
per phase. The formulas (2.13)-(2.15c)thus apply for single phase as well as
for three-phase system, as long as the 3-times phases sizes are inserted for
the three-phase relations of U and I.

2.3.3

Wye-Delta-Transformation

Each wye or delta connection can be transformed into an equivalent delta


or wye connection, respectively, consuming the same amount of power. If,
for example, only the power of a three phase consumer is given it can be
described arbitrarily as a wye or delta load.
We consider again figure 2.9. The impedances in the wye connection are
described by Z y , those of the delta connection by Z d . In a wye connection
the phase voltages of size Up are over the impedances. The overall apparent
power results in
U2
Sy = 3 R
(2.31)
Z y
At the impedances
of the delta connection, the phase to phase voltages of

size U = 3Up are over the impedances. The total consumed apparent power
is

2
3U R
Sd = 3
(2.32)
Z d
Both connections are then equivalent when they consume the same apparent
power. We therefore put
Sy = Sd
(2.33)
and obtain the following requirement for the impedances
Zy
3U 2R
1
=

2 =
Zd
3
3 3U R

(2.34)

A delta connection is therefore at each instant transformable in a wye connection by multiplying the resistances of the wye connection by a factor of

35

2.4. Summary

3 and connecting them the in delta formation. The new connection consumes the same amount of apparent power as the original one and the phase
conductor carries the same current.

2.3.4

Single Phase Calculation of Symmetric Three-phase Systems

As the same voltage and current situations arise in all three phases in a symmetric alternating current system the three phase system can be analyzed
on the basis of an equivalent single phase system. For this, the following
steps are taken:
1. All elements which are connected in delta form are transformed in a
equivalent wye connection (see section 2.3.3).
2. A single phase diagram for phase R is drawn .
3. The sought quantities are calculated by means of the single phase
diagram of phase R.
4. To obtain the quantities of phases S and T 120 and 240 , respectively,
are in each case added to the sizes of phase R.
5. If necessary, the elements in wye connection are transformed back in
delta connection for then calculating the phase to phase quantities.
Figure 2.10 shows an example of a symmetric three phase system. The
capacitors C connected in delta form are transformed into a wye connection.
Their reactance reduces by the transformation according to equation (2.34)
to a third comparing to the value of the delta connection. Due to
jXC =

1
jC

(2.35)

this corresponds to a multiplication by three of the capacity.


Having performed the transformation the single phase diagram can be
sketched. Thereby one needs to consider that all neutral points have the
same potential in a symmetric alternating current system. The points a,
f and e can by therefore be connected without any consequences. These
neutral points are in the single phase diagram on the same return current
conductor.

2.4

Summary

In single phase networks the power oscillates with double power frequency.
In symmetric alternating current systems the oscillating parts or the three
phases balance out and the power is constant in time.

36

2. Power in Alternating Current Systems

Up
L

R
c

3C

L
b

Up

3C
a

R
e

Figure 2.10. Symmetric three phase system and its single phase diagram. Having transformed the delta connection into a wye connection,
the capacities are in parallel to the resistances R. In the single phase
diagram, the neutral points a, f and e are all on the common return
conductor.

37

2.4. Summary
Table 2.1. Overview or real, reactive and complex apparent power.

Denotation
Symbol
Unit
Calculation

Active power
P
MW
U I cos = {S}

Reactive power
Q
MVar
U I sin = {S}

Complex apparent power


S
MVA
U I = P + jQ

The complex apparent power is defined as the product of voltage times


the conjugate complex current. The real part of the complex apparent power
describes the active power, the imaginary part the reactive power.
For the calculation of apparent, active and reactive power the same equations can be used for single phase and symmetricthree-phase systems.
For

the overall three-phase power in each case I = 3 I R and U = 3 U R is


inserted. In Table 2.1 the formulas and entities are summarized once more.

38

2. Power in Alternating Current Systems

3
Transformers
In this chapter a model for power transformers is developed. We start with
the principle of coupled coils and proceed with a idealized model of a single
phase transformer. This basic model is enhanced step wise to a model suitable
for power engineering applications. Finally, we consider models of threephase transformers.

3.1
3.1.1

Single Phase Transformer


Coupled Windings

A current through a coil produces a magnetic field. Two coils in a common


magnetic field will influence each other. Due to the magnetic coupling the
two circuits will interact so that every change of the current in one coil will
cause a current change in the other coil.
If two coils are placed on a common magnetically well conducting magnetic core, one reaches a situation where almost the same magnetic flux flows
through both inductances. Thus, the coupling is strong in this case. Figure
3.1 shows such a configuration. This is basically the setup of a single phase
power transformer.
primary side

secondary side

i2
i1

u1

u2

N2

N1

Figure 3.1. Basic setup of a single phase power transformer.

The limbs together with the yoke constitute a magnetically well conducting connection between the primary- and the secondary winding. The
39

40

3. Transformers

equidirectional

reverse

Figure 3.2. Labelling of the winding direction by points at the windings.

primary winding having N1 number of turns carries the primary current


i1 , whereas the secondary current i2 flows through the secondary winding
having N2 number of turns.1
In the iron core, the mutual flux h flows. This flux is flowing through
both windings and constitutes the magnetic coupling between them. The
higher its part on the overall flux, the stronger the interaction between the
windings. Additionally to the main flux, each winding is establishing its own
leakage flux 1 and 2 . These are relatively small in comparison with the
mutual flux for normal power transformers. The magnetic flux lines flow
only through one inductor and close via the air.2 The flux linkage in the
inductances result in
1 = N1 h + 1

(3.1a)

2 = N2 h + 2

(3.1b)

where 1 and 2 describe the flux linkages of the leakage fluxes 1 and
2 with the corresponding inductances.
The direction of the magnetic flux in dependency of the current is determined by the winding direction of the inductor. One distinguished between equi-directional and reverse windings. Usually, the winding direction
is marked by points at the coil taps (see Figure 3.2). If these points are
aligned, one refers to an equidirectional winding; if they lie diagonally the
configuration is in reverse. In the following investigations we always assume
equidirectional windings.

3.1.2

Ideal Transformer Model

The basic function of a transformer can be demonstrated with an idealized


model of two interlinked coils. The following assumptions are made for the
ideal transformer :
1. There are no losses in the transformer, neither in the windings nor in
the iron core.
1

The current i2 = i2 will be introduced later to simplify the derivation of the transformer model.
2
Not all flux lines of the leakage flux are interlinked with all windings of the inductor.

41

3.1. Single Phase Transformer

2. No leakage fluxes appear, i.e. the same mutual flux flows through both
windings. Hence, the coupling between the two coils is ideal.
3. The iron core has an infinitely high permeability (this corresponds to
an ideal magnetic conductance).
We will now examine an ideal transformer as depicted in Figure 3.1. Using
the induction law and the second assumption
1 = 2 = 0

(3.2)

we obtain the following voltages induced in the primary and secondary windings
d1
dh
= N1
dt
dt
d2
dh
= N2
u2 =
dt
dt
u1 =

(3.3a)
(3.3b)

When assuming a time varying flux, i.e.


dh
6= 0
dt

(3.4)

dh /dt can be eliminated from the equations (3.3a) and (3.3b) yielding
u1
u2
=
N1
N2

(3.5)

which means that the ratio between the voltage and the number of turns
is equal for the primary and secondary windings. The turns ratio, c, is now
defined from3
u1
N1
=
=cR
(3.6)
u2
N2
Since the number of turns is a real value for both windings the turn ratio is
real as well.4
Corresponding to the third assumption ( = ) the magnetic resistance
of the core is zero. Ohms law of the magnetic circuit gives a relation between
the magnetic voltage excitation , the magnetic flux and the reluctance
Rm :
= Rm
(3.7)
The magnetic voltage is the so-called ampere-turns, i.e. the product of current times the number of turns of a coil. For the ideal case Rm = 0, there
3
Strictly speaking, this concerns the no load turns ratio. The non ideality of real transformers considered in section 3.1.3 leads to a deviation which is load independent.
4
For three-phase transformers, a phase shift between the secondary and primary voltage
can occur which means that the turns ratio is complex (see section 3.2).

42

3. Transformers

u1

primary

secondary

i1

i2

N2

N1

u2

Figure 3.3. Ideal transformer.

is no magnetic voltage drop along the iron core. The sum of the ampere
windings of primary and secondary side result in
N1 i1 + N2 i2 = Rm h = 0

(3.8)

The quotient between the secondary and primary currents yields the turns
ratio
N1
i2
=
= c
(3.9)
i1
N2
Just as for the voltage also the current on the secondary side is dependent
only on the primary current and on the turns ratio. We now define
i2 = i2

(3.10)

and obtain the situation as depicted in Figure 3.3, and the relation between
the currents can be written as
i2
=c
i1

(3.11)

When assuming sinusoidal currents, the magnetic flux and the thereby
induced voltages are also sinusoidal. Sinusoidal currents and voltages can be
described by phasors according to equation (2.9).
In the equations (3.1a)-(3.11) u, i and can be substituted by the corresponding phasors U, I and .
Thus we know how to translate voltages and currents between the primary and secondary sides. Now we want to examine how to convert impedances from one to the other side in the case of an ideal transformer. This conversion is called impedance transformation. Considering Figure 3.4, we can
set up on each side of the ideal transformer an equation:
U 1 = E a1 = c E a2 = c (U 2 + I 2 Z 2 )
1
1
U 2 = E b2 = E b1 = (U 1 I 1 Z 1 )
c
c

(3.12a)
(3.12b)

43

3.1. Single Phase Transformer

U1

I1

I2

E 1a

E 2a

N1

Z2

Z1

U2

U1

I2

I1

E b1

N2

E b2

N1

U2

N2

impedance transformation
Figure 3.4. Impedance transformation from secondary to primary side.

These equations can be solved for c U 2 :


c U 2 = U 1 c I 2Z 2

c U 2 = U 1 I 1Z1

(3.13a)
(3.13b)

which means that


which can be rewritten as

0 = c I 2 Z 2 + I 1 Z 1 .

(3.14)

Z1
I2
=c
= c2
Z2
I1

(3.15)

Both circuits in Figure 3.4 behave equivalently as long as equation (3.15) is


fulfilled. The behavior of the two port is not changed if the impedances are
exchanged according to this ratio. Using equation (3.15) impedances can be
converted arbitrarily between the primary and secondary sides. 5
The transformations for voltages, currents and impedances were derived
above and we are now able to refer each of this quantities to the other side.
Recapitulating:
The voltages behave according to the turns ratio.
The currents act inversely to the turns ratio.
The impedances perform according to the square of the turns ratio.
As mentioned, the transformer was assumed to be ideal in this first representation. To obtain a more realistic model we extend the ideal model with
additional elements.

3.1.3

Realistic Transformation Model

The ideal transformer model in section 3.1.2 is now improved step by step
by considering three important non-idealities.
5

Equation (3.15) is not only valid for elements described as complex numbers but also
for R, L or C.

44

3. Transformers

Flux leakage
The leaking flux interlinked with only one winding can be considered to be
caused by separate coils in the primary and the secondary circuits. In each
case we obtain a leakage reactance for the primary and secondary side with
leakage inductance L1 and L2 , respectively. The leakage fluxes flow in the
air and are therefore linear:
1 = L1 i1

(3.16a)

L2 i2

(3.16b)

2 =

The coupling between windings becomes therefore incomplete due to the


leakage fluxes, they present a non-ideality of the transformer.
Copper Losses
Real windings have ohmic losses. Considering this, the primary and secondary terminal voltage result from the sum of the ohmic voltage drop over
the winding and the induced voltage
d1
d
= R1 i1 + (1 + N1 h )
(3.17a)
dt
dt
d2
d
u2 = R2 i2 +
= R2 i2 + (2 + N2 h )
(3.17b)
dt
dt
where R1 , R2 denote the resistances of the primary and secondary winding,
respectively. Incorporating equations (3.16a) and (3.16b) we obtain
u1 = R1 i1 +

di1
dh
+ N1
dt
dt

di
d
h
u2 = R2 i2 + L2 2 + N2
dt
dt

u1 = R1 i1 + L1

(3.18a)
(3.18b)

With equation (3.10) we can reformulate (3.18b) into:


di2
dh
+ N2
(3.19)
dt
dt
We incorporate these further deviations from the ideal behavior by inserting
the two elements in the transformer model (see figure 3.5).
As shown in section 3.1.2 we can now convert the impedances of the
secondary side to the primary side. Figure 3.6 shows the new equivalent
network with the elements
u2 = R2 i2 L2

Rt = R1 + c2 R2

(3.20)

Lt = L1 + c2 L2

(3.21)

and
This model includes non-idealities originating from flux leakage and ohmic
losses in the windings.

replacemen
45

3.1. Single Phase Transformer


R1

L1

i1

ideal

R2

L2

i2

u1

u2

N1

N2

Figure 3.5. Transformer model with primary and secondary leakage


inductances and winding resistances.
Rt

Lt

i1

ideal

u1

i2

u2

N1

N2

Figure 3.6. Transformer model with resistances and leakage inductances transformed to primary side.

Core Losses
The magnetic conductor, the iron core, is not completely ideal either. Voltage
and frequency dependent losses are occurring in the conductor. Therewith,
the transformer delivers less power on the secondary side than it consumes
on the primary side. The transformer carries a current on the primary side
although no current is flowing on the secondary side, as the losses in the
iron need to be modelled. These are referred to as no load losses. They are
composed of the primary winding losses and the core losses.
The reactive part of the losses in the iron can be modelled as a magnetization inductance. This consumes the magnetization current which is
calculated in the following.
We start again with the Ohms law of the magnetic circuit according
to equation (3.7). Considering the magnetic resistance of the core Rm 6= 0
yields
= Rm h = N1 i1 + N2 i2
(3.22)
Assuming a current of i2 = 0 on the secondary side we obtain the magnetization current as

Rm h
i1 i =0 = im =
(3.23)
2
N1

Taking into account this equation as well as eq. (3.22) the primary current

46

3. Transformers

results as a sum of the magnetization current and the secondary current,


translated to the primary side:
i1 = im

N2
i2
i2 = im +
N1
c

(3.24)

The primary current is the sum of the magnetization current and the transformed secondary current. The magnetization current is tapped off between
primary and secondary side and amounts to less than 1% of the nominal current for realistic power transformers. In the circuit we can incorporate this
current by introducing a magnetizing or main inductance Lh (see Figure
3.7). The value of this inductance can be derived from the equation for the
Rt

Lt

i1

i2

ideal
im

u1

u2

Lh

N1

N2

Figure 3.7. Transformer model with magnetizing inductance, Lh .

induced voltage
dh
dim
= Lh
(3.25)
dt
dt
For real transformers, this inductance is very large as compared with the
leakage inductances
Lh Lt
(3.26)
N1

The active losses in the core are modelled by a resistance Rh . In the circuit,
we introduce this in parallel with the magnetization inductance. Analogous
to the inductances we have
Rh Rt
(3.27)
Rt

Lt

i1

ideal
ir

u1

Rh

i2

im
u2

Lh

N1

N2

Figure 3.8. Complete transformer model.

47

3.2. Three-Phase Transformer

The model in figure 3.8 now takes all the above discussed deviations from
the ideal transformer into account. These are summarized again:
Flux leakage (inductive by Lt )
Winding losses (ohmic by Rt )
Core losses (ohmic by Rh and inductive by Lh )
Because of (3.26) and (3.27) the shunt elements Rh and Lh are often neglected at nominal operation conditions of the transformer. The transformer
can then be modelled as a complex serial impedance Z t = Rt + jLt and
an ideal transformer with turns ratio c = N1 /N2 . Figure 3.9 shows such
an equivalent model. This represents an important and often used model in
power system analysis. In no-load operations (i2 = 0), these shunt elements
can often not be neglected.
Zt

I1

ideal

I2

U2

U1

N1

N2

Figure 3.9. Simplified transformer model for energy transmission.

3.2

Three-Phase Transformer

To transform the voltages of all three phases of a three-phase system, three


individual single phase transformer transformations can be performed. Basically, there exist four possibilities to connect three-phase transformers using
wye and delta connections: primary and secondary windings can be connected in each case either in wye or delta configuration 6
The transformation ratio and the phase shifting between primary and
secondary winding as well as the behavior at asymmetric operation depend
on the so-called vector group of the transformer, apart from that each configuration features different grounding possibilities, more about this in later
chapters.
The labelling of the vector group is carried out along a standardized
scheme consisting of two letters and one number:
6

Apart from wye and delta connection there exists for example the possibility of a
so-called zig-zag connection, analogous labelled as Z or z connection. Other types of connections are also possible.

48

3. Transformers
1. The connection of the high voltage winding is labelled with a capital
letter (D for delta or delta and Y for wye).
2. The connection for low voltage winding is described with lower case
letters (d or y).
3. A number which indicates the phase shift between primary and secondary winding as a multiple of 30 is also added (n in equation (3.31),
counterclockwise).

The notation Yd5, as an example, represents high voltage wye connection,


low voltage delta connection and a phase shift of 5 30 = 150 . Equal
connections of primary and secondary sides yields the connection group Yy0
or Dd0.
For power infeed in high voltage nets the Yd configuration is often used.
First, the physical winding turns ratio is optimally utilized, second a neutral
conductor can be designed and grounded on the high voltage side. In Fig.
3.10 such a configuration is depicted. Between the star node of the transformer and earth a grounding impedance (earthing impedance) is connected.
Features and purpose of this arrangement will be discussed in chapter 7.
R

Transmission Grid

Generator

T
S

ZE

Figure 3.10. Yd-connection for connecting a generator or a distribution system to the transmission grid. On the high voltage side the
transformer neutral point is grounded via an impedance Z E .

3.2.1

Design of Three-Phase Transformers

It is also possible to put all primary and secondary windings on a common


magnetic conductor. Therewith, one saves material and weight which plays
an important role in relation to the transport. Apart from that, less space
is required compared with three single transformers. The savings can be
compared to those of a three-phase cable vs. three single phase cables.
Iron core and windings can be designed and configured in many different
ways. Figure 3.11 shows the basic construction of a five limb transformer.

49

3.2. Three-Phase Transformer

Three limb transformers are also common. Their principal design is similar
to the five limb configuration, only the additional not winded limb (left and
right) are missing.
primary winding
yoke

limbs

secondary winding

Figure 3.11. Five limb three-phase transformer.

In symmetric operation the flux in all three limbs add to


R + S + T = 0

(3.28)

This is the reason why the resulting flux in the yoke is zero. Only at asymmetric conditions the flux over the yoke and the possibly existing additional
limbs (fourth and fifth) are non-zero. The design of the iron parts has therefore a considerable influence on the operation behavior at asymmetric load.

3.2.2

Three-Phase Transformer Model

In principal, the single phase model from section 3.1.3 can be used for modelling each single phase of the three-phase transformer.
Considering the conditions at Yd or Dy configurations, namely for unequal connections of primary and secondary sides, two aspects can be observed:
1. The same voltages are not connected at primary and secondary inductances. The phase voltage is at one side, the phase to phase voltage at
the other side. Considering for example the Yd configuration in figure
3.10, the phase to phase voltage U lies over the generator side windings and the phase voltage U p is applied over the high voltage side
windings. Therewith, the actual turns ratio changes compared with
the winding ratio.
2. Phase to phase and phase voltages differ concerning the phase shift
too. Hence, additional to the amplitude a change of the phase occurs.
Depending on the vector group and the phase connection the possible
phase shifting amounts to an integer multiple of /6 = 30 .

50

3. Transformers
Zt

I1

I2

ideal

U1

ejn/6

U2

kN1 N2
Figure 3.12. Single phase transformer model taking the three phase
connections into account.

Both of these points influence the turns ratio. The effect in amplitude is
incorporated by the multiplication of the regular turns ratio with the factor
k. According to the reasoning above the following holds
1
Dy-configuration: k =
3

Yd-configuration: k = 3

(3.29a)
(3.29b)

We incorporate the phase shift into the model by adding a pure phase shifting
element to the ideal transformer in figure 3.9. The phase shift may add up
to an integer multiple of /6 = 30 . Accordingly, we multiply the turns ratio
by a phasor with magnitude 1 and phase n/6
ejn/6
where n Z (see figure 3.12). We remember that n occurs as well in the
labelling of vector groups, namely as a number after the labelling of the connections (e.g., Ydn). The norm of the transformation ratio is not influenced
due to
|ejn/6 | = 1
(3.30)
However,the resulting turns ratio contains also a phase shift and is therefore
described by a complex number. Finally, we obtain
k

N1 jn/6
e
=cC
N2

(3.31)

Figure 3.13 shows a single phase equivalent network of this model. The ideal
transformer and the phase shifting element are merged here to one element.
In a real power transmission grid many connections exist via transformers. The different voltage levels, generators and loads are connected via transformers. If transformers operate in parallel, the vector group connections
need to be observed in special.

51

3.2. Three-Phase Transformer

Zt

U1

I2

I1

U2

Figure 3.13. Single phase transformer model with complex turns ratio
modelling a three phase transformer.

52

3. Transformers

4
Per Unit System
This chapter deals with the description with per unit values. This is nothing
else than relating a size to a predefined base unit and then describing it as a
multiple of this unit. This description is common in power engineering and
can be helpful e.g. when dealing with certain quantities in electric grids.

4.1

Reason for the Calculation with Per Unit Values

Normally, physical quantities are described as a product of a number and a


unit, e.g, U = 400 kV. These units are normally SI units and are uniquely
and globally defined. This has normally significant advantages, but sometimes other methods of describing quantities can be practical.
On the other hand, one can choose a description where the base is not
defined by an independent base system, e.g. by the SI system, but the base
units are selected so that certain other advantages are achieved. This is done
when one uses the so called per unit or p.u. system, which will be elaborated
in the following. Thereby, the value of a quantity is given as a multiple of a
predefined base value, i.e. the size is related to this base value. This value is
dimensionless and is referred to as the per unit (p.u.) value of this quantity.
In this system, the various quantities are given as
actual value
base value
Choosing the base value appropriately, the p.u. values turn out to be of great
practical value.
If, for example, the voltage in a node is related to the nominal voltage
of the grid, e.g. 400 kV the value u = 0.93 p.u. is more useful at first glance
than U = 372.03 kV. One notices instantly that the voltage lies 7% below
nominal value.
Systems with pieces of equipment with different ratings, e.g, transformers
with different nominal currents, are often more manageable and easy to
compare when describing different values in p.u. rather than by specifying
their absolute values. Describing, for example, the current values of two
transformers in each case related to the maximal allowed operating current
with i1 = 0.98 p.u. and i2 = 0.35 p.u., one recognizes instantly that the first
transformer is operating close to its loading limit whereas the seconde one
is far off its limit.
A further advantage of using the p.u. system occurs in power system
calculations with computers. By choosing the base values appropriately, a
quantity in p.u. =

53

54

4. Per Unit System

description of the p.u. values within the same range can be achieved although
the original systems were of different ratings (e.g. choose 100 MVA = 1 p.u.
for the calculation of large grids and 1 MVA = 1 p.u. for smaller grids). This
aspect can be advantageous when using numerical calculation methods.
Another advantage of these systems is essential for the description of
electric machines and transformers. For example, the leakage reactance of a
transformer in p.u., with the transformer ratings as base units, see below,
corresponds to its relative short circuit voltage, which usually is in the order
of 0.05 - 0.20 p.u., whereas its value in ohm can vary over a very wide range
depending on size and voltage level.

4.2

Example of Use of P.U. Values

We will now represent the values of a simple circuit in the p.u. system. For
1
this, we consider the example in figure 4.1, where XL = L and XC = C
.
For the series connection of the elements R, jXL and jXC we obtain
I

jXL

jXC

Figure 4.1. RLC series circuit.

U = RI + jXL I jXC I

(4.1)

By introducing a base voltage UB , by which the equation above is divided,


one obtains
U
RI jXL I jXC I
=
+

(4.2)
UB
UB
UB
UB
where u is the voltage related to the base voltage UB . u is a dimensionless
value and is described in p.u.. The base voltage UB is often chosen close to
the nominal voltage of the system. The index B, stands for Basis, and
denotes the nominal value.
Furthermore, a base power SB will be introduced. The choice of SB is
arbitrary because two base values of the quantities in the relation
u=

SB = UB IB =

UB2
ZB

(4.3)

can be chosen arbitrarily. After introducing the related power


s=

S
SB

(4.4)

55

4.2. Example of Use of P.U. Values

and voltage in equation (4.2) the base values for current and impedance are
defined by the following relations:
SB
UB
U2
UB
ZB =
= B
IB
SB
IB =

Introducing the p.u. current

(4.5)
(4.6)

I
IB

(4.7)

RIB
jXL IB
jXC IB
i+
i
i
UB
UB
UB

(4.8)

i=
equation (4.2) yields
u=

The base value of the nominal impedance ZB is the base for all impedances
within the system, real and imaginary and complex ones. The p.u. values of
the impedances are thus:
R
ZB
XL
xL =
ZB
XC
xC =
ZB
r=

(4.9a)
(4.9b)
(4.9c)

Now, all p.u. values can be placed into the original voltage equation (4.1)
and we obtain the following in p.u.:
u = ri + jxL i jxC i

(4.10)

Instead of the equation above one could also specify


U = ZI

(4.11)

for the circuit depicted in figure 4.1 where the overall impedance of the
circuit is
Z = R + jXL jXC
(4.12)
with the overall impedance in p.u.
z = r + jxL jxC =

Z
ZB

(4.13)

equation (4.11) is now, in p.u.,


u = zi

(4.14)

It is seen that this equation has the same form and structure as eq. (4.1).

56

4.3

4. Per Unit System

How to Choose Base Voltages for Interconnected


Circuit Elements

Based on the preceding sections the following statements are valid for a
linear system consisting of circuit elements.
As only one base power shall exist within the whole grid, it is useful
to chose the base power SB according to the most often occurring power
values. These are mostly described as three-phase powers. Typical values in
high voltage nets are 1 MVA for medium or 100 MVA for higher voltage
levels.
In the case of base voltages, it is common to define a phase to phase
base voltage for all nodes within each part of the net located between two
transformers (i.e. at the same voltage level). This grid base voltage is often chosen in order to enable the most suitable interpretation of the p.u.
node voltages, i.e. to reflect the operational status of the system. Often, the
nominal voltage is chosen as base voltage.
To illustrate these fundamentals of the p.u. calculation an example regarding the conversion of / kV/ MVA-Modell into the p.u. system is given
below.
5.5

1
2:1

5.5

1.375 : 1

120 kV

1 kA
Load

Figure 4.2. Example with -, kV- and kA values.

Given is the system shown in figure 4.2. The voltage of the source on
the left hand side is 120 kV. All impedances as well as all transformer turns
ratios are known. The load on the right hand side consumes a current of 1
kA. To be determined is the overall p.u. model and the p.u. voltages of this
system with the base voltages of 120 kV at node 1, 50 kV at nodes 2 and 3
and 40 kV at node 4.
Figure 4.3 shows the solution of the example in SI units. This can be
achieved by means of the standard methods of network theory.
The scheme above is depicted as p.u. model in figure 4.4. A base power
of SB = 100 MVA has been chosen. Based on the given base voltages of each
node all voltages, currents and impedances are translated into p.u. values.
Of interest is the description of the transformer turns ratio in p.u. Thereby,
the base value can be set on e.g. the nominal turns ratio of the transformer.

4.3. How to Choose Base Voltages for Interconnected Circuit Elements 57


5.5

5.5

2:1

1.375 : 1

120 kV 118 kV

59 kV

55 kV

40 kV

37 kV

1 kA

Load

Figure 4.3. Solution of the example in , kV und kA values.

This yields a turns ratio of 1 p.u. in our example, if primary and secondary
voltages amount to 1 p.u.:
cB =

U1B
120 kV
=
= 2.4 : 1
U2B
50 kV

(4.15)

The actual turns ratio of the transformer results from the actual connected
voltages
118 kV
c=
= 2.0 : 1
(4.16)
59 kV
Now we can relate the actual turns ratio of the transformer c to the base
value cB which yields the turns ratio c in p.u.:
c =

c
2.0
=
= 0.833 : 1 p.u.
cB
2.4

(4.17)

Analogously, the turns ratio of the second transformer can be related to the
base value.
0.038 p.u.

1
0.833 : 1 p.u.

4 0.188 p.u. 5

0.22 p.u. 3
1.10 : 1 p.u.

0.4 p.u.

1.0 p.u.

Load

Figure 4.4. Example with p.u. values.

The resulting node voltages are presented in Figure 4.5, given in p.u.
values.
From this description, one can easily check the percentage of the corresponding voltage drop or deviation from the given rated voltage level, respectively. The reasons of clarity and comparability as mentioned in section
4.1 are thus well illustrated.

58

4. Per Unit System

1.1 : 1 p.u.

0.188 p.u.

0.93 p.u.

1.00 p.u.

1.10 p.u.

0.98 p.u.

0.833 : 1 p.u.
1.0 p.u.

0.22 p.u.

1.18 p.u.

0.038 p.u.

0.4 p.u.
Load

Figure 4.5. Solution of the example in p.u. values.

This simplification enabled by the p.u. system is also applicable to the


conversion of transformer impedances from primary to secondary side. According to equation (4.15) the normalized value of the turns ratio is defined
as
U1B
cB =
U2B
When a primary transformer reactance Z 1 is given, its per unit value is
Z1
Z1
z1 =
= 2
(4.18)
Z1B
U1B /SB
If converting this impedance according to equation (3.15) to the secondary
side, its normalized value results in
z2 =

Z 1 /c2B
Z1
Z1
2 /S = c2 U 2 /S = U 2 /S = z 1
U2B
B
B
B
B 2B
1B

(4.19)

One notifies the same value on both sides of the transformer impedances in
p.u. as long the base transmission ratio cB is given according to (4.15).
The series impedance of real transformers (refer to the simplified model
in section 3.1.3) is about |z t | 0.05 . . . 0.15 p.u.

4.4

Conversion between Different p.u. Systems

When converting between two different p.u. systems one has to recalculate
alt into a new system with
a given p.u. base system having the base values XiB
new
the values XiB . Practically, one needs to perform this for example when
the p.u. data of a generator needs to be adapted to the base values of the
connected power grid
Generally, the way via the base system neutral / kV/ MVA-system leads
always to the goal. Namely, one transforms all quantities given in p.u. system
A into / kV/ MVA sizes and chooses then an arbitrary new unified base
system B, valid for all sizes. However, this way is relatively tedious.
It is easier to transform the quantities directly to the new base values.
Therefore, one needs to multiply the actual p.u. value with the ratio of the
old to the new base value.

4.4. Conversion between Different p.u. Systems

59

Example: Conversion of impedance The actual value of an impedance Z


in must remain equal when changing to a new base value:
old
new
Z = zold ZB
= z new ZB

(4.20)

Hence, the new p.u. value is given by


z new = zold

old
ZB
new
ZB

(4.21)

As the base value of the impedance originates from the base values of voltage
and power, according to equation (4.6), the following holds hold
z

new

old

=z

UBold
UBnew

2

new
SB
old
SB

(4.22)

60

4. Per Unit System

5
Lines
In this chapter a very important system component in power transmission
systems the power line will be modelled and analyzed. First, we examine distributed line parameters of various line configurations and discuss the
electromagnetic fields close to overhead lines. Next, we introduce a line model
which allows the derivation of a equation of the line in steady state conditions. Finally, we investigate the line models for different line lengths with
different approximations.

5.1
5.1.1

Introduction
Characteristics of Power Lines

When the current is flowing through the metallic conductors electric and
magnetic fields are created and the electric power is carried by these fields
(Poyntings vector S = E H).
In power systems the primary task of a line is the secure transport of
powers between different locations in the grid. Utilized are either overhead
lines or cables. They differ in their construction and the operating conditions
and therefore their usage is different. A considerable difference is the cost:
at equal voltage levels the costs for a cable can be up to about 10 till 20
times higher than those of an overhead line. Therefore the construction of
overhead lines is preferred from an economical point of view.
The geometry of overhead lines or cables is characterized by that it
is not varying along one direction, namely along the direction of the line.1
Therefore, a parallel configuration of the conductors is considered. This is an
approximation which considerably simplifies the mathematical calculation
without compromising the accuracy of the results.
Power lines can be considered as a homogenous construction possibly including multiple single conductors. Thereby, the properties are characterized
by distributed parameters (e.g., inductance per line length in mH/km). A line
can therefore a priori not be described by lumped parameters as known from
the theory of circuits. Simplified, a line section can though be modelled by a
-configuration using lumped circuit elements, i.e. impedances. Depending
on the type of analysis, this simplification is valid up to a certain line length
1

A certain sag between two towers evolves at overhead lines due the weight of the
conductors. This can be incorporated approximatively by an assumption of an average
distance to earth, refer to equation (5.51).

61

PSfrag
62

5. Lines
Generator

PG

PL

UG

Um

Load

UL

Figure 5.1. A simplified energy distribution system

as will be discussed and further elaborated in this chapter.

5.1.2

Reasons for High Transmission Voltages

Figure 5.1 shows a simplified direct current transmission system. Here, a


load is fed with electric power by a generator through a power line. As the
line is resistive, a voltage drop will occur along the line: The voltage at the
generator UG is higher than the voltage at the load UL . Um denotes the
voltage in the middle of the line.
The average power transferred over the line is thus
Ptrans = Um I

(5.1)

If R is the overall line resistance, this yields the following transmission losses
P :
2

P = RI = R

Ptrans
Um

2

(5.2)

Relating the power losses to the transmitted power we obtain the the relative
power loss as
P
Ptrans

=R

Ptrans
Um 2

(5.3)

Equation (5.3) illustrates the inverse proportionality of the transmission


losses to the square of the line voltage.
To keep the losses low, energy transmission lines are operated at voltages
which should theoretically be as high as possible. However, increasing the
line voltage implies an increase of the insulation costs of the components
(e.g. transformers). As shown in figure 5.2 decreasing losses imply increasing
equipment costs. The optimal voltage level results at minimal overall costs
as indicated qualitatively in figure 5.2.

63

5.2. Line Parameters


K
Kg
Kv

Ka

Kmin

Uopt

Figure 5.2. Costs K vs. transmission voltage U . The overall costs Kg


are obtained by the loss costs Kv and the equipment costs Ka . The
economically optimal transmission voltage Uopt results at minimal costs
Kmin .

R l

L l

I
E

G l

C l

l
Figure 5.3. Left hand side: magnetic (H) and electric (E) field of a
high voltage overhead line; right hand side: line model of a short section
of a line with concentrated parameters.

5.2

Line Parameters

For each line different characteristic parameteters, so called line parameters,


are obtained depending on the geometry of the line configuration. These
parameters are normally measured per line length and denoted by primed
quantities, i.e. as X . Thereby one distinguishes between
resistance per unit length R in /km
inductance per unit length L in H/km
conductance per unit length G in S/km
capacitance per unit length C in F/km
The physical meaning of these quantities is illustrated in figure 5.3. The series resistance of the line originates from the ohmic resistance of the metallic
conductor. The series inductance and the shunt capacitance result by the effects of magnetic and electric fields induced by the voltage and current. The

64

5. Lines

Figure 5.4. Bundled conductors: duplex-, triplex- and quadruplex conductors.

shunt conductance is due to the leakage currents occurring in the insulators


and in the insulating medium (e.g., air, SF6 ). The shunt currents are normally very small wherewith the shunt conductance can often be neglected.
All the other line parameters are used for modelling of high voltage power
lines.
The aim is now to find formulae for the above mentioned parameters for
three-phase lines which are applicable to arbitrary forms and configurations
of conductors.

5.2.1

Power Line Conductors

A power circuit consists basically of two main components: conductors,


whose task is to carry the current and an insulating system guaranteeing
the necessary insulation of the current carrying parts between themselves
(phases) as well as to ground. In the case of overhead lines, the insulation
consists of insulators, e.g. made of porcelain, glass, silicone, located at the
towers and of air along the conductors. Depending on the operating voltage specific minimal distances need to be respected in order to obtain the
demanded voltage strength of the insulating path. These distances (and
therefore the operating voltage of the power line) is of importance for the
design of the line towers.
Each phase of a transmission line consists of one or multiple conductors.
If several conductors are used for one phase, these are referred to as bundle conductors. Depending on the number of conductors per phase they are
called duplex, triplex, quadruplex etc. conductors (see Fig. 5.4). These bundle conductors result in a lower series impedance of the line and in reduced
electric field strength at the conductor surface. Hence, the corona discharges
and therefore the shunt conductance of the line are reduced.
Nowadays, almost only aluminium or aluminium alloys are used as conductor material. In order to obtain a high mechanical strength a design
with a steel core with additional aluminium strands is often used (see Figure 5.5). These conductors are referred to as ACSR (Aluminium Conductor
Steel Reinforced) or Stalu.
As the specific resistance of steel is much higher than the one of aluminium (F e /Al 4 . . . 10), the major part of the current flows in the
aluminium wires. Additionally, the skin effect yields a displacement of the

65

5.2. Line Parameters


Aluminium

steel

Figure 5.5. Cross section of a ACSR/Stalu conductor.

current to the outer conductor regions.


As already mentioned, alloys of aluminium, magnesium and silicon are
utilized as material for the conductor lines. By this, a comparable mechanical
strength as ACSR is obtained but with much lower weight. However, the cost
is considerably higher.

5.2.2

Inductance of a Power Line

Inductance of an Infinitely Long Straight Conductor


In order to derive the inductance of a multiple phase conductor system we
start with the examination of an infinitely long straight conductor. This
infinitely long and straight-lined conductor is an approximation of a long
power line and simplifies the calculation of the parameters. This approximation does not correspond to the model of a real transmission line, but only
represents the first step in the modelling process. As we will see later, the
required equations can be found by simple considerations and superposition.
We start with the derivation of the inductance of an infinitely long line
with radius r where the current i is distributed uniformly (constant current
density) inside the conductor. The field lines of the H field are concentric
circles (see figure 5.6). For points outside the conductor, x > r, it holds
Ha (x) =

i
2x

(5.4)

Inside the conductor (x r) the total current inside 2 in figure 5.6 is


i(x) =

x2
i
r 2

(5.5)

Therewith, the magnetic field strength inside the conductor is obtained as


Hi (x) =

x
i
2r 2

(5.6)

66

5. Lines
y
dl
Wire

x
r

Figure 5.6. Cross section of a conductor of a power line.

The overall inductance is summed up by the fields inside and outside the
conductor.
Outer Inductance For determining the part of the inductance originating
by the field outside the conductor we consider the magnetic flux from the
conductor surface within a radius R. Each field line surrounds the conductor
exactly once, therefore the flux linkage2 corresponds to the flux within the
rectangle with radius r until R according to figure 5.7. The length of the
rectangle is assumed as one length unit (1 l.u.). The flux linkage is calculated
as
Z
Z R
Z R
i
0 i R
a =
Ba (x)da =
Ba (x)dx = 0
dx =
ln
(5.7)
2
r
A
r
r 2x
where we have used da = 1 dx and Ba (x) = 0 Ha (x) as differential area.
The relation between the flux linkage a and inductance la is
a = la i

(5.8)

By this, we obtain the outer inductance per length unit as


la =

0 R
ln
2
r

(5.9)

2
The concept of the flux linkage is normally applied for coils. Thereby, the flux linkage
is the part of the flux which intersects all windings of the coil(s). The infinitely long
conductor can be imagined as a coil with one single winding which is closed at infinity
( and + are assumed to be the same point).

67

5.2. Line Parameters

Infinitely long
wire with
radius r

area da

1l.u
.

Ba

0
i

dx

Figure 5.7. Flux through rectangle R r, 1.

Inner Inductance For the derivation of the inner inductance we first compute the magnetic energy within one length unit of the conductor:
Z
Z r
1
1
x 2
0 i2
2
Wm = 0
Hi dV = 0
i
2xdx
=
(5.10)
2
2
2r 2
16
V
0
represents the relative permeability of the conducting material. In the case
of copper or aluminium one can assume 1.
Denoting the part of the inductance resulting of the flux linkage within
the conductor by li , we obtain
Wm =

1 2
li i
2

(5.11)

The inner part of the inductance per length unit is then


li =

0
8

(5.12)

From equation (5.12) one can conclude that the inductance is independent
of the conductor radius r.
If we now sum the outer and inner inductances of equations (5.9) and
(5.12), we obtain the total inductance of the conductor per length unit.


0
R
l = li + la =
+ ln
(5.13)
2 4
r
If inserting the value for 0 in this equation the inductance per km can be
computed:


R
4
l = 2 10
+ ln
H/km
(5.14)
4
r
By using that /4 = ln e/4 equation (5.14) can be written as
l = 2 104 ln

R
H/km
r

(5.15)

68

5. Lines
i2
in
Phase n

Phase 2

R1
i1
Phase 1
i3
Phase 3
Figure 5.8. System with n phases.

where
r = re/4

(5.16)
r

For 1, i.e. for copper or aluminium conductors, 0.78 r holds.


Considering the expression for the inductance in equation (5.15) one notices that this value becomes infinite for an infinite integration area, namely
R . This does not correspond to a real physical system. This derived
model for a single wire is incomplete, because in a real physical system a
return current must flow somewhere within a finite distance R of the conductor.
In the following paragraphs it is shown that the physical model becomes
consistent when considering this circumstance.
Multi Phase Systems
To correct the unphysical result for the single conductor we now examine a
system with n conductors as shown in figure 5.8. In the following we denote
a single conductor as a phase, according to figure 5.8. Thus, we do not
consider a bundle conductor consisting of n single conductors in which the
phase current, in , is split, but we consider n phases each of them carrying
a (phase) current.3 Each phase can, however, consist of a bundle conductor
with multiple single conductors (see figure 5.11) which together carry the
phase current. We will come back to this situation later.
First of all, we are interested in the total flux linkage of phase 1 from
the origin until radius R1 . In addition to the flux of phase 1 the flux of
phases 2, 3, . . . , n need to be considered. The flux linkage of phase 1 has
been introduced in the previous section and yields
11 = l1 i1 =
3

i1 0 R1
ln
2
r

(5.17)

In the case of three-phase systems the phases 1, 2, . . . , k, . . . , n are denoted with R, S


and T.

69

5.2. Line Parameters

d1k

d1k

Phase 1

R1
R1 d1k

Phase k

a
2

b
3

Figure 5.9. Contribution of phase k {2, 3, . . . , n} to the flux linkage of phase 1.

We consider now the part originating from phase k of the flux linkage
with phase 1 (see Figure 5.9). As shown in figure 5.9 the magnetic field lines
are represented by ik concentric circles with the center in the conductor
of phase k. The field line 1 does not include the conductor phase 1 and
yields consequently no contribution to the flux linkage. The field line 3 is
interlinked with phase 1, 5 as well, though outside of radius R1 . The field
lines 2 and 4 represent the borders between phases 1 and k where the flux
contribute to the flux linkage of phase 1 if integrating up to R1 . Hence, only
those field lines yield a contribution to the flux linkage which intersect the
axis A between the points a and b.
The flux linkage between phase 1 and phase k can be computed with
equation (5.7):
ik 0 R1 d1k
1k =
ln
(5.18)
2
d1k
The overall flux linkage of phase 1 can now be found by superposition of the
single parts of all phases and results in


n
X
0
R1
R1 d12
R1 d1n
1 =
1k =
+ . . . + in ln
i1 ln + i2 ln
2
r1
d12
d1n
k=1
(5.19)
This expression can be split into two parts:


0
1
1
1
i1 ln + i2 ln
+ . . . + in ln
+
1 =
2
r1
d12
d1n
n
0 X
+
ik ln (R1 d1k )
(5.20)
2
k=1
|
{z
}
A

where we have set d11 = 0. The term A can be simplified by applying a


few mathematical manipulations. We add a null term B, reformulate the

70

5. Lines

equation and then we will let R1 :


A=

n
X
k=1

ik ln (R1 d1k )

n
X
k=1

ik ln R1 =
{z

B=0

n
X
k=1

dik
ik ln 1
R1

(5.21)

P
Since nk=1 ik = 0 it holds that B = 0. To calculate the overall flux linkage,
we let R1 and obtain


n
X
dik
A (R1 ) = lim
ik ln 1
=0
(5.22)
R1
R1
k=1

Finally, the total flux linkage of phase 1 results thus according to (5.20) in


0
1
1
1
i1 ln + i2 ln
1 =
+ + in ln
(5.23)
2
r1
d12
d1n

The flux linkage of phase 1 therefore depends on all currents i1 , i2 , . . . , in .


The above equation can be reformulated in
1 = l11 i1 + l12 i2 + + l1n in

(5.24)

where the inductances l1k are dependent only of the geometry of the conductor configuration. The overall flux linkage for phase k results in


0
1
1
1
k =
i1 ln
+ + ik ln + + in ln
(5.25)
2
dk1
rk
dkn
with obvious notation.
In the following two sections two important special cases regarding multiphase systems are considered, namely the two- and three-phase system.
System with Two Phases
Consider a system according to the left part of figure 5.10 consisting of two
identical conductors (phases) with radius r with a separation distance D.
With equation (5.25) and the assumption i1 + i2 = 0 we obtain


0
1
1
1 =
i1 ln + i2 ln
2
r
D


1
1
0
=
i1 ln i1 ln
2
r
D
0
D
=
i1 ln
(5.26)
2
r
The same holds for 2 . Therewith, the inductances of both conductors per
length unit are the same and equal
0 D
ln
(5.27)
2 r
The overall inductance is composed of the self and mutual inductance.
l1 = l2 =

71

5.2. Line Parameters

2r

2r

Figure 5.10. Systems with two and three phases.

System with Three Phases


Consider a system according to the right part of figure 5.10 consisting of
three identical conductors (phases) with radius r and in each with a separation distance of D, i.e. the three phases are placed at the corner points of
an equilateral triangle with side length D. Here as well, we obtain the flux
linkage with equation (5.25) and i1 + i2 + i3 = 0


0
1
1
1
i1 ln + i2 ln + i3 ln
1 =
2
r
D
D


0
1
1
=
i1 ln i1 ln
2
r
D
D
0
i1 ln
(5.28)
=
2
r
The same holds for phases 2 and 3. Therefore, the inductance of phases 1, 2
and 3 are
0 D
l1 = l2 = l3 =
ln
(5.29)
2 r
The comment stated above regarding self- and mutual inductance holds here
too. Due to symmetry, i.e. the equidistant configuration of the phases and
the requirement i1 + i2 + i3 = 0, we obtain the simple expression for the
overall inductance in equation (5.29).
Bundle Conductors
In section 5.2.1 it was mentioned that at higher voltage levels, typically over
200 kV, each phase consists of multiple single conductors (see figure 5.11).
For the calculation of the inductance the procedure introduced above can
be applied. Thereby, expressions similar to those for the case of one single
conductor per phase are obtained. It is outside the scope of this lecture to
elaborate this derivation.
Actually, the same formulae can be applied for the calculation of the
inductances with a minor modification. Instead of the conductor radius r
and instead of the distance d
a so-called equivalent radius req
ik in figure

72

5. Lines

2r

Figure 5.11. Symmetric configuration of the bundled conductor of one phase.

5.9 a geometric average value of the distance between the single conductors
of the phases should be used. It can be shown that the equivalent radius
of a bundled phase with n symmetrically placed single conductors can be
calculated by the following relation:

= nRn1 r
(5.30)
req
The notation corresponds to figure 5.11. It is easy to show that the equivalent
radius is enlarged by splitting one phase in a bundled conductor consisting
of multiple single phases having the same cross-section area. Accordingly,
the inductance is reduced.
Of course, the equivalent radius is also computable for the physical conductor radius r (see figure 5.7):

n
req = nRn1 r
(5.31)
The distance between the phases is much larger that the distances of the
single conductors within the phase bundle. Hence, the distance between the
midpoint of the conductor bundle instead of the geometric average of the
single distances can be inserted.
Transposition of Phases
In the previous subsection a general relation for the flux linkage in a multiphase system has been derived. Thereby we assumed that the single phases
are placed in a symmetric configuration, e.g. placed at the corner points of
an equilateral triangle. However, often the three phases are not arranged
exactly symmetrically.
Quite often all three phases are placed horizontally or vertically. That
is why the single phases are linked with different fluxes. If these differences
are too large, countermeasures need to be taken. A general method consists of continuously exchanging the position of the phases (see figure 5.12),
wherewith a transposition of the conductor is obtained.
We consider a transposed line with a phase line configuration according
to figure 5.13 with following assumptions:

73

5.2. Line Parameters

Transpositions
Position
1

Section 1

Section 2

repetition

Section 3

Figure 5.12. Transposed three-phase line.

2
d12

d23

1
d13

Figure 5.13. General configuration of three-phase lines.

74

5. Lines
1. Each phase (R, S And T) takes each position (1, 2 and 3) over equal
length sections of the line.
2. iR + iS + iT = 0

We calculate now the average flux linkage per length unit over three line
sections for phase R:

1  (1)
(2)
(3)
R =
R + R + R
(5.32)
3
(i)

R represents the flux linkage per length unit within section i of the line,
with i = 1, 2, 3. Inserting equation (5.25) in equation (5.32), we obtain
1 0 
1
1
1
R =
+ iT ln
iR ln + iS ln
3 2
r
d12
d13
1
1
1
+ iR ln + iS ln
+ iT ln
r
d23
d12

1
1
1
+ iR ln + iS ln
+ iT ln
(5.33)
r
d13
d23

Each line in equation (5.33) corresponds to an application of equation (5.25)


on phase R in the particular positions 1, 2 and 3. Now the terms can be
added:


0
1
1
1
R =
iR ln + iS ln
+ iT ln
(5.34)
2
r
dm
dm
with

dm =

p
3

d12 d23 d31

(5.35)

The distance dm is the geometric mean of the distances between the three
phases.
Applying the second assumption stated above, i.e. iR + iS + iT = 0,
equation (5.34) can be further simplified to
R =

0
dm
iR ln
2
r

(5.36)

and we obtain the average inductance as


lR =

0 dm
ln
2
r

(5.37)

Due to the transposition of the line the same average inductance per unit
length for all three phases is obtained:
lR = lS = lT

(5.38)

The average inductance of a transposed line is the same for all phases and
amounts per length unit to
L =

0 dm
ln
2
r

(5.39)

75

5.2. Line Parameters

Contribution from equation (5.41)


++

q2
q3

qk

(k)

++
+

++
++

(k)

++
++

qn

++
++

q1
Contribution = 0
++
+

Phase 1

q4

Figure 5.14. Integration of the electric field strength.

In practice, transposition is often done only in the (existing) substations


whereby the individual sections are not exactly equidistant anymore. Thus,
only a limited balancing of the different configurations is obtained. Normally,
the phase conductors are transposed only at very high voltage levels outside
substations.

5.2.3

Capacitance of a Power Line

The derivation of the shunt capacitance of a conductor can be obtained


analogously to the series inductance. Consider a system as depicted in figure
5.14. The conductor k carries the charge qk per length unit. The electric field
of the assumed infinitely long conductor is radial and homogenous and its
absolute value is
qk
Ek =
(5.40)
20 R(k)
The voltage between the points and can be obtained by integration
of the electric field along an arbitrary curve from point till point . Of
course, the integral along concentric circles, which represent equipotential
lines of the radial homogenous electric field, is zero. Of importance is only
the distance from the conductor k.
The integral of the field strength over the radial path between the concentric circles, on which the points and lie, yields the voltage between
the two points (or between the equipotential lines, respectively), which is
originating from the charge of conductor k:
(k)
u

(k)

(k)

(k)

qk
qk
R
dR =
ln (k)
20 R
20 R

(5.41)

76

5. Lines
(k)

where R is the radial distance from conductor k to point , analogously


for .
For a configuration of multiple conductors the overall voltage is obtained
by superposition of the single voltages. For n conductors this yields
u =

n
(i)
1 X
R
qi ln (i)
20
R
i=1

(5.42)

We now assume that the charges of all conductors add to zero:


q1 + q2 + + qn = 0

(5.43)

Similar to the deviation of the magnetic flux linkage, we assume the radius
(i)
to be R , this means we relate the voltage of the conductor to a
infinitely far-off point. With equation (5.43) we obtain
u =

n
1 X
1
qi ln (i)
20
R
i=1

(5.44)

Of interest is the voltage at a conductors surface, for example of conductor


k. Assuming rk as conductor radius, the voltage at the conductors surface
is computed to


1
1
1
1
uk =
q1 ln
+ + qk ln
+ + qn ln
(5.45)
20
dk1
rk
dkn
If the phase conductors consist of multiple single conductors (bundle), rk
needs to be replaced with rk,eq in order to calculate the voltage (see equation
5.31). In the case of bundled conductors, the electric field strength is reduced
at the surface of the conductor. Therewith, the leakage currents over the air
and therewith the corona sound emissions of the conductor are reduced as
well.
Equation (5.45) can also be written in matrix format:
u = Fq

(5.46)

The matrix F contains the potential coefficients. The capacitances can be


found from the inverse relation:
q = Cu

(5.47)

with C = F1 . Figure 5.15 shows the equivalent conductor capacitances


which occur in a three-phase power line.
The per phase and per unit length capacitances of a transposed threephase line results in
20
C =
(5.48)
dm
ln
req

77

5.2. Line Parameters


2

C12

C23

C13

C10

C20

C30

Figure 5.15. Capacitances of a three-phase power line.

This equation generally yields quite accurate results. In the derivation we


assumed the reference potential to be located at infinity and to be zero.
Although earth is not an ideal conductor without resistance the potential
at the earths surface can be assumed to be zero.4 This fact needs to be
considered mainly when the distances between the phases are of the same
order of magnitude as the distances to earth.
In order to represent the earths surface as ground potential mirror
charges can be used. They carry the same amount of charge but with opposite sign as the phase conductors. They are placed at equal distance to
the surface but at the other side of the surface, namely below the earths
surface. Geometrically, a reflected image of the situation above earth is obtained below earth, only the charge signs are exchanged (see figure 5.16).
For a three-phase system, an equivalent system with six phase conductors
is obtained: three above earth (real conductors) and three below earth with
opposite equally sized charges (mirror conductors).
For this system, the same derivation as for the real three-phase system
can be done. The capacitance per phase and per length unit for a transposed
three-phase line results in

=
Cm

20


2H dm
ln
A req

(5.49)

with dm and req as above. H is the geometric mean of the phase line distances
from the earths surface (see figure 5.16)
p
H = 3 H1 H2 H3
(5.50)

As depicted in figure 5.17, a sag appears between two towers due to the
weight of the conductor lines. By this, the distance between conductor and
4

Already a very small charge density at the earths surface is sufficient to obtain this.

78

5. Lines
2
d12

d23

1
d13

H2

H1

H3
A23

Earth

A13
A12
3
1

2
Figure 5.16. General configuration of a three-phase power line with
mirror conductors.
h
Hmin

Figure 5.17. Sag of the conductor lines between two towers.

the earth surface is not constant. This sag can be incorporated by defining
a reduced distance
1
H = Hmin + h
(5.51)
3
in the equation stated above.
The geometric mean of the distances between the phase conductors and
the mirror conductors is (see figure 5.16)
p
A = 3 A12 A23 A13
(5.52)

If the height of the phase lines is large compared with the distances between
the phases, 2H/A 1, holds. By this, the equations (5.48) and (5.49) yield
the same result for the capacitances per length unit.

5.2.4

Ohmic Losses

Real power lines are not lossless. Thus, an additional series resistance and
a shunt conductance need to be incorporated besides the inductance and
capacitance per unit length.

79

5.2. Line Parameters


Series Resistance

In the case of direct current, the ohmic resistance of a conductor can relatively simply be calculated by its cross-section area, length and specific
conductivity. But when the conductor carries an alternating current, the
skin effect needs to be incorporated when calculating the ohmic resistance.5
This effect implicates that the current is not distributed homogenously over
the conductor cross-section but is pushed to the outer boundary of the conductor, i.e. the current density is increasing outwards. Thereby, the effective
conductor cross-section is decreased compared with the direct current leading conductor. In a 50 Hz system, this increase typically amounts to a few
percent.
When having a configuration as shown in figure 5.5 the ohmic alternating current resistance approximately corresponds to the direct current
resistance of the aluminium conductor. Thereby, the increasing effect of the
resistance due to the current displacement approximately balances out the
current part which flows in the steel core. More exact values can be gained
by measurements of the nominal operation frequency, i.e. 50 or 60 Hz, and
the operational temperature. These values are provided by the conductor
manufacturer.
The ohmic series resistance per unit length R gives rise to losses in the
conductor, where part of the losses are dissipated to air by convection and
radiation. The allowable current capability is limited by a maximal heating
of the conductor; long term conductor temperatures above 80 C can lead to
a weakening for commonly used conductor materials (aluminium and steel).
For a conductor temperature of , the resistance can be calculated by6

R = R20
(1 + ( 20 C))

(5.53)

denotes the per unit length resistance at 20 C. It is computed


where R20
by means of the specific resistance of the conductor material at 20 C 20
(in mm2 /m) and the effective cross-section area A (in mm2 ):

R20
=

20
A

(5.54)

Shunt Conductance
The per unit conductance G represents the corona discharges at the surface of conductors and leakage current losses across the insulators in the
5

Apart from the skin effect, other effects which influence the ohmic resistance occur
within the conductor line: The proximity effect describes the current displacement in the
conductor by fields of adjacent conductors. The spirality effect describes the increase of
the ohmic resistance due to the design of the individual conductors.
6
For aluminium and copper a temperature coefficient of = 0.004 K1 is normally
used.

80

5. Lines

case of overhead lines. In the case of cables, the dielectric losses within the
solid isolators contribute to G . The operational losses are almost identical
with the open-circuit losses, which for overhead lines are weather dependent.
For example, on a foggy and humid day, one can clearly hear the cracking
noise due to partial discharges or corona discharges. However, the corona
phenomenon occurs less on a dry summer day, whereby the corona losses
decrease considerably. For overhead lines only approximate values based on
measurements can be gained for G . On the other hand, exact test measurements are possible for cables.

5.2.5

Conductor Parameters for Cables and Overhead Lines

The conductor parameters of cables and overhead lines differ significantly.


The reasons for this are:
1. In a cable the distances between the phase conductors are considerably
lower than for overhead lines. The same holds for the distances between
the phase conductors and earth.
2. In a cable the conductors are mostly surrounded by other metallic
materials such as shielding and sheaths.
3. For overhead lines, air serves as insulating material between the phase
conductors. For cables, other materials are used such as paper, oil, or
SF6 .
From equations (5.39) and (5.48) two important differences for the distributed parameters for overhead lines and cables result:
1. The per unit length series inductance of overhead lines is generally
higher than the one of cables.
2. The per unit length shunt capacitance is normally significantly higher
for cables than for overhead lines.
In the case of overhead lines the ohmic losses in a conductor is dissipated to
the surrounding air by convection or radiation. In the case of cables the heat
needs to be transported through the dielectric insulation, so convection can
not take place. In order not to exceed the maximal operational temperature
of the cable the heating and therefore the ohmic losses need to be kept
low. Thus, cables have at same nominal power lower series resistances than
overhead lines.
In Tables 5.1 and 5.2 some typical values for line parameters are given.

81

5.3. Electromagnetic Fields of Overhead Lines

Table 5.1. Typical conductor parameters for overhead lines at 50 Hz [10].

nominal voltage in kV
R in /km
XL = L in /km
YC = C in S/km

230
0.050
0.407
2.764

345
0.037
0.306
3.765

500
0.028
0.271
4.333

765
0.012
0.274
4.148

Table 5.2. Typical conductor parameters for cables at 50 Hz [10].

nominal voltage in kV
R in /km
XL = L in /km
YC = C in S/km

5.3

115
0.059
0.252
192.0

230
0.028
0.282
204.7

500
0.013
0.205
80.4

Electromagnetic Fields for Three-phase High Voltage Overhead Lines

In relation with the increased use of mobile phones during the last decades
more concerns about possible harmful effects of electro-magnetic fields to
human beings have been expressed. Overhead lines cause both electric and
magnetic fields. An important difference between the fields generated by
antennas used for mobile communication and high voltage lines is the operating frequency. The fields of high voltage lines have a low frequency of
50 Hz, or 60 Hz, whereas the fields from antennas lie in the GHz domain.
Moreover, the very aim of antennas is the generation and radiation of fields
whereas this is only an undesirable side effect in the case of high voltage
lines.
To gain a feeling how strong the electromagnetic fields of a high voltage
line close to the ground are, a high voltage line with the dimensions given
in figure 5.18 is considered in this section.
Since we consider a symmetric three-phase system, the voltage in the
single conductors are in each case phase-shifted by 120 , and the voltages
and currents in the conductors vary with the power frequency of 50 Hz.
These voltage and current variations influence the electromagnetic fields.
The field strength is dependent on the considered time instant, i.e. the fields
are also alternating.

5.3.1

Electric Field

The electric field of a high voltage line depends on the voltage, the form of the
tower and the configuration of the conductor lines. The analytic calculation
of the electric field is very difficult because the earth impedance needs to be
incorporated. Hence, we abandon the derivation and present only the results

82

5. Lines

12 m

12 m

20 m
y

Figure 5.18. Dimensions of the considered high voltage line.


umax

uS (t)

uR (t)

uT (t)

umax

t1

t2

t3

Figure 5.19. Phase voltages in all three phases.

of a simulation for the configuration given in figure 5.18.


In figure 5.19 the behavior of the voltages of all three phases is shown.
Regarding figure 5.18 uR (t) corresponds to the voltage in the left conductor,
uS (t) to the one of the middle conductor and uT (t) to the one in the right
conductor. In the following simulations three time instants t1 , t2 and t3 are
considered, as indicated in figure 5.19.
The resulting values of the electric fields are strongly dependent on the
voltage. The electric field for phase to phase voltages of 110 kV, 220 kV
and 400 kV (root mean square values) are considered. The electric field
disappears exactly on the ground because the potential is zero there. Related
to the coordinate system in figure 5.18, the values of the fields are shown in
figure 5.20 at a distance of 2 m above the ground, i.e. y = 2 m, up to 80 m
in positive and negative x direction.
Only at time instants where the voltage of the middle conductor receives
its maximal value or passes through zero, i.e when the voltage values of the
left and right conductors are equal, a symmetric pattern is obtained. At

83

5.3. Electromagnetic Fields of Overhead Lines

E (kV/m)
2, 5
400 kV
2, 0
1, 5

220 kV

1, 0
110 kV
0, 5
0
80

60 40

20

20

40

60

80x (m)

20

40

60

80x (m)

20

40

60

80x (m)

E (kV/m)
2, 5
2, 0
1, 5
1, 0
400 kV
0, 5

220 kV
110 kV

0
80

60 40

20

E (kV/m)
2, 5
400 kV

2, 0
1, 5

220 kV
1, 0
110 kV

0, 5
0
80

60 40

20

Figure 5.20. Electric field at time instant t1 (top), t2 (middle) and t3


(bottom) 2 m above earth (conductor 20 m above earth).

84

5. Lines

the same time, the field assumes on average the maximal value (t1 ) or the
minimal value (t2 ) at these time instants. The time instant t3 is an example
where the field is not symmetric.
As a comparison, the natural static air field obtains a field strength of
around 0.1 kV/m. In connection with a thunderstorm, values up to 20 kV/m
are possible. Also in a house we are constantly exposed to electric fields from
different appliances. Within a distance of 30 cm field strengths up to 0.5
kV/m can be measured around electric devices. The human skin serves as a
safety shield and is sufficient for not enabling electric fields of transmission
lines to enter. In buildings, the electric field coming from outside and varying
with 50 Hz is reduced to at least one tenth. The values of electric fields which
are generated by high voltage lines at 2 m above ground (figure 5.20) are
thus not much stronger than those we are exposed to from other sources.

5.3.2

Magnetic Field

In contrast with the electric field the magnetic field can be computed analytically quite easily as the earth influences the magnetic field only insignificantly. Using the law of Biot-Savart, a straight current-carrying conductor
creates a magnetic field with the value
0 I
(5.55)
2R
Thereby R denotes the distance between the point at which the field is computed and the conductor, and the direction of the field is always tangential
to the circles around the conductor.
This can now be applied for the calculation of the magnetic field for
a three-phase line. For each conductor the value and the direction of the
magnetic field is determined. Then, the three field vectors resulting of the
three phases are summed up and the overall value of the field is computed.
As the field of a straight conductor only depends on the distance between
the conductor and the point at which the field is computed (5.55), we can
use a two-dimensional model to derive the formulas. For a single conductor
the relations are given in figure 5.21.
The value of the magnetic field at location (xB |h) results with (5.55) in
|B| =

|B| = B =

0 I
(xB xw )2 + (h yw )2

(5.56)

Thereby we assume that I flows in the positive z direction, i.e. out from the
paper. In opposite direction I becomes negative.
In order to sum up the magnetic fields of all three phases, B is split into
x and y coordinates:
Bx = B cos
By = B sin

(5.57)
(5.58)

85

5.3. Electromagnetic Fields of Overhead Lines


y
I

(xw |yw )

R
B

.
(xB |h)
x

Figure 5.21. Magnetic field of a straight line.

With
sin =
cos =
this results in
Bx =
By =

p
p

xB xw

(xB xw )2 + (h yw )2
yw h
(xB xw )2 + (h yw )2

0 I(yw h)
2((xB xw )2 + (h yw )2 )
0 I(xB xw )
2((xB xw )2 + (h yw )2 )

(5.59)
(5.60)

(5.61)
(5.62)

The calculation of the overall field is carried out by summing up the


single fields in the x and y directions. This field varies in time just as the
electric field does. In figure 5.22 the root mean square (rms) values of the
overall field for the configuration in figure 5.18 and for the phase to phase rms
currents of 1000 A, 700 A and 500 A at a distance of 2 m above the ground
are
given. (This means that the phase currents are these values divided by
3.) It is assumed that the conductor height is constantly 20 m. Considering
a line sag would result in higher values for the magnetic field. The rms is
calculated by the instantaneous values as follows:
s
Z
1 T 2
Beff =
B (t)dt
(5.63)
T 0
The calculation of the magnetic field at large distances of the conductor
can be simplified by some approximations. As the direction of the magnetic
field is always tangential to the circle around the conductor, we can conclude

86

5. Lines

Beff (T)
8

1000 A

700 A

500 A
2

0
80

60

40

20

20

40

60

80x(m)

Figure 5.22. RMS values of the magnetic field at different locations


at a height 2 m above ground (conductor 20 m above earth).

a
B
x

Figure 5.23. Magnetic field at large distances.

5.3. Electromagnetic Fields of Overhead Lines

87

that the overall field is more or less perpendicular (normal) to earth at large
distances (figure 5.23).
We assume that the distance between P and the conductor corresponds
to the distance in x direction, for example the distance for the conductor on
the left hand side amounts to x + a. In phasor notation the magnetic field
at P results in


0
IR
I
I
B=
+ S+ T
(5.64)
2 x + a
x
xa
With the approximation

1
1
1+

for small this results in




0 (x a)
1
(x + a)
B=
IR + IS +
I T 7, a x
2
x2
x
x2

(5.65)

(5.66)

From the condition that the sum of the currents is equal to zero at each
time instant
IR + IS + IT = 0
(5.67)
the simple formula
B=

0 a

(I I R ) , a x
2 x2 T

(5.68)

is derived, and therewith also the absolute value of the magnetic field at
large distances

aIph
3Iph a
0
= 3.5 107 2 , a x
(5.69)
B=

2
x2
x
Thereby Iph denotes the phase current in A; with x and a in meter a flux
density B in T results. For the calculation of the field directly below an
overhead line, (5.69) cannot be employed, because the assumption that the
distance between the conductors is much smaller that the distance to the
point of computation does not hold anymore. Regarding (5.69), the field
would be infinite at x = 0, which obviously is not correct. However, the
formula for the calculation of the magnetic field at large distances is useful
and shows the quadratic decrease of the field with increasing distance.
Although magnetic fields are hardly damped by the human skin, they
induce only small currents within the body. Electric household devices, such
as mixers or electric irons generate magnetic fields with amounts up to 10 T
at a distance of 30 cm. For causing heart problems for humans alternating
field strengths at a frequency of 50 Hz and 1 Tesla are needed. This is ten
thousand times more than the field strength below a high voltage line.
7

1
x+a

1
x

1
a
1+ x

1
x

a
x

(xa)
x2

88

5. Lines

Although hundreds of research studies by scientists all over the world


have concluded that no harm or damage to the human being due to high
voltage lines can be proved, intense research is undertaken. In many European countries anyhow maximum limit values concerning magnetic field
strengths have been set by authorities which are not allowed to be exceeded
by high voltage power lines.

5.4

Line Model and Solution of the Wave Equation

The model of a power line forms an important prerequisite for planning and
operation of electric power systems. Thereby we assume the system to be in
steady state, which is actually never exactly fulfilled in practice, but quite
often a sufficiently good approximation. The always present load variations
keep the system always in a dynamic state. However, for many tasks it is
enough to make a snapshot of the operating state and to neglect the short
term transients. This approach is justifiable because the load variations are
fairly small over a short period of time. More significant load changes extend
over a longer time period (a number of ten minutes) at normal operation.
Therefore, a model with constant loads describes the instantaneous operation state with a sufficient accuracy. The normal operation state can thus
be described by a system of nonlinear algebraic equations which are derived
in this chapter.
Transient processes, such as a large load step or the outage of a power
station, cannot be modelled by these equations. Only after the dynamic
transients have decayed the equations derived in this section can be used for
the calculation of the stationary operation state.
How long this computed stationary operation state is valid depends on
the load variations at normal operation. At fast load changes (e.g. in the
surrounding of day load peaks) the stationary solution needs to be redetermined more often than at times with only few load changes (e.g. during
night hours).
An important precondition for the validity of the circuit-based model to
be developed is that the distances between the conductors are small compared with the characteristic electro-magnetic wave length. Thereby, it is
ensured that the field between the conductors approximately coincides with
the static field. This holds both for electric fields and for magnetic fields. In
addition, we can assume that these fields are practically decoupled.
Thereby it is possible to associate the charges and the surrounding flux of
the conductors with the voltage and the current by usage of capacitances and
inductances. Capacitances and inductances are considered to be distributed
and given per length unit. For also ensuring the validity of the capacitor and
inductor model in axial line direction, a short piece of the line is considered

5.4. Line Model and Solution of the Wave Equation

89

(length element x, see figure 5.24). For each line element it holds
Q = C U
= L I
where Q denotes the charge and the flux per line unit length. Both quantities can be determined by measurements of a finite line element. As these
quantities are proportional to the length of the line element, the capacitance
per unit length C and the inductance per unit length L can be defined.
The behavior of a power line is determined by its length and the per unit
length parameters R , L , G and C . The actual design of overhead lines
is often very dependent on local conditions. This means, the characteristic
values cannot be determined by lab measurements, as it is the case for devices
like generators and transformers. This holds generally also for cables, which
are delivered to the site. The per unit length inductance and capacitance of
power cables depend on site-specific factors such as installation type (e.g.,
configuration and distance of phase conductors), connection and grounding
of the cable sheaths or shields as well as on the electric characteristics of the
surrounding soil.

5.4.1

Equivalent Circuit Diagram of a Line Element

The line model is derived starting from a single line element. By usage of
the per unit length parameters derived in section 5.2 an infinitesimally small
line element, as shown in figure 5.248 , is considered. In this equivalent circuit
diagram of the line element, voltages and currents are shown, with which
the general differential equation of the homogenous line can be derived.
In order to analyze the phenomena on a power line one would need to use
the general theory of electromagnetic phenomena, i.e. Maxwells field theory.
As certain assumptions can be made regarding the geometric configuration of
the line, several model simplifications can be assumed concerning the circuit
model. The per unit length capacitance and inductance are assumed constant
along the power line. Thereby, there exists a field pattern perpendicular to
the conductors, which corresponds to the static field lines. The propagation
velocity is dependent on the dielectric and the magnetic materials of the
power line, see below.
The propagation of waves on power lines is an important characteristic
of the power line theory. A line section, attached to a source at one end,
behaves differently than a lumped circuit element. Depending on the length
of the line, approximations can be introduced, and these are often used in
practice. They are discussed more in detail later on.
8
For visualization of the line model often a series connection of inductances and capacitances is shown, which is a useful representation for the analysis. However, such a
circuit does not reflect the wave propagation properties for a power line. Hence, we do not
consider this model in more detail.

90

5. Lines
i(x, t)

u(x, t)

R x

L x

G x

i(x + x, t)

C x

u(x + x, t)

x + x

Figure 5.24. Equivalent circuit diagram of a line element x of a


lossy homogenous power line.

The theory is simplified when assuming the power line as lossless (R =


= 0). When studying many phenomena this approximation can be used
without any significant loss of the validity of the results.
The propagation of waves on conductors is always connected with an
energy transport. The basic application of the Poynting vector on a field
pattern between two conductors shows that the energy flows in the axial
direction of the line.The energy transport is thus carried out in the field
between the conductors. Because of the predefined geometry and the simplifying assumption we can determine the current and voltage at a specific
point on the conductor from a given field. This is the reason why the field
approach is used only in rare cases. Mostly, one uses current and voltage
when describing power lines at steady state, or at low-frequency transients.
In the electric power transmission power and energy are of central importance. As operation frequency in steady state, the values 16 23 , 50 and
60 Hz are most often used.9 Sometimes, more than one single frequency is
needed to model the system accurately: many transients contain components
from a wide frequency spectrum, and these need to be considered in order
to give the correct time response. The general power line model needs thus
to incorporate both descriptions in time domain and the frequency domain.
However, the analysis of high-frequent transients goes beyond the scope of
this lecture.
A fundamental assumption for the application of the power line model
to be derived in the following is the linearity of the field relations. This
means that the permeability and dielectric constant are constant in the
whole considered space. Therewith, the law of superimposition holds, which
G

In Switzerland, the Swiss federal railways (SBB (Schweizerische Bundesbahnen)) use


a frequency of 16 23 Hz for their power network. This frequency is also used by many other
countries for the power supply of railway systems.

91

5.4. Line Model and Solution of the Wave Equation


plays an important role when passing from frequency to time domain.

5.4.2

The Differential Equation for the Power Line The Telegraph Equation

In the following we assume that each small element of the line can be modelled by the equivalent circuit diagram depicted in figure 5.24.
Voltage u and current i are functions of the location x along the line and
of the time t:
u = u(x, t)
i = i(x, t)
When moving along the line from position x a small step x, voltage and
current at position x + x are according to Kirchhoffs laws
i (x, t)
(5.71a)
t
u (x + x, t)
i (x + x, t) = i (x, t) G x u (x + x, t) C x
t
(5.71b)

u (x + x, t) = u (x, t) R x i (x, t) L x

These equations can be rewritten as


u (x + x, t) u (x, t)
i (x, t)
= R i (x, t) L
x
t
u (x + x, t)
i (x + x, t) i (x, t)

= G u (x + x, t) C
x
t

(5.72a)
(5.72b)

Assuming that the line element is infinitesimal, i.e.


x 0

(5.73)

we obtain the following differential equations using equations (5.72a) and


(5.72b)10 :


u


= R +L
i
(5.74a)
x
t


i


= G +C
u
(5.74b)
x
t
The equations (5.74a) and (5.74b) constitute a system of partial linear first
order differential equations with constant coefficients and are referred to as
the Differential Equations of Power Lines.
10

We have here assumed that the derivatives of the functions i and u exist, and these
functions are hence continuous.

92

5. Lines

We obtain, by using equations (5.74a) and (5.74b), equations with only


variable (u or i), by differentiating the first equation with respect to x and
the second with respect to t (or the other way round) and then eliminating
i or u. Thus we obtain the Telegraph equation (J.C. Maxwell, 1860):
2

2u


u
u
=
R
G
u
+
R
C
+
L
G
+
L
C
x2
t
t2
2
2

i


i
i
+
L
=
R
G
i
+
R
C
+
L
G
C
x2
t
t2

(5.75a)
(5.75b)

Both partial differential equations (5.75a) and (5.75b) are valid for arbitrary
time dependencies of voltage and current, hence also for transients. They are
valid for inhomogeneous conductors too when incorporating the locational
dependency of the linear electric constants (e.g., R = R (x)).
In the next section well solve these differential equations for sinusoidal
excitations, in which case the equations often are referred to as the Power
Line Equations.

5.4.3

Wave Equation and its Solution

The differential equations (5.75a) and (5.75b) are solved here for the special
case of pure sinusoidal excitation. This excitation form is very important in
physics as transfer functions can be defined for this form of excitation. The
sinusoidal function is namely, besides a constant, the only function which
appears in the same form at the output of a linear physical network as it
enters at the input, albeit with changed amplitude and phase.11
If the line is excited by an arbitrary signal form, the signal can be decomposed into its frequency components by means of the Fourier transformation.
For each component of the signal, the solution of the differential equation is
determined and one obtains the solution for the given excitation signal as a
sum of the single solutions (superposition theorem).12
The bases for the following considerations are the Telegraph equations
(5.75a) and (5.75b). These are linear partial differential equations of second
order with constant coefficients and are identical for voltage and current.
Hence, solutions (for u(x, t) and i(x, t)) are both of the same form.
We now introduce phasors for u(x, t) and i(x, t):



u(x, t) = 2 U (x) ejt



i(x, t) = 2 I (x) ejt

where U and I denote the voltage and current phasors and = 2f is the
angular frequency. These phasors are now inserted into equations (5.74a)
11

This is due to the fact that the function sin (t), its derivative and each linear combination of it also have a sinusoidal form.
12
Of course, this is valid only for a linear system.

93

5.4. Line Model and Solution of the Wave Equation


and (5.74b), respectively, and one gets:

dU
= R + jL I
dx

dI
= G + jC U
dx

(5.77a)
(5.77b)

The phasor I can now be eliminated from equations (5.77a) and (5.77b):


d2 U
= R + jL G + jC U
2
dx

(5.78)

In contrast with equation (5.75a) the time is not explicitly appearing anymore in equation (5.78). By introducing the phasors, the partial differential
equation (5.78) has been converted into an ordinary linear differential equation with respect to x of second order with constant coefficients.
By elimination of U from equations (5.77a) and (5.77b) we find an equation for the current analogous to (5.78):


d2 I
= R + jL G + jC I
2
dx

(5.79)

The equations (5.78) and (5.79) are the Wave Equations of the lossy line
and are normally written in the following form:
d2 U
= 2U
dx2
d2 I
= 2I
dx2 p
with = (R + jL ) (G + jC ) = + j

(5.80b)

U (x) = U a + U b = U a0 ex + U b0 ex

(5.81a)

(5.80a)

(5.80c)

Thereby denotes the complex propagation constant with the dimension


1/length.
The solutions of the equations (5.80a) and (5.80b) are of the following
form:

I(x) = I a + I b = I a0 e

+ I b0 e

(5.81b)

If (5.81a) is differentiated with respect to x and inserted together with equation (5.80c) into equation (5.77a), we find a relation between the line voltages
and currents:
s

1
dU
G + jC
x
x
I (x) =

=
U
e

U
e
(5.82)
a0
b0
R + jL dx
R + jL

94

5. Lines

By comparing the coefficients of equations (5.82) and (5.81b), we obtain


the characteristic impedance, or surge impedance, as quotient of voltage and
current:
s
U a0
U b0
R + jL
ZW =
=
=
(5.83)
G + jC
I a0
I b0
The dimension of the surge impedance is .
By means of equation (5.83), equation (5.82) can be expressed as follows:
I(x) =

5.4.4

1
(U a0 ex U b0 ex )
ZW

(5.84)

Interpretation of the Wave Propagation

The solutions in equations (5.81a) and (5.81b) will now be studied in more
detail. Eq. (5.81a) written out as time function reads
u(x, t) =

n
o
2 U a0 ex ejt + U b0 ex ejt
|
{z
} |
{z
}
a)

(5.85)

b)

We can split up the wave propagation constant to real and imaginary part:
=
with

(R + jL ) (G + jC ) = + j

(5.86)

. . . Damping Constant in Np/m (Neper/Meter)


. . . Phase Constant in rad/m

Often is given in dB/m, with the conversion done as follows

= 8.686
dB/m
Np/m

(5.87)

where the logarithmic definition of dB has been used.


The part a) in equation (5.85) corresponds to a forward, in the positive
x-direction, travelling damped voltage wave. Part b) corresponds to one
backward, thus in the negative x-direction travelling, damped voltage wave.
Which is the velocity vp of the wave in the positive x-direction, i.e.
how fast does the zero crossing of the wave travelling in positive x direction
t x = 2 move? The reference point x = x0 +vp t moves with the phase
velocity vp . The phase (t (x0 + vp t)) remains then constant, when the
phase velocity amounts to
vp =

(5.88)

95

5.4. Line Model and Solution of the Wave Equation


|Ua |
x

|Ub |
x

Ua

Ub

Ua

|U |

Ub
U

Figure 5.25. Forward and backward travelling voltage wave.

The phase velocity of the wave in the other direction, part b) in the
equation (5.85), is by a similar reasoning found to be
vp =

(5.89)

The phase velocity in overhead power lines reaches almost the speed of light
in vacuum or air, see below.
The left picture in figure 5.25 shows the forward travelling damped wave,
the middle picture in figure 5.25 shows the backward travelling damped
wave, and the right picture in figure 5.25 shows the sum of the forward
and the backward travelling waves. These pictures show however only theoretically the behaviour of the line voltage: In real voltage waves the wave
length is approx. 6000 km, thus the shadowed part in picture 5.25 represents
already a line length of 750 km!

5.4.5

Inclusion of Boundary Conditions

The constants U a0 and U b0 in the voltage equation (5.81a) result from the
boundary conditions at the beginning and the end of the power line. These
constants can thus be determined if these conditions are known, as will be
shown in the following. We will study two cases that are common in practice,
i.e. that the voltage and current are known at the beginning and at the end
of the line, respectively. Often the beginning of the line is referred to as the
sending end and the end of the line to as the receiving end.

96

5. Lines

Specification of Current and Voltage at the Beginning of the Line


If current and voltage at the beginning of a line are given, equation (5.81a)
yields
U (x = 0) = U 1 = U a0 + U b0

(5.90)

And from equation (5.82) results


I (x = 0) = I 1 =

1
(U a0 U b0 )
ZW

(5.91)

Hence, the constants are given by


U 1 + ZW I1
2
U 1 ZW I1
=
2

U a0 =

(5.92a)

U b0

(5.92b)

Consequently we obtain the following expressions for voltage and current at


location x along the line:
1
1
(U 1 + Z W I 1 ) ex + (U 1 Z W I 1 ) ex
2
2


U1
1 U1
1
x
I(x) =
+ I1 e

I 1 ex
2 ZW
2 ZW

U (x) =

(5.93a)
(5.93b)

A mathematically more elegant form results when combining both parts


according to U 1 and Z W I 1 , respectively:
U (x) = U 1

ex + ex
ex ex
ZW I1
2
2

(5.94)

and analogously for I(x). Using the cosh and sinh functions we get


U (x) = U 1 cosh x Z W I 1 sinh x
(5.95a)


U1
I(x) = I 1 cosh x
sinh x
(5.95b)
ZW
Specification of Current and Voltage at the End of the Line
If current and voltage at the line ending are specified, equation (5.81a) yields:
U (x = l) = U 2 = U a0 el + U b0 el

(5.96)

and from equation (5.81b) for the current at the end of the line
I (x = l) = I 2 =


1 
U a0 el U b0 el
ZW

(5.97)

5.4. Line Model and Solution of the Wave Equation

97

Then, the constants are given by


U 2 + Z W I 2 l
e
2
U 2 Z W I 2 l
=
e
2

U a0 =

(5.98a)

U b0

(5.98b)

These are now inserted into the general wave equation, eqs. (5.81a) (5.81b),
and we obtain
1
1
U (x) = (U 2 + Z W I 2 ) e(lx) + (U 2 Z W I 2 ) e(lx)
(5.99a)
2
2


1 U2
1 U2
I(x) =
+ I 2 e(lx)
I 2 e(lx)
(5.99b)
2 ZW
2 ZW
This is the general form of the line equations with current and voltage specified at the line end. By introducing the hyperbolic functions we obtain
analogously:


(5.100a)
U (x) = U 2 cosh (l x) + Z W I 2 sinh (l x)


U2
I(x) = I 2 cosh (l x) +
sinh (l x)
(5.100b)
ZW

The comparison of these expressions with (5.95a) and (5.95b) shows,


that these latter equations can be obtained by substituting x by l x and
inserting the current and voltage values at the line end.

5.4.6

Wave Equation of Lossless Line

A special and important case is the lossless line. As the inductive reactance
dominates for high voltage power lines (For voltages from around 130 kV
R
L . 0.1) and the losses due to the shunt conductance are very low, the
power transmission line can be modelled as lossless as a simplification, i.e.
R = G = 0

(5.101)

This neglect of R and G influences of course the line equations as shown


in below.
The following equations serve as a basis for our calculations:
ex + ex
ex ex
ZW I1
2
2
ex + ex
U 1 ex ex
I(x) = I 1

2
ZW
2

U (x) = U 1

(5.102a)
(5.102b)

Thereby, the values of current and voltage at the beginning of the line are assumed to be known. The propagation constant becomes purely imaginary
in this special case:

= j L C = j
(5.103)

98

5. Lines

The surge impedance is purely real in the special case of a lossless line and
is given by
ZW =

L
C

(5.104)

For the sine and cosine functions the following relations hold:
sin (z) =

ejz ejz
2j

cos (z) =

ejz + ejz
2

(5.105)

When applying the relations from equation (5.105) for equations (5.102a)
and (5.102b), we obtain the wave equations for the lossless line.
U (x) = U 1 cos (x) jZW I 1 sin (x)
U1
sin (x)
I(x) = I 1 cos (x) j
ZW

(5.106a)
(5.106b)

Analogously, we obtain the wave equations for the lossless line, when current
and voltage at the end of the line are known.
U (x) = U 2 cos ( (l x)) + jZW I 2 sin ( (l x))
U2
sin ( (l x))
I(x) = I 2 cos ( (l x)) + j
ZW

(5.107a)
(5.107b)

For the lossless line we can now calculate the phase constant, , by using
the equations derived in section 5.2 for
 
0
dm
L =
ln

2
req
20
C =  
dm
ln
req

We thus obtain
v
 
u
dm
u
ln

req

= L C = t0 0   0 0 =
c
m
ln dreq

(5.108)

where c denotes the light velocity in vacuum or air. Therewith, the phase
constant for the lossless line is only dependent on the operation frequency

and amounts at 50 Hz and = 1 to about = 1006 km . From eq. (5.88) the


wave propagation velocity is calculated to be c.

99

5.5. Line Models

5.5

Line Models

In practice, one does not always want to use the relatively complicated wave
equation; the exact propagation of current and voltage along a line is not
always of interest. Often, one is only interested in determining e.g. the voltage drop along a line or in calculating the total reactive power demand. For
these kinds of examinations some simplifications of the wave equation can
be made without losing any accuracy of the results. Simplified line models
are thus obtained which are suitable for different problems. In the following
a few line models and their application areas are discussed.

5.5.1

General Two-Port Model of a Line

A Two-Port model (T or equivalent) with lumped parameters can be


derived from the line model with distributed parameters (per unit length
values). This description of a line or a line section can simplify further examinations. For example, in many cases only the voltage at the line beginning
and ending is of interest, but not the distribution of the voltage along the
line. One can therefore use models with lumped parameters. Such a model
can be derived as follows:
1. First, the model of a element (as depicted in figure 5.26) is set up.
2. Then, the coefficients are compared with those of the wave equation
and the values of the elements of the equivalent circuit diagram can
be computed.
Figure 5.26 shows a model of a line section. For the derivation it is
useful to split the overall shunt admittance Y q in half and connect it to
each side of the circuit. The relations between current and voltage at the
element in figure 5.26 are

!
!
Yq
1
+
Z
Z
l
l
U1
2
U2


= Y 
(5.109)
Yq
Yq
q
I1
I2
2
+
Z
1
+
Z
l 2
l 2
2
|
{z
}
A1

Zl

I1

U1

Yq
2

I2

Yq
2

U2

Figure 5.26. equivalent circuit diagram of a homogenous power line.

100

5. Lines

The same relations can be expressed with the wave equations (5.100a) and
(5.100b):

!
!
!
cosh l
Z W sinh l
U1
U2
=
(5.110)


1
I1
I2
cosh l
Z W sinh l
|
{z
}
A2

where U 1 = U (x = 0) and I 1 = I(x = 0). When comparing the coefficients


of matrices A1 and A2 , the following expressions can be obtained for the
elements of the equivalent circuit diagram:

Z l = Z W sinh l
(5.111a)

 
cosh l 1
l
Yq
1
=
tanh
=
(5.111b)
2
ZW
2
Z W sinh l
These elements correspond to the exact relations between current and voltage and which result from the solutions of the wave equation for x = 0 and
x = l.
For the case |l| 1 the expressions for Z l and Y q can be simplified:

(5.112a)
Z l = Z W sinh l Z W l = Z l
 

l
Yq
1
1 l
Y l
=
tanh

=
(5.112b)
2
ZW
2
ZW 2
2

or by use of the parameters R , L , G and C as



Z l = Z l = R + jX l
Yq
Y
(G + jB )
=
l=
l
2
2
2

(5.113a)
(5.113b)

Of importance for the modelling of a line with the above simplifications is


how well |l| 1 is fulfilled. The larger this product is the less exact is the
representation of the relations of the wave equation by the model with
the lumped parameters calculated as in eqs. (5.112) or (5.113). Applying
the equivalent circuit diagram, one needs to be aware that the results differ
more and more from the exact current and voltage relations when increasing
the line length l. Depending on the kind of line, the approximation is valid
up to the following line lengths 13 :
until about 300 km for overhead lines
until around 100 km for cables14
13

This length is here always related to the operation frequency of 50 Hz. At higher
frequencies the limits become shorter according to the shorter wave length.
14
We will later see that AC cables can be utilized anyway only up to a length of around
60 km. The reason for this is the high capacitive current which results from the high
capacitance of the cable.

101

5.5. Line Models

In practical power flow calculations, longer lines are often split into several
short pieces, which are then each modelled as elements and connected
together.

5.5.2

Lossless Line

A further simplification consists in neglecting the ohmic elements of the line,


i.e. we set
R = G = 0
For this special case the surge impedance becomes purely real and the wave
propagation constant purely imaginary, i.e. ZW R and = j. The corresponding simplification in the equivalent circuit diagram is achieved by
neglecting the ohmic elements in the equations (5.113):
Z l = Z l = jX l
Yq
Y
B
=
l=j l
2
2
2

(5.114a)
(5.114b)

For overhead lines sometimes only the resistive part of the shunt conductance is neglected:
G = 0
The elements of the equivalent circuit diagram are with this simplification
Z l = Z l = R l + jX l

Yq
Y
=
l=j l
2
2
2

(5.115a)
(5.115b)

This model yields a very good approximation for lines with nominal voltage
starting at around 130 kV.

5.5.3

Additional Simplifications

Models for overhead lines can be further simplified. For example, the overall
shunt impedance can be neglected under certain circumstances and the line
can be modelled as a pure series impedance. Normally, lines are divided into
three categories according to their line length: 15
Short Lines (until 100 km): Lines with a length of around 100 km normally have a very low shunt capacitance and the ohmic part of the
shunt conductance is low as well. These short lines can be modelled as
a single series impedance Z = R l + jL l.
Medium Length Lines (100 until 300 km): For line lengths in the range
of 100 until ca. 300 km the simplified model (5.113) can be used
without great loss of accuracy.
15

The lengths given here serve as approximate benchmarks and not as strict rules.

102

5. Lines

Long lines (over 300 km): For line lengths over about 300 km the effects of the distributed parameters are significant. Thus, these lines
should be modelled as series connections of shorter line elements or
the lumped parameters should be computed with the exact relations
according to (5.111).

5.5.4

Comparison of Different Line Models

In the following the results of different line models are discussed on the basis
of an example. Given are the surge impedance and the propagation constant
of a 230 kV line:
Z W = 382.2 j 16.5

= + j = 0.0001 Np/km + j 0.0011 rad/km


We calculate now the voltage at the end of the open line. At open-circuit
operation the current at the line end is I 2 = 0, the voltage at the line
beginning as |U 1 | = 1.0 p.u. |U 2 | is computed in the following way:
a) With the exact equation (5.100a):
U (x = 0) = U 1 = U 2 cosh l

b) With the wave equation for a lossless line (5.107a):


U(x = 0) = U 1 = U 2 cos (l)
c) With the equivalent circuit for lines of medium lengths. From the
propagation constant and the wave resistance we can compute the
elements of the equivalent circuit according to (5.112):
Zl ZW l

und

Yq
1 l

2
ZW 2

With (5.109) the relation between the voltages results in


!


(l)2
Z lY q
U1 = U2 1 +
= U2 1
2
2
d) With the line modelled as a pure series impedance. For a series impedance no voltage drop or increase occurs at open-circuit operation, the
output voltage equals the input voltage.
Table 5.3 and figure 5.27 show the results for the different models above.
One can notice how the results of the different models differ more with
increasing line length. For lines with a length until around 50 km, all four

103

5.5. Line Models


Table 5.3. Absolute value of the voltage at the end of a 230 kV overhead line running open-ended (no-load) for various line lengths l, computed with different line models. Values in p.u.

l in km
50
100
300
500

|l|
0.0552
0.1105
0.3314
0.5523

a)
1.0015
1.0060
1.0565
1.1710

b)
1.0015
1.0061
1.0570
1.1730

c)
1.0015
1.0060
1.0540
1.1503

d)
1.0000
1.0000
1.0000
1.0000

1.15
a)
b)
c)
d)

|U 2 | in p.u.

1.10

1.05

1.00

0.95

50

100

200

300

400

500

l in km
Figure 5.27. Absolute value of the no-load voltage at the line end,
computed with different models. a) exact b) with lossless wave equation, c) with model (5.112), d) with serial impedance.

models yield almost identical values for the output voltage. At longer lines,
the series impedance model d) differs considerably from the other models.
At a line length of 300 km the deviation between the results of the complete
wave equation a) and those of the model c) amounts to around 0.24%, at
500 km to 1.77%. Obviously, the lossless wave equation b) models the exact
conditions very accurately, the results almost coincide with the exact values:
At a line length of 500 km the difference is only 0.17%. By this reason, the
lossless model is used very often in practice for high voltage power lines
when voltage drops should be calculated. For loss calculations, this model is
of course of no value.

104

5. Lines

6
Power Transmission Fundamentals
In this chapter the steady state current and voltage characteristics as well as
the active and reactive flows of three-phase power lines will be described. After some basic considerations the operational behavior of different situations
will be analyzed.
In the discussion of the operational characteristics of power lines the
analysis will be restricted to one-phase equivalent systems, because for the
time being only symmetrical conditions will be considered. In the chapter 7
we are going to discuss how non-symmetrical systems could be described as
three decoupled one-phase equivalent systems.
The second very important assumption of the following analysis is that
the system is in steady state. For the calculations of dynamic phenomena
(e.g. transient oscillations after switching operations) other models must be
used.

6.1

Decoupled quantities

The power transmission over a line can be described in several ways. As


presented in figure 6.1, usually the quantities complex voltage, active and
reactive power are of interest at the beginning and at the end of the line.1
For a power transmission over a line eight real quantities are of importance:
U1 , 1 , U2 , 2 , P1 , Q1 , P2 , Q2
1

Instead of power the (complex) current could be used as well.

P2 + jQ2

P1 + jQ1

U1 1

Power line

Figure 6.1. Power transmission over a line

105

U2 2

106

6. Power Transmission Fundamentals

Not all of these quantities are independent of each other. The angles
of voltages U1 and U2 (1 and 2 ), for instance, are coupled through the
properties of the line (length, phase constant). We will later see that in all
formulae only the difference between 1 and 2 will occur. One of the angles
can then be chosen as reference (e.g. 2 = 0), thereby there are only seven
quantities of interest.
The power at the beginning and the end of the line are not independent of
each other, either. If e.g. active and reactive power are given at the beginning
of the line , the powers at the end of the line are a function of the line
parameters and the transmission voltage. Either U1 or U2 can be given. If
the complex voltages at the beginning and at the end of the line (three real
values) are given, this will uniquely determine the active and reactive powers
at the two ends of the line.
The complex current and voltage along the line are related according
to the two complex equations (5.95a) and (5.95b). Thus, the number of
independent variables is reduced by two complex or four real quantities.
The above discussion concerning the quantities describing the power
flows in a line can be summarized as:
number of quantities
8
a voltage angle as reference
1
2 complex = 4 real U /I-equations 4
independent real quantities
3
The number of the independent quantities in figure 6.1 is thus reduced
to three real quantities. This means that from eight real quantities, which
describe the power transmission, only three can be chosen freely and the
other will then be given by the power line equations. These three quantities
are referred to as independent (decoupled) quantities.
Not every combination of three independent quantities is meaningful
from a practical point of view. In power system analysis and operation the
following three combinations of independent quantities are most common:
U1 , 1 , U2 : The active and reactive powers are then given by the line
equations. This case occurs, e.g. when the power line is connecting two
large networks where the voltage and angle are fixed and independent
of the loadings in the systems. This is a typical situation for lines in a
meshed system with good voltage control.
U1 , P2 , Q2 (or P1 , Q1 , U2 ): If the voltage is given at one end of the
power line and the powers at the other end. This case corresponds to
when a distant load is fed via a power line and the voltage is controlled,
e.g. by a generator, at the sending end.
U1 , P1 , Q1 : In this case the voltage magnitude and the powers are
given at the same end of the line. This is e.g. the case when a power
station feeds a certain power over a line into a power grid.

107

6.2. Surge impedance loading

6.2

Surge impedance loading

By the surge impedance loading of a circuit one understands the power which
is transmitted when the power line is loaded with an impedance equal to its
surge impedance.

6.2.1

Surge impedance loading of a lossless power line

We regard a lossless circuit with R = G = 0, where an impedance equal to


its purely real surge impedance is connected. The surge impedance can be
calculated from the circuit parameters:
r
L
Z2 = ZW =
(6.1)
C
The power at the output of the circuit amounts to

PSIL = P2 =

|U 2 |2
|U 2 |2
=
Z2
ZW

(6.2)

and the current to


I2 =

U2
U2
=
Z2
ZW

(6.3)

If we now insert equation (6.3) into the equations for the lossless circuit
(5.107a) and/or (5.107b), we obtain the voltage and current at the beginning
of the circuit (x = 0):
U2
sin (l)
ZW
= U 2 (cos (l) + j sin (l)) = U2 ejl
1
U2
sin (l) +
cos (l)
I 1 = jU 2
ZW
ZW
U2
=
(cos (l) + j sin (l)) = I 2 ejl
ZW

U 1 = U 2 cos (l) + jZW

(6.4a)

(6.4b)

The results show that current and voltage are phase shifted over the line with
the same angle, l, while the amplitudes are unchanged. If one calculates
the quantities along the line (x 6= 0), then the phasors are phase shifted
proportionally to the distance from the circuit end. The amplitudes of the
2
Here we use the convention from chapter 2 : For the computation of the three-phase
quantities
P , Q and S the phase quantities of the current and voltage are multiplied with

3. In the following it is assumed that the current and voltage


values at the beginning

and end of the line are the phase quantities multiplied by 3.

108

6. Power Transmission Fundamentals

current and voltage along the circuit remain the same. The power at the
beginning of the circuit is the same as at the end of it:
P1 =

|U 1 |2
= P2
ZW

(6.5)

because no losses appear on the circuit. This special load case is called surge
impedance loading (SIL) PSIL .
PSIL =

|U |2
ZW

(6.6)

With this load one achieves in a sense optimum power transmission conditions. Current and voltage have constant amplitudes along the line and
the angular phase shift between the beginning and the end of the circuit
amounts to l. However, in practice the transmitted powers deviate most
often from the surge impedance loading.
At surge impedance loading the power line is in a state of reactive power
balance. This means that the series inductance consumes exactly the same
amount of reactive power as the shunt capacitances produces. The reactive
power produced in the shunt capacitances depends on the shunt capacitances
and the operating voltage. The reactive power consumed in the series inductance depends on the line current and on the value of the series inductance.
At a loading equal to the surge impedance loading the two reactive powers
are equal:
QC = QL U 2 C = I 2 L

U2
L
2
=
= ZW
I2
C

(6.7)

The surge impedance of overhead circuits lies within the range of approximately 200 to 400 . Dominant parameter is the series inductance,
its reactance is very high in comparison with the shunt reactance. The reactive power consumed in the series inductance QL exceeds already for a
relatively low current the reactive power generated in the shunt capacitance
QC . For this reason overhead circuits are usually operated above their surge
impedance loading.
Cables have relatively small surge impedances compared with overhead
circuits, approximately within the range 30 to 50 . The influence of the
shunt capacitance is dominating. Hence a higher surge impedance loading
for the same operating voltage U results. The surge impedance loading lies
most often above the thermal power limit of the cable. While overhead circuits can transmit powers several times higher than the surge impedance
loading, cables are always operated below their surge impedance loading.
The characteristics and features of cables are described more in detail in the
section 6.10.

109

6.2. Surge impedance loading

Here, we introduce the software PowerWorld (URL of PowerWorld web


site: http://www.powerworld.com). The PowerWorld Simulator is an interactive power systems simulation package designed to simulate high voltage
power systems operation. The software contains a power flow analysis package capable of solving systems with up to 100,000 buses. A demo version
of the PowerWorld Simulator for educational use, which can handle systems
with up to 12 buses, is freely available and can be downloaded from the PowerWorld web site. On the web site of this course, you can download some
example cases. The file 2 Bus Example contains a simple example in which
one can study the behavior of a power line.
The surge impedance loading can also be applied to a lossy power line.
The output and input powers are then, of course, not equal anymore. This
case will be examined in the next section.

6.2.2

Surge impedance loading of a lossy power line

In the case of a lossy power line the surge impedance is a complex quantity.
For following investigation an impedance equal to the surge impedance is
connected at the end of the line. The current at the end of the line is equal
to
I2 =

U2
U2
=
Z2
ZW

(6.8)

The complex power at the end of the circuit can now be calculated
S 2 = P2 + jQ2 = U 2 I 2 =

|U 2 |2
Z W

(6.9)

If the equation (6.8) is inserted in the circuit equations (5.100a) and (5.100b)
the voltage and current at the beginning of the line can be calculated


U2
U 1 = U 2 cosh l + Z W
sinh l
ZW


= U 2 cosh l + sinh l = U 2 el


U2
U2
I1 =
cosh l +
sinh l
ZW
ZW


U2
=
cosh l + sinh l = I 2 el
ZW

(6.10a)

(6.10b)

The complex power at the beginning of the circuit is


S 1 = P1 + jQ1 = U 1 I 1 = U 2

U 2 2l
e = S 2 e2l
Z W

(6.11)

The phase relationship between current and voltage remains constant also
in the lossy case over the entire circuit, but the amplitudes are not the same

110

6. Power Transmission Fundamentals

anymore. Active and reactive power increase toward the beginning of the
circuit, the power at the input is higher with a (real) factor e2l than at
the output. Also the amplitudes of the voltage and the current rise along
the circuit (when moving along the line from end to the beginning, i.e. from
point 2 to point 1) proportional to ex . The angle phase shift is still x. The
distance x is here measured from 2 (line end) towards 1 (line begin).

6.2.3

Typical values for overhead lines and cables

Typical values for the surge impedance and the surge impedance loading of
overhead power lines and power cables are given in the tables below.
Overhead power lines [13]:
nominal voltage in kV
ZW in
PSIL in MW

132
150
50

275
315
240

115
36.2
365

230
37.1
1426

380
295
490

Power cables [10]:


nominal voltage in kV
ZW in
PSIL in MW

6.3

500
50.4
4960

Open circuit and short circuit

After the discussion of the surge impedance loading, the behavior of lines
in the two extreme cases of no load, or open line, and short circuit will be
examined. For the sake of simplicity the lines are considered in each case as
lossless.

6.3.1

Open circuit

The case of open line operation can occur if an unloaded line is energized,
or if the load at one end of the line is switched off. This case is characterized
by the fact that the current at the end of the line is I 2 = 0. Thus from the
line equations we get (5.100a) and (5.100b):

(6.12a)
U 1 = U 2 cosh l

U2
I1 =
sinh l
(6.12b)
ZW
The voltage U 1 is assumed to be the phase reference, i.e. it coincides with
the real axis, and the equations (6.12a) and (6.12b) when R = G = 0 can

111

6.3. Open circuit and short circuit


be simplified to
U1 = U2 cos (l)
jU2
I1 =
sin (l)
ZW

(6.13a)
(6.13b)

If the voltage is fixed, then the voltage U2 at the end of the line becomes
U2 =

U1
cos (l)

(6.14)

and the input current of the line can be calculated to


I1 =

jU2
jU1 tan (l)
sin (l) =
ZW
ZW

(6.15)

Obviously, the amplitude of the voltage over the open-circuited line rises
towards line end, while the current amplitude decreases (see figure 6.2). The
rise of U1 to U2 over the open-circuited line is called the Ferranti Effect. In
extreme cases resonant conditions can arise if l = 2 = 90 , i.e. the line
length for 50 Hz is around 1500 km. Even if no resonant conditions arise,
still in practice the Ferranti effect can cause substantial overvoltages at the
open end if no measures are taken.
Here again the software PowerWorld can be mentioned. An example that
illustrates the Ferranti effect can be found under the set of examples provided
on the course web site - the file Long Line Example. With changing load
at the end of the line, the voltage distribution along the line changes too.
By switching the load off the voltage increase along the line becomes most
significant.
Finally we calculate the input impedance of the open-circuited high voltage transmission line:
U1
ZW
= Z 1 = j
I1
tan (l)

(6.16)

From the impedance representation it becomes obvious that the impedance


of the open line is capacitive. With a line length of 750 km, i.e. l = 45 ,
the value of the input impedance is as large as that of the surge impedance.
The arising currents (charging currents 3 ) are considerable.
In reality it must be remembered that the voltage at the beginning of
the line is not fixed, but by the capacitive charging currents often also will
increase. The occurring overvoltages at the end of an open line must therefore
be considered from line lengths of 300 km.
3

The charging current is defined as the current that flows through the line at no-load
conditions.

112

6. Power Transmission Fundamentals

U1 = U (x = 0)

U2 = U (x = l)

U (x), I(x)

P2 = 0
Q2 = 0

x
l = 300 km
1.05

U (x)
p.u.
1.00

0.3

I(x)
p.u.

0.2

0.1

0.0
0

100

200

300

x in km
Figure 6.2. Voltage and current characteristics for a 300 km long line
in no-load operation with a fixed voltage at the beginning of the line
(L = 1 mH/km, C = 10 nF/km).

6.3.2

Short circuit

A short circuit at the end of the line can be analyzed in a similar way as the
no-load case. In the case of an ideal short circuit at the end of the line the
output voltage is U 2 = 0, which for a lossless lines implies
U 1 = jI 2 ZW sin (l)

(6.17a)

I 1 = I 2 cos (l)

(6.17b)

In particular the short circuit current at the end of the line becomes equal
to
I2 =

I1
cos (l)

(6.18)

Again the case of resonance is also here recognizable with l = 90 (I 2 ).


The input impedance in the case of the short circuit at the end of the line

6.4. Reactive power demand of a line

113

is given by
U1
= Z1 = jZW tan (l)
I1

(6.19)

For 50-Hz systems the short circuit impedance is inductive for line lengths
up to 1500 km.

6.4

Reactive power demand of a line

If power is transmitted over a power line, then losses, active and reactive,
arise. In the resistive elements of the line (series resistance and shunt conductance) active power losses arise, in the reactive elements (series inductance
and shunt capacitance) reactive power is either consumed or generated. The
complex power at one end of the line differs consequently from the complex
power at the other end of the line, and the difference between these two
complex powers gives the active and reactive losses of the line. The active
losses are of course always positive, but as will be shown the reactive losses
can be either positive or negative. Negative reactive power losses means that
the line is a net producer of reactive power. For the voltage profile of the
line the reactive power is particularly important as will be discussed in the
following.
We consider the wave equation for a lossless line:
  
 
cos (l)
jZW sin (l)
U1
U2
=
j
cos (l)
I1
I2
ZW sin (l)
The power input at the beginning of the line S 1 = U 1 I 1 = P1 + jQ1 depends
on the power output at the end of the line S 2 = U 2 I 2 = P2 + jQ2 . For given
voltage U 2 and surge impedance loading PSIL = |U 2 |2 /ZW , the power at
the line input is




1 |S 2 |2
P1 + jQ1 = P2 + j Q2 cos (2l) +
PSIL sin (2l)
(6.20)
2 PSIL
For a lossless line P1 = P2 . The available reactive power Q1 is determined
besides Q2 to a high degree by the net consumption Q of the line. With
the approximation cos (2l) 1 one gets


1 |S 2 |2
Q = Q1 Q2
PSIL sin (2l)
(6.21)
2 PSIL
The reactive power demand of a line consists, according to this, of two parts:
inductive part:
QL =

1 |S 2 |2
sin (2l)
2 PSIL

(6.22)

114

6. Power Transmission Fundamentals


capacitive part:
1
QC = PSIL sin (2l)
2

(6.23)

The two parts are balanced at surge impedance loading of the line, i.e. the
shunt capacitances produce exactly the amount of reactive power as the
series inductance consumes:
Q = QL QC = 0 with

|S 2 | = PSIL

(6.24)

With |S 2 | > PSIL the line absorbs reactive power (QL > QC ), with |S 2 | <
PSIL the line generates reactive power (QL < QC ). The reactive power during no-load operation QC is capacitive, i.e. reactive power is produced by
the line, and is referred to as the charging power. This accounts for a
lossless line with R = G = 0
1
QC = PSIL sin (2l)
2

(6.25)

short line with U1 U2 U


QC = PSIL l =

|U |2
l = |U |2 C l
ZW

(6.26)

If the line is in no-load operation, i.e. S 2 = 0, then according to equation (6.22) the inductive reactive power demand QL is in both cases (lossless/short line) equal to zero.
We consider now the reactive power demand in dependence of the transmitted active power. We assume again a lossless line. The equation (5.106a)
with x = 0 is
U 1 = U 2 cos (l) + jZW I 2 sin (l)

(6.27)

The current at the end of the line can be calculated by use of the apparent
power at the end of the line as


P2 jQ2
U 1 = U 2 cos (l) + jZW sin (l)
(6.28)
U 2
We now introduce the transmission angle, , i.e. the phase angle difference
between U 1 and U 2 , and using U 2 as the phase angle reference, i.e. U 2 is
parallel with the real axis
U 1 = U1 (cos + j sin )

(6.29)

115

6.4. Reactive power demand of a line


800 km 600 km
0.6
400 km

Q2 = Q1

0.4

300 km
200 km

0.2

Q1
PSIL

0.2

0.4

0.6
0

0.2

0.4

0.6

0.8

1.0

1.2

1.6

1.4

P2
PSIL

Figure 6.3. The reactive power demand at the beginning of the line
as function of the active power at the end of the line. (U1 = U2 , lossless
line)

If we split the voltage into real and imaginary parts, then we achieve
{U 1 } = U1 cos = U2 cos (l) + ZW sin (l)
{U 1 } = U1 sin = ZW sin (l)

Q2
U2

P2
U2

(6.30a)
(6.30b)

The real part of the above equation gives the following expression for the
reactive power at the end of the line:
Q2 =

U2 (U1 cos U2 cos (l))


ZW sin(l)

(6.31)

The reactive power at the line input is thus


Q1 =

U1 (U2 cos U1 cos (l))


ZW sin (l)

(6.32)

In case the amplitudes of U 1 and U 2 are identical, it follows that:


Q2 = Q1 =

U12 (cos cos (l))


PSIL (cos cos (l))
=
ZW sin (l)
sin (l)

(6.33)

Figure 6.3 shows the reactive power demand at the beginning of the line
according to the equation (6.33) for lines of different lengths as function of
the active power transmitted over the line.
Both the active and the reactive power can be related to the surge
impedance loading of the line, PSIL . If the active power at the end of the
line is smaller than the surge impedance loading of the line (P2 < PSIL ),

116

6. Power Transmission Fundamentals


R + jX

U1

P2 + jQ2

B
2

Pload + jQload

U2 0
B
2

Figure 6.4. -model of a transmission line.

the line can be regarded as a capacitance and generates at both line ends
reactive power. In case P2 > PSIL the line behaves like an inductance and
reactive power has to be supplied. Thus, high voltage transmission lines can
be operated with variable load and approximately constant voltages at both
line ends only when at both line ends sufficient sources of reactive power are
available.

6.5

Voltage drop along a line

In this section expressions for the voltage drop along a transmission line will
be derived for different power transfers.
We assume the model in figure 6.4. The parameters R, X and B will
be calculated from the distributed parameters of the line, R , L and C (G
is neglected in the following). At the receiving end of the line the complex
power Pload + jQload is delivered.
The shunt admittances, B/2, of the line model consist of the shunt capacitances, i.e. they deliver reactive power. We obtain the reactive power
flow through the series inductance for the right part of the -model Q2 by
subtracting the reactive power, which is produced in the capacitive shunt
element B/2 from the inductive reactive load Qload . The transmitted active
power P2 corresponds to the active part of the load Pload . We thus get
P2 = Pload

(6.34a)

Q2 = Qload Qc

(6.34b)

whereby Qc corresponds to the reactive power produced in the right shunt


admittance. Thus we can replace the model in figure 6.4 by a simpler one
as shown in figure 6.5.
We can now express the current as function of the apparent power
S 2 = P2 + jQ2 and voltage U2 giving
I2 =

S 2
P2 jQ2
=
U 2
U2

(6.35)

117

6.5. Voltage drop along a line


U1

R + jX

U2 0

P2 + jQ2
Figure 6.5. Model of a transmission line for the computation of the
voltage drop.

U1
RP2
U2

I2

U2

RI 2

XP2
U2

jXI 2
RQ2
U2
XQ2
U2

Figure 6.6. Relation between the phasors U 1 and U 2 .

Again we select the phase angle of the voltage U 2 as reference (U 2 = U2 ).


Thus the voltage U 1 is easily obtained as
P2 jQ2
U 1 = U 2 + Z l I 2 = U2 + (R + jX)
U2




RP2 + XQ2
XP2 RQ2
=
U2 +
+j
U2
U2
The amplitude of the voltage at the beginning of the line is then
s



RP2 + XQ2 2
XP2 RQ2 2
|U 1 | = U1 =
U2 +
+
U2
U2

(6.36)

(6.37)

A phasor diagram of these relations is shown in fig. 6.6. For a lossless line
with R = 0 equation (6.37) is simplified to
s



XQ2 2
XP2 2
+
(6.38)
U1 =
U2 +
U2
U2
This simplified equation between U 1 and U 2 is given for a lossless line as a
phasor diagram in fig. 6.7. For realistic cases |XP2 /U2 | U2 , therefore the
right part of the equation (6.38) can be neglected and we get
U1 U2 +

XQ2
U2

(6.39)

118

6. Power Transmission Fundamentals

U1
j

jXI 2

XP2
U2

U2

I2

XQ2
U2

Figure 6.7. Simplified relation between the phasors U 1 and U 2 .

From fig. 6.7 it can be seen that the phase angle difference between the
two voltages on the first line is determined by the active power flow P2 , and
that the voltage (amplitude) difference U = U1 U2 is caused mainly by
the reactive power Q2 .
In the above investigations we assumed that voltage and powers are
known at the same end of the line. However, it is often so that the voltage
at one end of the line, e.g. at the generator end, and the powers at the
other end of the line, e.g. at the load, are given. In that case the equation
(6.37) is still valid, but now the voltage U1 is the known and U2 is the
unknown quantity. U2 can then be calculated from U1 , P2 and Q2 through
transformation of the equation (6.37).

6.6

Efficiency of high voltage transmission lines

Now we want to determine the efficiency of a high voltage transmission


line based on a realistic example. For a 200 km long 420-kV-line (= U2 ) the
efficiency of the transmission at surge impedance loading is to be calculated.
The primary data of this line are:
R
L
C
f

=
=
=
=

0.031 /km
1.06 mH/km
11.9 nF/km
50 Hz

The resistive shunt conductance is neglected (G = 0). Using the equation


(5.86) we can calculate the propagation constant from the line parameters:
=

(0.031 + j0.333) j3.74 106 = (0.052 + j1.117) 103

1
km

(6.40)

Together with the line length we obtain


l = l + jl = 0.0104 + j0.2234

(6.41)

6.7. P -U -Diagram

119

In the characteristic impedance we neglect the imaginary part and get


ZW = 298.5

(6.42)

The active power, which is consumed by the load at the end of the line
amounts to
U22
(420 kV)2
= 591 MW
=
ZW
298.5

(6.43)

In this case the phase current amounts to


U2
420 kV
I2 =
= 812.4 A
=
3ZW
3 298.5

(6.44)

The losses can be approximated by


P = P1 P2 3 R l I22 = 3 0.031 200 812.42 = 12.3 MW

(6.45)

The total power consumption of the line results thereby to P1 591+12.3 =


603.3 MW. The effect of the distributed line capacitance is in this case
neglected. Nevertheless the estimation coincides very well with the following
more exact calculation:
P1 = P2 e2l = 603.6 MW

(6.46)

Thus, the efficiency amounts to e2l = 0.979 or 97.9%. The efficiency of


power transmission over an overhead power line is consequently very high!

6.7

Voltage-Power relation of a power line

In case too much power is consumed at the end of a high voltage transmission
line, a substantial voltage drop can arise along the power line. This can lead
to stability problems. Therefore the relation between delivered power and
voltage of a power line is of importance.
We are going to discuss this relation for a 300 km long, lossless overhead
line with = 0.0013 rad/km. The voltage at the beginning of the line is
fixed and amounts to 1.0 p.u., at the end of the line a power of P2 + jQ2
is delivered. From these data the current at the end of the line can be
calculated:
P2 jQ2
(6.47)
I2 =
U 2
If we could now put the equation (6.47) into the line equation (5.107a) with
x = 0, we get
U 1 = U 2 cos (l) + jZW I 2 sin (l) = U 2 cos (l) + jZW

P2 jQ2
sin (l)
U 2
(6.48)

120

6. Power Transmission Fundamentals


1.5

cos = 0.90 cap.

1.0
0.98

stable

1.0
0.98

U2
U1

cos = 0.90 ind.

0.5

unstable

0.0

0.5

1.0

1.5

2.0

2.5

P2
PSIL

Figure 6.8. Voltage at the end of the line as function of the delivered
active power for different line loadings.

The equation (6.48) can be now solved for U 2 , which yields


U2 =

2
U 1 jZW P2 jQ
sin (l)
U
2

cos (l)

(6.49)

We will now examine this equation in more detail. The voltage at the end of
the line U 2 appears twice in the equation: once on the left side, and a second
time as complex conjugate on the right side. Since the complex voltage U 2
appears in different forms in equation (6.49) an analytic solution is not
easily obtainable. However U 2 can easily be calculated in an iterative way:
One begins with an initial value (usually 1 p.u. real) for U 2 and calculates
with this value the right side of (6.49). The complex conjugate of U 2 is
then inserted into the right side of the equation and a new estimate of
U 2 is calculated. One repeats this iteration until the complex voltage does
not change significantly anymore between two iterations, i.e. the result has
converged.
The figure 6.8 shows the voltage at the end of the line as function of the
delivered active power at the end of the line for different load cases. Because
of its form this representation is often called the nose curve of the power
line.
From figure 6.8 it is noticed that there exists a maximum transferable
active power, and that this limit is highly dependent on the power factor
of the load(cos ). Further, as shown in the figure 6.8, the power factor of

6.8. P --Diagram

121

the load has a significant influence on the voltage at the end of the line at
the maximum transferable power. The maximum transferable active power
and the voltage at the end of the line are lower in the case of an inductive
power factor. For a capacitive power factor of the load, the voltage profile at
the upper range of the curve in figure 6.8 becomes flatter and the maximum
transferable active power larger. That means that the voltage at the end of
the line can be adjusted by additional capacitances at the end of the line
(= reactive shunt compensation).
Figure 6.8 shows also that an active power transfer below the maximum
transferable active power can be done theoretically at two different voltage levels. Normally the higher value is selected, with a voltage magnitude
around 1.0 p.u. For the lower voltage value, the current must be larger, so
that still the same amount of active power can be transferred over the line.
Thus the power losses in the line would increase and thus the line could come
closer to (or even exceed) its thermal limits. The operation of the line at the
lower voltages (below the dashed line in figure 6.8) can lead to transmission
instabilities.

6.8

Active Power as function of phase shift between


voltage phasors

Now we will examine another important relation, i.e. between transferred


active power and the phase angle difference of the voltages at the beginning
and the end of the line.
For this we assume the voltage amplitude at the line ends as fixed. The
phasors have thus constant amplitudes, but the phase angle difference (shift)
between them can vary. The transmitted active power as function of this
angle difference will now be calculated.
The line is approximated as a purely inductive series impedance jXL =
jL. The voltages at the beginning and the end of the line can be written
in polar representation as
U 1 = U1 ej1
U 2 = U2 ej2
Since the line was assumed lossless, the active powers at both ends are equal
P1 = P2 = {U 1 I 1 }

(6.51)

The current and its complex conjugate are calculated as follows:


U1 U2
U1 ej1 U2 ej2
=
jXL
jXL

j
j
2
1
U1 e
U2
j 
I 1 =
=
U1 ej1 U2 ej2
jXL
XL
I1 =

(6.52a)
(6.52b)

122

6. Power Transmission Fundamentals

One of the voltages can be brought into a pre-defined position. We do this


by defining the angle of the voltage at the end of the line: 2 = 0. The
transferred active power can now be computed with the following equations:
P1 = {U 1 I 1 }



j1 j
j1
U1 e
U2
= U1 e
XL


1 j1
2 1
= jU1
jU1 U2
e
XL
XL


1
1
2 1
= jU1
jU1 U2
cos 1 + U1 U2
sin 1
XL
XL
XL
1
1
= U1 U2
sin 1 = U1 U2
sin
XL
XL

(6.53)

= 1 2 is here the angle difference between the voltage at the beginning and the end of the line and is often referred to as the transmission
angle. The reactive power Q1 follows directly from the imaginary part of
equation (6.53):
Q1 = U12

1
1
U1 U2
cos
XL
XL

(6.54)

Figure 6.9 gives a graphical representation of equation (6.53). In the


upper picture the active power is shown as function of the transmission
angle, in the lower diagram the voltage in the center of the line (Um ) as
function of the active power flow.
If the transmission angle is increased from zero, then the transferred
active power increases, which causes a reduction of the voltage in the middle
of the line and an increase of the current. Up to a certain point the increase of
the current dominates as compared with the voltage drop in the middle of the
line. If the transmission angle reaches = /2 = 90 , then the transferred
power over the line is at its maximum. For larger values of the transmission
angle the voltage in the middle of the line decreases more than the current
increases, and the transferred power drops consequently. Once the point
of maximum transferable active power is reached, then, if the operation
point enters the lower branch, the voltage decreases for a reduction of the
transferred power. For this operating mode, thus the voltage would drop
although the power is reduced, which is an abnormal situation in power
systems. All points at the lower branch of the voltage curve in figure 6.9 are
therefore called unstable. The maximum transferred active power constitutes
the steady-state limit.
We consider now a 400 km long, lossless power line with l = 0.52 rad,
which is the line used in figure 6.9. Here we have a theoretically maximum
transferable power of 2.012 PSIL . In practice however a line is never loaded

123

6.9. Load diagram and load limits

2.0

P2
PSIL

2.012

1.0

0.0

29.8

180

90

Transmission angle in

1.5

1.0

Um
U1

0.5
2.012

0.0

0.5

1.0

P2
PSIL

1.5

2.0

2.5

Figure 6.9. Top: P --Diagram. Bottom: Voltage in middle of the line


as function of the transmission angle. Both figures for a 400 km line
with l = 0.52 rad.

to 100% for stability reasons, but normally only to approx. 70% of this limit,
which in this case is 0.7 2.012 PSIL = 1.4084 PSIL .
The analysis in this chapter represents an idealized situation. The assumption that voltage amplitude at the end and the beginning of the line is
constant and independent of the line loading, is not always valid.

6.9

Loadability limits of power lines

Pioneering work in this area was carried out by H. P. St. Clair [14], who
first derived the curves for load limits of high voltage transmission lines
from practical experiences, which subsequently were further elaborated by
other engineers.
High voltage transmission lines have certain loading limits. Fundamen-

124

6. Power Transmission Fundamentals

Pmax
1.0

Plimit

Pmax sin

P2
PSIL

0.0

29.8

180

90.0

Transmission angle in

Figure 6.10. Transmitted active power as function of the transmission angle.

tally the power lines are subject to three different limits originating from
different physical phenomena:
Thermal limit: If the temperature of the conductor gets too high due to
the resistive losses, this will lead to an increased sag of the power line
(see p. 78). An increased sag implies that the clearing distances to
objects on the ground, e.g. trees, decreases and the risk for flashovers
rises. High temperatures in power cables can lead to accelerated aging
of the insulation.
Voltage drop: For operational reasons and for reasons of power quality
the voltage at a bus should not deviate too much from the nominal
voltage (typically less than 10%).
Transmission angle: For stability reasons a certain transmission angle
should be not exceeded but a steady state stability margin should be
kept. This is defined as
steady state stability margin =

Pmax Plimit
100%
Pmax

(6.55)

Figure 6.10 shows the dependance on the transmitted active power of


the transmission angle. Plimit indicates the maximum power, for which a
stability margin of at least 30% is maintained. The maximal transmission
angle should therefore be less than around 40 .
Figure 6.11 shows the maximum transferable active power in p.u. (power
base = the surge impedance loading, PSIL ) as function of the line length. A
maximum voltage drop of 5% of the nominal voltage and a stability margin
of 30% are assumed. As suggested in this figure, the load of high voltage
transmission lines is determined by the following approximate limits

125

6.10. Power cables

3.0

2.5

Line

2.0

Pline
PSIL

1.5

1.0

0.5

2
160

3
320

480

640

800

960

Line length in km
Figure 6.11. Maximum active power loading as function of the line length.

1. up to 80 km: thermal limit


2. between 80 and 320 km: voltage drop
3. above 320 km: stability limit
Figure 6.11 also shows that for line lengths from approx. 400 km the line
loading must be smaller than the surge impedance loading. As previously
mentioned, the power transmission capability of (long) power lines can be
increased by reactive compensation, shunt or series.
The limits of the power transmission have been represented here in a
simplified way in order to give a general overview and understanding of the
physical phenomena. In large, complex power systems, detailed investigations must be undertaken for the determination of the transmission capacities. However, the basic mechanisms that determine the maximum power
transfers are still the ones described above.

6.10

Power cables

A three-phase AC cable consists of three phase conductors, whereby a conductor usually consists of several twisted copper or aluminium wires. A highvoltage cable has at the outside of the metallic conductor, which is placed
in the center, an electric insulation consisting of either solid (e.g. a polymer), in layers solid/liquid (e.g. oil-impregnated paper4 ), or gaseous (e.g.
4

Instead of the oil-impregnated paper today also a light oil, under pressure, is used as
isolating medium, which decreases the creation of voids (places of partial discharges) and
thus the aging of the insulating medium.

126

6. Power Transmission Fundamentals

Figure 6.12. Example of a high voltage cable: nominal voltage


132 kV, nominal power 315 MVA, copper conductor with cross section 2000 mm2 , XLPE isolation (cross-linked polyethylene), weight
40 kg/m. source: ABB.

sulfur hexafluoride SF6 ) dielectric insulation medium. The outside shield of


the insulating layer is formed by one or more metallic layers. An outer nonmetallic layer serves often as passive corrosion protection for the metallic
layers. Figure 6.12 shows a high-voltage cable for AC operation.
The circuit data of cables, more precisely L and C , differ substantially
from those of overhead lines, although the conductor cross sections lie in the
same order of magnitude. Reasons for this were discussed in section 5.2.5.
Now the operational characteristics of a cable will be discussed for a
specific example. We consider a 420kV cable with the following data:
R
L
C

=
=
=

75 m/km
0.35 mH/km
0.2 F/km

If the resistance R is neglected, one obtains for the surge impedance


r
L
ZW =
= 42
(6.56)
C
Therefore the surge impedance loading (see section 6.2) amounts to
PSIL =

2
Unominal
(420 kV)2
=
= 4.2 GW
ZW
42

(6.57)

To transmit this power a phase current of I = 5774 A would be necessary. With a realistic conductor cross section of 500 mm2 this would imply
a current density of 11.5 A/mm2 and losses of 2.5 MW/km. These are values which are neither technically nor economically realistic. Based on the
possible heat dissipation and the economic optimization a nominal power
of approximately 500 MVA is realistic. The nominal phase current is then
approximately 700 A.

127

6.10. Power cables

It should be noted that in no-load operation of the cable considerable


currents can occur because of the low surge impedance. The large shunt
capacitance leads to large charging currents in longer cables. The current in
no-load operation for nominal voltage is given by
I=


U
tanh l
ZW

(6.58)

The propagation constant is equal to:


=

(R + jL ) jC = 0.00085 + j0.00276 rad/km

(6.59)

The phase constant, , of the cable is about three times as large as of a typical overhead line of the same nominal voltage. The characteristic impedance
on the other hand is only approximately 1/10 of the surge impedance of a
comparable overhead line. If a value of 700 A is taken as limit of acceptable
current load, then the open-circuited cable (without compensation) may be
not much longer than approximately 50 km. This length can be determined
by means of equation (6.13b):
I1 =

jU2
sin (l) 3 790A
ZW

for a line with a length of l = 50 km.

(6.60)

128

6. Power Transmission Fundamentals

7
Symmetrical components in three-phase
systems
In this chapter unbalanced operation conditions in three-phase systems will
be investigated, which e.g. arise due to ground faults or short-circuits. The
so called symmetrical components will be introduced in order to analyze unsymmetries in an efficient way. The influence of different ways of neutral
point grounding will also be discussed.

7.1

Unbalanced operating conditions

Unbalanced operating conditions occur due to faults, asymmetrical loads or


because of untransposed, and thus unsymmetrical, lines. By faults are here
understood unwanted changes of the voltage, insulation or switching status
of the network. Faults are usually classified as shunt and series faults.
A shunt fault is a fault where an unwanted connection between one or
more phases and ground or between phases arise. The possible shunt faults
in three phase systems are:
Ground faults
single fault (ground contact in one location)
double fault (ground contact simultaneously in two different locations)
triple (ground contact simultaneously in three different locations)
two-phase short-circuit
without ground contact
with ground contact
three-phase short-circuit
without ground contact
with ground contact
A special fault type is the double or the triple ground fault, consisting of two
or three spatially separated single ground faults. Short-circuits are the most
129

130

7. Symmetrical components in three-phase systems

dangerous fault types in electrical power systems because of the high mechanical and thermal stresses due to the high short circuit currents through
different pieces of equipment. The short circuit current can be several times
higher than the normal load currents. They can cause substantial damage to
the equipment if this is not designed to withstand these stresses, and they
can also be a hazard to personnel.
By series faults one understands failures along the power line, that
causes undesirable interruptions or connections in the circuit. Examples of
series faults are if a conductor is disrupted or a breaker does not trip the
three phases as desired. Possible series faults in three-phase systems are:
Single-phase disruption
Two-phase disruption
Three-phase disruption
Due to series faults asymmetrical operating conditions can occur, which
could damage the equipment. For example three-phase asynchronous machines in two-phase operation, after the loss of one phase, can be subject to
a high thermal loading which can in worst case lead to the destruction of
the machine.
During three-phase short-circuits/interruptions the symmetrical conditions remain, all other faults will imply unsymmetrical conditions. An efficient method for analyzing unsymmetrical conditions in multi-phase systems
is by the so called symmetrical components, which consist in a mathematical transformation of the three physical phase quantities (RST) into a new
system with three new quantities called the positive, negative and zero sequence components. Through this coordinate transformation unsymmetrical
conditions can often be analyzed in a more efficient way.

7.2

Symmetrical components

A multi-phase system with linear relationships between currents and voltages can generally be described by system matrices, e.g. the admittance and
the impedance matrices Y and Z. The voltages and the currents at a component, e.g. a generator, transformer, or consumer, are a linear function of
the voltages of the feeding generators in the system. A simple example will
clarify this.
A three-phase generator feeds a delta connected load. The complex currents, magnitude and phase, can be calculated with the help of the admittance matrix from the complex generator voltages. For linear circuits we
have
I=Y E
(7.1)

131

7.2. Symmetrical components


UR
ER

Zl

IR

US

Zd

ES

Zl

IS

UT

Zd

Zd

ET

Zl

IT

Figure 7.1. A simple three-phase system.

where Y represents the (complex) admittance matrix of the circuit, I and E


are the currents and the generator voltages, respectively, of the three-phases
R, S and T.
For the system shown in figure 7.1 the phase currents can be calculated
on the basis of the phase voltages as follows:
UR US UR UT
+
Zd
Zd
US UT US UR
IS =
+
Zd
Zd
UT UR UT US
IT =
+
Zd
Zd

IR =

(7.2a)
(7.2b)
(7.2c)

In these equations the phase voltages are given by the source voltages and
the phase currents
U R = ER Zl I R

(7.3a)

U S = ES Zl I S

(7.3b)

U T = ET Zl I T

(7.3c)

and each phase current is consequently a function of the source voltages,


which in matrix form can be written as


IR
2 1 1
ER
1
IS =
1
2 1 E S
(7.4)
3 Zl + Zd
IT
1 1
2
ET
| {z } |
{z
} | {z }
I

In general the admittance matrices of three-phase symmetrical power


systems can be assumed to have cyclic symmetry. This assumption is the

132

7. Symmetrical components in three-phase systems

basic requirement for the transformation resulting in the symmetrical components.




IR
ER
I S = E S
IT
ET

|
{z
}

(7.5)

cyclicly symmetric

Multi-phase systems become more difficult to analyze above all if the


system matrices have few zero elements, i.e. an output quantity is coupled
to many input quantities. This is most often not the case for real systems,
whose system matrices are sparse.
Only if the circuit is completely symmetric it can be completely described
by a single-phase equivalent circuit. In case of a contingency, where the system becomes asymmetrical, this is no longer possible. Of course the effects
of any unbalance, e.g. a single-phase ground fault on a line, can be calculated by use of the system equations for the three physical phases (RST)
by standard network analysis methods. However, the relations between the
triggering causes and the outputs are in most cases rather incomprehensible. The symmetrical components enable a procedure that makes possible
a systematic analysis of unsymmetrical operating situations in three-phase
circuits.
The symmetrical components allow a three-phase AC circuit to be represented by three single-phase systems, namely the positive, negative, and zero
sequence systems. Mathematically considered, the basic idea of this method
is the transformation of the physical system (RST system) into symmetrical
components. In this new system, the calculations can be performed much
easier. Subsequently, the results of the calculations are transformed back
into the RST system.
The desired simplifications in the new system are due to the fact that
the symmetrical components are the eigenvectors of the admittance matrix.
A vector x is called eigenvector of a matrix A with eigenvalue if
Ax =x

(7.6)

The vector x does not change direction by multiplication with the matrix,
but its magnitude is only changed by a (scalar) factor . A further characteristic of the eigenvectors is that the matrix A can be diagonalized by the
transformation matrix T, whose columns are the eigenvectors of A:
D = T1 AT

(7.7)

Not all matrices can be diagonalized but normally the admittance matrix
Y of a three-phase circuit can be diagonalized. If an admittance matrix

133

7.2. Symmetrical components


Table 7.1. Eigenvectors and eigenvalues of cyclic symmetric matrix

Symmetrical components

Eigenvectors

Eigenvalues

Positive sequence

x1 = a2
a

1 = + a2 + a

Negative sequence

Zero sequence


1
x2 = a
a2

2 = + a + a2

x0 = 1
1

0 = + +

is diagonal, this means that the currents in eq. (7.1) are each dependent
only on one voltage and independent of the others, and we have decoupled
single-phase systems.

It can be shown that a general cyclic symmetric matrix, as in equation


(7.5), has the eigenvalues and eigenvectors given in table 7.1, where the
constant a is

1 + j 3
j120
=
(7.8a)
a=e
2
1 j 3

and a2 = ej240 =
(7.8b)
2
This complex constant is frequently used in calculations in three-phase systems. It has a magnitude

a = ej120

=1

(7.9)

and a phase of +120 . Thus, the elements of the eigenvectors correspond to


three phasors, each of the same magnitude but phase-shifted by 120 , which
vectorially add to zero.
1 + a + a2 = 0

(7.10)

This very important property of symmetrical components is illustrated in


figure 7.2.

134

7. Symmetrical components in three-phase systems


a
1

a2
1

a2
Figure 7.2. Eigenvectors of a cyclic symmetric matrix.

One should also mention that the eigenvectors in table 7.1 (contrary to
the eigenvalues) are not dependent on the elements , and of the cyclic
symmetric matrix. The eigenvectors can consequently be chosen as identical
for all cyclically symmetrical matrices and are thus suitable for calculations
in arbitrary three-phase systems.
With

1
1
1

T = x1 x2 x0 = a2 a 1
(7.11)
a a2 1
the admittance matrix can be diagonalized:

Y1 0
1 0 0
0
Y 120 = T1 Y T = 0 2 0 = 0 Y 2 0
0 0 0
0
0 Y0

(7.12)

Hence and from the table 7.1 it is obvious that for = the positive and
negative sequence admittances are equal.
The voltage sources and current vectors E120 , I120 of the symmetrical
components are consequently
I120 = Y 120 E120
120

I
TI

120

= T

(7.13)
120

YTE

= Y T E120

Hence and from (7.1) it follows that




IR
I1
I = I S = T I 2 = T I120
IT
I0


ER
E1

E = ES
= T E 2 = T E120
ET
E0

(7.14)
(7.15)

(7.16)

(7.17)

With these relations and the equations (7.3a), (7.3b) and (7.3c) it can also
be shown that

135

7.2. Symmetrical components


U1
UR

U = U S = T U 2 = T U120
UT
U0

(7.18)

The matrix Y 120 is diagonal, i.e. there are no coupling impedances between the positive, negative and zero sequence systems, and the symmetrical
components are thus totally decoupled from each other. Therefore calculations are much more easily done in the 120 system. It should be noted that
the admittance matrices in the RST system and the corresponding one in
the 120 system contain the same information.
The basic steps for the analysis of three-phase circuits with symmetrical
components have thus been explained. The transformations RST 120
and 120 RST are linear, where the back-transformation matrix T has as
columns the eigenvectors of the cyclically symmetric admittance matrix. The
eigenvectors could be multiplied with any scalar factor 6= 0, without losing
their particular characteristic. The choice of eigenvectors made here means
that elements of T have the magnitude 1, which implies that a symmetrical
three-phase system quantity corresponds to a positive sequence quantity
of the same amplitude. This means that the transformation is amplitude
invariant. (Other choices of the magnitudes of the eigenvectors can e.g. result
in power invariant transformations.)
The transformation matrix S , which is used for the transformation RSTsystem 120-system, is calculated as the inverse of T . The definition of
the eigenvectors having elements of magnitude 1 results in the fact that all
elements of the matrix S have the magnitude 1/3.

1 1 1
T = a2 a 1
a a2 1

S = T 1

1 a a2
1
= 1 a2 a
3
1 1 1
(7.19)

The transformation between the RST and 120 systems and of the corresponding back-transformation can be summarized in one equation. Thereby
Y = T Y 120 S
is used in (7.1) resulting in:

(7.20)

136

7. Symmetrical components in three-phase systems

IR
I S =
IT

1
= a2
a
|



ER


ES =
(7.21)
ET



1 1
Y1 0
0
1 a a2
ER
1
a 1 0 Y 2 0 1 a2 a E S
3
0
0 Y0
a2 1
1 1 1
ET
|
{z
}
First step: Transformation

{z

Second step: Solution

{z

Third step: Back transformation

}
}

The expansion of this equation from right to left corresponds to the following
steps:
1. Transformation (RST 120) by multiplication with S (This step is
sometimes called symmetrization.)
2. Calculation in the 120-system by use of the diagonal admittance matrix
Y 120
3. Transformation (120 RST) by multiplication with T (This step is
sometimes called de-symmetrization.)
As an overview the formulae, which are needed for the above transformations, are given in eq. (7.22).

RST
I
U
E
Y
I

=
=
=
=

120
T
T
T
T

I120
U120
E120
Y120 S

= YE

I120
U120
E120
Y120
I120

=
=
=
=

S
S
S
S

I
U
E
YT

= Y 120 E120
(7.22)

Thus, we can now transform a the physical RST three-phase system


mathematically into a 120 system, i.e. into symmetrical components. All
variables with index 1 belong to the positive sequence system, those with
index 2 belong to the negative sequence system and those with index 0 to
the zero sequence system.

137

7.2. Symmetrical components

We consider now the calculation of the phase voltages of the RST system
from the phase voltages of the symmetrical components in order to get a
better understanding of the positive, negative and zero sequence systems.
UR = U1 + U2 + U0

(7.23a)

U S = a2 U 1 + a U 2 + U 0

(7.23b)

UT = a U1 + a U2 + U0
Positive sequence system
system are

(7.23c)

The phase voltages of the positive sequence


U 1R = U 1

(7.24a)

U 1S
U 1T

(7.24b)

= a U1

(7.24c)

= a U1

This corresponds to the multiplication of the eigenvector x1 with the voltage


phasor U 1 . Figure 7.3 shows the graphical representation of the equations
above. The phasors have equal magnitudes and a phase shift of 120 with
respect to each other.

U 1T
a

a2

U 1R

U 1S

Figure 7.3. Voltage phasors of the positive sequence system.

In a balanced, three-phase system, symmetrical conditions are assumed.


Only symmetrical voltages and currents occur, i.e. equal magnitudes and
phase shifts of 120 between the phases, which is also the case in figure
7.3. Therefore only positive sequence quantities are present for balanced
conditions in a three-phase power system. Thus, the positive sequence system
describes the normal operation. As the sum of the three current phasors is
always zero, the current through the neutral points of wye connections is
thus also zero. Consequently, it does not make any difference whether the
neutrals are grounded are not. However, if the currents are unbalanced so
that the sum is not zero anymore the grounding of the system plays a crucial
role as will be seen later.

138

7. Symmetrical components in three-phase systems

Negative sequence system The two other eigenvectors x2 and x0 describe


deviations from the ideal balanced operation. The phase voltages of the negative sequence system are
U 2R = U 2

(7.25a)

U 2S = a U 2

(7.25b)

U 2T

= a U2

(7.25c)

The multiplication of the eigenvector x2 with the voltage phasor U 2 gives


phase voltages of the negative sequence system. Figure 7.4 shows the corresponding phasors. Again it is a balanced three-phase system, however with
the phase sequence R-T-S, i.e. reverse direction of rotation of all currents
and voltages as compared with the positive sequence system. That explains
the name negative sequence system.

U 2S
a

U 2R

a2
U 2T

Figure 7.4. Phase voltages of the negative sequence system.

Also in the negative sequence system all neutrals of wye connections


have currents equal to zero. Lines, transformers, and passive loads, which
are indifferent to the direction of rotation, have the identical properties in the
positive and negative sequence systems. Thus, the corresponding impedances
1
Z 1 = Y 1
1 and Z 2 = Y 2 are identical. In rotating machines, many quantities are very sensitive to the direction of the rotation and the positive and
negative sequence system properties are thus quite different.
Zero sequence system

The phase voltages of the zero sequence system are


U 0R = U 0

(7.26a)

U 0S
U 0T

= U0

(7.26b)

= U0

(7.26c)

The phase voltages correspond thus to the eigenvector x0 multiplied by the


voltage phasor of the zero sequence system. Figure 7.5 shows the phasor
voltages of the zero sequence system. Obviously, the zero sequence system
is a triplet of single-phase voltages and currents of the same magnitude and
phase in all three phases.

139

7.2. Symmetrical components

U 0R = U 0S = U 0T

Figure 7.5. Voltage phasors of the zero system sequence system.

Even if a positive sequence system is the dominating one, the phase voltage phasors become unequal due to the zero sequence system. Zero sequence
system currents can only flow through the neutrals of the wye connections,
and are always zero if such a return path does not exist. Contrary to positive
and negative sequence systems, the grounding of the neutrals is therefore of
critical importance for the zero sequence system. The impedances Z 0 = Y 1
0
of the lines and transformers in the zero sequence system are generally bigger than in the positive and negative sequence systems.
It must always be clearly stated whether the components are expressed
in RST or in the symmetrical component system. From eq. (7.23a) it is seen
that the corresponding phasors of the symmetrical component system can
be added vectorially when transforming from the 120 system to the RST
system. For the systems in the figures 7.3, 7.4 and 7.5 this is shown in the
figure 7.6.
U 1T
U 2T
U 0T
U 0S

U 2S

UT

US

UR

U 1R

U 0R
U 2R

U 1S

Figure 7.6. Change from symmetrical component system to the RST system.

During normal, i.e. balanced, operation of a three-phase system only the


positive sequence system is active and all voltages and currents of the negative and zero sequence systems are zero. Unbalanced operation implies that
the negative and zero sequence system components are not zero any longer.
The system conditions are now simpler to analyze in the symmetrical component system than in the original RST system, because the coupling between
symmetrical components is zero, or at least much smaller. In circuits that
are designed for balanced conditions, only a zero sequence system voltage

140

7. Symmetrical components in three-phase systems

will give rise to zero sequence currents.

7.3

Powers in symmetrical component system

Now the powers in the symmetrical component system will be considered.


Generally, the complex power in three-phase systems is calculated from the
sum of the three-phase powers
S = U R I R + U S I S + U T I T = UT I

(7.27)

We will now express the complex power using the symmetrical components
of the three-phase system. For the currents and the voltages in the RST
system we can use the vectors in the 120 system multiplied by the matrix
T:
U = T U120
I = TI

120

Inserted in equation (7.27) we obtain for the total complex power


T

S = T U120 T I120 =
T

= U120 T T (T) I120
| {z }

(7.28)
(7.29)

(7.30)

3E

We can simplify this equation by calculating the middle part



T
T
T
T T (T) = ((T ) ) T = 3 T1 T = 3 E

(7.31)

where E represents the identity matrix. Thus the equation (7.30) gives
T

S = 3 U120 I120 =
= 3 (U 1 I 1 + U 2 I 2 + U 0 I 0 )

(7.32)

Due to the choice of the eigenvectors and the T matrix a factor 3 arises
between the expressions of the powers in the RST and symmetrical component systems. The symmetrical three phase system is transformed into a
positive sequence system of the same amplitude, i.e. the phasors of the 120
system can be just added vectorially, in order to obtain the values in the
RST system, as done in figure 7.6. Therefore the transformation used here
is called amplitude invariant. Generally the conversion of the powers is quite
straightforward, because each element of the RST system carries exactly the
triple power as compared with its equivalent in the symmetrical component
system.
The reason for this factor 3 can also be explained as follows. In section
7.2 we saw that the phase voltages in the RST system consist each one

7.3. Powers in symmetrical component system

141

of three voltages in the 120 system. For instance, the U R is composed of


U 1R , U 2R and U 0R , and I R of I 1R , I 2R and I 0R . These three voltages and three
currents have the same amplitude as U 1 and I 1 . We can now calculate the
power with the help of these voltages and currents, which for the R-phase is
U R I R = U 1R I 1R + U 2R I 2R + U 0R I 0R

(7.33)

Since the voltage U 1R and the related current I 1R arise from the same phase
angle shift U 1 , and I 1 respectively, the product is the same as from U 1 I 1 .
The same applies for the second and third term of the right side of (7.33).
If the same is done for the S and T phases, we see that
U R I R + U S I S + U T I T = 3 U 1 I 1 + 3 U 2 I 2 + 3 U 0 I 0

(7.34)

From these considerations it should be clear that in each phase of the RST
system positive, negative, and zero sequence currents flow as well. Since
these differ only by phase shifts, they can be calculated from single positive,
negative, and zero sequence systems.
We have already mentioned that the used positive sequence transformation is amplitude invariant. One can also define a power invariant transformation between the two systems. The transformation and back transformation matrices differ from those in (7.19) by a scalar factor. For the power
invariant transformation one has

(7.35)
S = U 1P I 1P + U 2P I 2P + U 0P I 0P =
= (SP U)T (SP I)
T

= 3 (S U) (S I)

(7.36)
(7.37)

The transformation matrix SP in the power invariant transformation is thus

SP = 3 S
(7.38)
and the back transformation matrix is
1
TP = T
(7.39)
3
This means that the transformation matrix of the power invariant transformation correspondsto that of the amplitude invariant transformation
multiplied by a factor 3 and the back transformation matrix to that of the
amplitude invariant transformation multiplied by a factor 13 . Consequently

U 120
=
3 U120
(7.40)
P

120
120
IP
=
3I
(7.41)
For the calculation of the phasors in the RST system from those in the 120
system of the power invariant transformation, the amplitudes of the 120
phasors must be multiplied first by 13 and then they can be added.
In the following sections we will use the amplitude invariant transformation.

142

7. Symmetrical components in three-phase systems

7.4

120 equivalent circuit

So far we have discussed the transformation of the voltages and currents


between the RST system and the 120 system and vice versa, and the phasors
of the positive, negative and zero sequence systems. But how can we conceive
a circuit in the symmetrical components system?

7.4.1

The basic power system

The starting point is the Thevenin equivalent of a three phase power system
in figure 7.7. Here Z S are the self impedances of the three-phase system and
Z M are the mutual impedances.
Z1
ER

ZS

IR

~
ES

ZS

ZM

~
ZS

UR
Z2
US
E2

ZM

IT

U1

ZM
IS

ET

E1

I1

I2
U2

UT
Z0

ZE

I0

IR + IS + IT
E0

U0

Figure 7.7. Thevenin equivalent of a three phase power system.

For phase R, for instance, we obtain the following relation:


U R = (I R + I S + I T )Z E + E R Z 1 I R Z 2 I S Z 2 I T

(7.42)

For the complete three-phase system, we have the following system of


equations:


ER
ZS + ZE
E S = Z M + Z E
ET
ZM + ZE

ZM + ZE
ZS + ZE
ZM + ZE


ZM + ZE
IR
UR
Z M + Z E I S + U S (7.43)
ZS + ZE
IT
UT

7.4. 120 equivalent circuit

143

With the aid of (7.22) this can be transformed into the 120 system:


E1
Z1
0
0
I1
U1
E 2 = 0 Z 2

0
I 2 + U 2
(7.44)
E0
0
0 Z0
I0
U0

with Z 1 = Z 2 = Z S Z M and Z 0 = Z S + 2 Z M + 3 Z E . If the mutual


impedances are neglected, the impedances of the positive and negative sequence systems are equal to the self impedances of the symmetric three
phase system. In the case that the ground impedance is equal to zero, then
the impedance of the zero sequence system is also equal to the the self
impedance.
This RST system with cyclically symmetric impedances can thus be represented by an equivalent system consisting of three decoupled circuits, the
positive, negative and zero sequence systems. In the following, we will discuss
these three systems.

Positive sequence system The circuit of the positive sequence system corresponds to the single-phase equivalent circuit of the balanced three phase
system. The voltage of the generator in the positive sequence system, which
is equal to the voltage E R of a symmetrically operated generator, feeds
the circuit. The impedances of the positive sequence system are included
with their impedances Z 1 . It should be noted that the Z 1 is independent
from neutral to ground impedance Z E . For the positive sequence system the
grounding of the neutrals is irrelevant, since the sum of the currents is zero.
Delta connected elements must be transformed into wye connections. In the
symmetrical, three-phase circuit all neutral points have the same potential,
it does not matter whether they are connected or not. In the equivalent
circuit of the positive sequence all neutral points can thus be regarded as
connected with each other.
Negative sequence system The equivalent circuit of the negative sequence
system is fundamentally derived analogously to the positive sequence system.
However the voltage source component of the generator voltage is zero, so
that usually no supply voltage exists in the circuit. In the passive part of
the network the impedances Z 2 , are equal to Z 1 . Hence, the neutral point
grounding has likewise no influence on the negative sequence system. In the
equivalent circuit of the negative sequence system, delta wye transformations
must be done and all neutral points must be connected to each other.
Zero sequence system The zero sequence system is fed by the zero sequence component of the generator voltage, which is zero for symmetrical
generators. The zero sequence impedances Z 0 of the passive elements, which
generally differ from the positive and negative sequence system impedances,

144

7. Symmetrical components in three-phase systems

are to be included. The treatment of the neutral points is very important in


the zero sequence system. As mentioned above zero sequence currents can
only flow through the neutral point connections. Neutral points could therefore only be connected in the equivalent circuit of the zero sequence system
there and only there, where in the real physical system actually connections
do exist. Possible impedances at the neutral point connections must be included with the triple of the physical impedance values. The threefold value
is necessary because the triple of the real zero sequence current flows through
the neutral ground connection, and thus, to achieve the same voltage drop,
the factor three is needed.
Figure 7.7 shows the resulting equivalent circuits of the positive, negative
and zero sequence systems. For the representation of a three phase system
often the circuit shown in figure 7.8 is used. This figure represents the 120
system as well as the RST system.
UR
E1, E2, E0
Z 1, Z 2, Z 0

US
UT

Figure 7.8. Schematic representation of a three phase system.

If the three-phase system is symmetrically operated, we have


E S = a2 E R

ET = a ER

Hence the generator voltages in the 120 system are

ER
E1
1 a a2
ER
1
E 2 = 1 a2 a a2 E R = 0
3
E0
1 1 1
a ER
0

(7.45)
(7.46)

(7.47)

The generator voltages in the negative and in zero sequence systems are thus
zero. In figure 7.9 the resulting equivalent circuits are given.

7.4.2

Neutral point grounding

As pointed out already several times, the treatment of the neutral point is
very important. So far we have assumed that the generator neutral point is
grounded through an impedance Z E , but this is not always the case. In figure

145

7.4. 120 equivalent circuit


Z1
E1

I1
U1

Z2

I2
U2

Z0

I0
U0

Figure 7.9. Thevenin equivalent of a balanced three phase system.

7.10, a three phase system, whose generator neutral point is not grounded,
is shown.
In the preceding sections, it was already mentioned that for a system as
in figure 7.10 no zero sequence current can flow. But what does this actually
mean? In figure 7.10 the corresponding 120 system is given. The positive and
negative sequence systems remain unchanged. However, in the zero sequence
system the ground connection is interrupted. Thus no current can flow there.
One could also regard the zero sequence impedance Z 0 as infinitely large,
which has the same effect as the circuit interruption.

146

7. Symmetrical components in three-phase systems


Z1

ER

E1

ZS

I1
U1

UR

~
ES

Z2

ZS

I2

US

~
ET

U2

ZS
UT

Z0

I0
U0

Figure 7.10. Balanced three phase system with ungrounded generator


neutral point.

7.4.3

Wye connected load

A further neutral point of importance is that of a wye connected load, as


shown in figure 7.11, which shows a symmetrical load.
The resulting 120 System is given in the right side of figure 7.11. The load
impedances Z L of the wye connected load give rise to additional impedances
Z 1,L , Z 2,L , Z 0,L in parallel to the impedances discussed earlier, i.e. Z 1,S =
Z 2,S = Z 0,s = Z S . These additional impedances have all the same value as
the wye connected load impedances, i.e.
Z 1,L = Z 2,L = Z 0,L = Z L

(7.48)

It is of importance whether the neutral point of this wye connected load


is grounded or not. Again the positive and negative sequence systems are
identical in both cases, with or without grounded neutral point. The grounding or non-grounding affects only the zero sequence system. In the case of
grounding (dashed line), the impedance Z 0,L is connected at both ends; with
ungrounded neutral point the lower connections will be open.
Of course the load could also be delta connected. But, since a delta
connected load can be transformed into an ungrounded wye connected load,
this case is already covered by the above considerations.

7.4.4

Transformers

Transformers are very important for the analysis using symmetrical components and for the neutral point grounding of systems. Foremost the zero se-

147

7.4. 120 equivalent circuit


Z 1,S
ER

E1

ZS

Z 2,S

ZS

U1

I2

US

~
ET

Z 1,L

UR

~
ES

I1

Z 2,L

U2

ZS

ZL
ZL

UT
Z 0,S
ZL
Z 0,L

I0
U0

Figure 7.11. Three phase system with a wye connected load.

quence system is strongly dependent on the type of the transformer that is to


be connected. The positive and negative sequence systems are on the other
hand hardly affected. The process of the transformation of a three-phase
system including transformers into symmetrical components is as follows:
1. First of all, it is decided which voltage reference for the calculations
should be used (As an example, in case that a fault current is to be
calculated, of course suitable is the level where the fault occurs.).
2. All impedances and sources are referred to this voltage level.
3. The 120 system is derived without transformers. Thus at the place
where the transformer should be, there is now an interruption.
4. In the positive and negative sequence systems the appropriate transformer impedances are connected in series.
5. The appropriate transformer type with respect to the zero sequence
system is selected from figure 7.12. The zero sequence impedance for
transformers with ungrounded neutral is set to Z Ep = , and Z Es =
, respectively).
6. The appropriate zero sequence impedance is inserted into the circuit.
The calculation of the zero sequence impedance Z T 0 is rather complicated and the derivation is outside the scope of these lecture notes. The
reader is referred to Appendix A, where a somewhat deeper discussion is
provided. The approximate values are indicated in figure 7.12.

148

7. Symmetrical components in three-phase systems

Primary side

Nr.

Secondary side

Zero sequence system

Z T 0 + 3 Z Ep + Z Es
1

S
T

Z Ep

Z Es

ZT 0 3 5 Z1

ZT 0
2

ZT 0 Z1
R

ZT 0
S

3Z Ep
S
T

Z Ep

ZT 0 Z1

Figure 7.12. Zero sequence impedance of different three phase transformer connections.

149

7.5. Fault analysis using symmetrical components

In figure 7.13 a three phase system with a transformer is given. If one


follows the steps described above, one will get the 120 system shown in figure
7.14.
ER

ZS

~
ES

Transformer
Z T1 , Z T2 , Z T0
Z Es

ZS

UR

US

~
ET

ZS

UT

ZL
ZL

ZL

Figure 7.13. Three phase circuit with transformer.

In the following section different unbalances caused by ground faults and


short circuits will be examined more closely. It will be shown how these fault
conditions can be analyzed with the use of the 120 transformation.

7.5

Fault analysis using symmetrical components

In this section we will investigate some typical fault situations in electrical


power systems. We will neglect the transient events just after the fault and
assume that the fault currents are sinusoidal. We will see that in comparison
with classical network analysis using the physical phase representation the
calculation with symmetrical components introduces many simplifications.
In the examples only ohmic fault impedances will be used, but in the general
case complex impedances can also be introduced. In many cases a zero fault
impedance implies the most severe stresses.

7.5.1

Single-phase to ground fault

As single-phase, or sometimes also called single-line, to ground fault arises


when there is a low impedance connection between one of the three phases
and the ground.1 A cause for the fault could be that one conductor of an
1

If the neutral point of the feeding system is grounded with a low impedance, then
one speaks also of ground short-circuit. In case that different phases have ground contact
at different places, then depending upon the number of affected phases one has a double
or triple ground fault. Another type of fault is the two-phase short-circuit with earth
contact, which occurs if two phases have contact with the ground at the same location
and are short-circuited.

150

7. Symmetrical components in three-phase systems


Z 1,S

E1

Z T1

I1

Z 1,L

Z 1,S

U1

Z T2

I2

U2

Z 2,L

Z 0,S

Z 0,L

Z T0

I0

3Z Es

U0

Figure 7.14. 120 system with transformer.

overhead line has fallen down and touches the ground, or that a tree has
come in contact with one of the conductors. The single-phase ground fault
is the most frequent contingency in transmission grids. In high-voltage transmission grids with nominal voltages higher than 130 kV this fault constitutes
about 80% of all disturbances. Depending upon the type of disturbance the
fault resistance between the phase and the ground can vary significantly.
Generally each connection between a single phase and the ground can be
considered as single-phase unbalance, thus e.g. also a single-phase load.
The starting point of the analysis is the symmetrical grid on figure 7.8. In
case of a fault the grid is loaded additionally by a fault resistance R between
a phase and the ground (Figure 7.15). Thus an unsymmetrical system of
currents and voltages will occur, which with the use of the three symmetrical
components can be analyzed. The left side corresponds to the symmetrical
system, the right side represents the faulted part of the grid with a fault
resistance, R, between the phase R and the ground.
For the analysis of the unsymmetrical part of the circuit we will use
the right part of the circuit in figure 7.15. The unbalanced conditions will be
described as relations between the voltages U R , U S , U T and the currents I R ,
I S , I T , at the fault location. In order to describe the unbalance completely,
three equations are needed, which contain the six quantities or some of them.

151

7.5. Fault analysis using symmetrical components

E1
E2 = E0 = 0
Z 1, Z 2, Z 0

IR

UR

IS

US

If

IT

UT

Figure 7.15. Equivalent circuit of a single-phase ground fault.

In this case we have


R IR = U R

(7.49a)

IS = 0

(7.49b)

IT = 0

(7.49c)

Generally each unfaulted phase current I = 0. Each short-circuit to ground


implies U = 0. If a non-zero fault resistance exists, Ohms law must be
considered.
The equations are transformed now into the 120 system:
R (I 1 + I 2 + I 0 ) = U 1 + U 2 + U 0

a I1 + a I2 + I0 = 0
a I 1 + a2 I 2 + I 0 = 0

(7.50a)
(7.50b)
(7.50c)

By subtraction of the equation (7.50b) and (7.50c) as well as division by


(a a2 ) one obtains
I1 = I2
(7.51)
With the use of eq. (7.10) together with eq. (7.50b) one obtains
I1 = I2 = I0

(7.52)

R 3 I1 = U 1 + U 2 + U 0

(7.53)

and from the equation (7.50a)

The unsymmetrical load of the three-phase circuit is described completely


by the equations (7.52) and (7.53).
By the series coupling of the equivalent circuit of positive, negative and
zero sequence systems in figure 7.16 the above equations are fulfilled, i.e. the
same current flows through the three positive, negative and zero sequence
systems according to equation (7.52). According to equation (7.53) the sum

152

7. Symmetrical components in three-phase systems


Z1

Positive sequenceE 1

I1

U1

Z2

I2

Negative sequence

U2

Z0

Zero sequence

3R

I0

U0

Figure 7.16. 120 system equivalent circuit for a single-phase to ground


fault with a fault resistance R.

of the voltages of the three sequence systems equals the voltage drop over the
resistance 3R caused by the identical sequence currents. From the equivalent
circuit of the symmetrical components the identical currents of the three
symmetric component currents can now be calculated:
I1 = I2 = I0 =

E1
Z 1 + Z 2 + Z 0 + 3R

(7.54)

The voltage E 1 is the open-circuited voltage of the positive sequence system,


i.e. the voltage at the fault location prior to the fault. The voltages at the
fault location before the fault are E R , E S = a2 E R and E T = aE R , thus one
obtains for the internal voltage of the positive sequence system
1
E 1 = (E R + aE S + a2 E T ) = E R
(7.55)
3
The fault current I f is, as seen from figure 7.15, equal to the current in
phase R:
If = IR = I1 + I2 + I0
(7.56)
With this and with the equations (7.54) and (7.55) the fault current is given
by
ER
If = 1
(7.57)
3 (Z 1 + Z 2 + Z 0 ) + R

7.5. Fault analysis using symmetrical components

153

From the circuit in figure 7.16 the fault voltages can also be calculated.
One obtains
U 1 = E1 Z1I 1 = E1 E1

Z1
Z 1 + Z 2 + Z 0 + 3R

(7.58a)

U 2 = Z 2 I 2 = E 1

Z2
Z 1 + Z 2 + Z 0 + 3R

(7.58b)

U 0 = Z 0 I 0 = E 1

Z0
Z 1 + Z 2 + Z 0 + 3R

(7.58c)

From these equations the phase voltages of the RST system can be calculated. It is assumed that in the used three-phase system = applies (see
section 7.2), so that Z 1 = Z 2 . We now introduce Z:
Z1 = Z2 = Z

(7.59)

By means of equations (7.58a), (7.58b) and (7.58c) the phase voltages at the
fault location are calculated as
U R = U 1 + U 2 + U 0 = ER

3R
2Z + Z 0 + 3R

(7.60a)

U S = a2 U 1 + aU 2 + U 0 = E S E R

Z0 Z
2Z + Z 0 + 3R

(7.60b)

U T = aU 1 + a2 U 2 + U 0 = E T E R

Z0 Z
2Z + Z 0 + 3R

(7.60c)

From these equations one can easily recognize that the zero sequence impedance of the system has a substantial influence on the voltage of thehealthy,
i.e. unfaulted, phases (in this case phases S and T). Two cases will be discussed below.
First we assume that the neutral point is solidly grounded, i.e. no impedance in the ground connection, and that the line is short, whereby the
zero sequence impedance Z 0 is equal to the positive sequence impedance Z 1
and the negative sequence impedance Z 2 . A further simplifying assumption
comes from the fact that the fault resistance R is very low, which is the
case for solid short-circuits. With these assumptions (Z 0 = Z and R = 0)
equations (7.60a), (7.60b) and (7.60c) become
UR = 0

(7.61a)

U S = ES

(7.61b)

U T = ET

(7.61c)

154

7. Symmetrical components in three-phase systems


UT

ET

ET

UT
ER

ER

UR
US

ES

UR
US

ES

Figure 7.17. The three phase voltages during a single line to ground
fault. Left part: Z 0 = Z and R = 0. Right part: Z 0 = und R = 0.
(The dashed lines show the voltages prior to the fault, and the solid
lines the voltage phasors during the fault)

This means that with a ground fault in a system with solidly grounded neutral point, the voltages in the healthy phases are not affected. In a realistic
system with a solidly grounded neutral point some of the above assumption
might not be valid, but still the voltages in the unfaulted phases are affected
marginally.
The second case which is going to be investigated is the case of an ungrounded (isolated) neutral point (Z 0 = ). Thus the above equations
change to
UR = 0

(7.62a)

U S = ES ER

(7.62b)

U T = ET ER

(7.62c)

From the result one sees that with a single-phase to ground fault in a system
with ungrounded
neutral point the voltages of the healthy phases increases
with the factor 3. The figure 7.17 shows the voltages for both cases.
Unbalances of the sources, i.e. the generators, imply sources in the negative and zero sequence systems, in addition to the (desired) one in the
positive sequence system. Unbalances in the passive part of the system, i.e.
in impedances, loads, or faults, are considered by connections (couplings)
between the three equivalent circuits of the 120 system. It should be remembered that in a completely symmetrical (balanced) system the equivalent circuits of the symmetrical components are completely uncoupled. In
the above regarded case of a single phase fault in phase R, the introduced
unbalance, as given by equations (7.52) and (7.53), can be imposed by connecting the three circuits of the symmetrical components in series so that
the three currents I 1 , I 2 , and I 0 are equal. The value of the fault resistance
R must be multiplied by three in order to fulfill eq. (7.53).
From the three values in the 120 system, e.g. voltages U 1 , U 2 and U 0 ,
the corresponding values in the RST system, magnitude and phase, can be

155

7.5. Fault analysis using symmetrical components


calculated with the use of

UR
US = T
UT

the T -matrix

U1 + U2 + U0
U1
U 2 = a2 U 1 + aU 2 + U 0
U0
aU 1 + a2 U 2 + U 0

(7.63)

A single-phase fault in phase S or T can be analyzed in a similar way.


However the component systems must then be multiplied by a and/or a2 , in
order to be able to accomplish the series connection. A simpler way would
be to rename the phases, for instance to consider the phase S as phase R
for a single phase fault in phase S. Thus the phase R becomes phase T and
the phase T phase S. The designation of the phases is always a matter of
reference. One must just pay attention to that the correct order of the phases
is maintained after the phases have been renamed, i.e. that the phase shift
from phase R to phase S amounts to 120 .
The derived equivalent circuit for the single phase to ground fault will
also be used in the section 7.6 when resonant grounding of the neutral point
is discussed.

7.5.2

Two-phase short-circuit without ground contact

As a further unsymmetrical fault we will consider the two-phase shortcircuit, i.e. a low ohmic connection between two phases. If a ground contact
does also exist, then one speaks of two-phase ground fault.
We consider now a two-phase short-circuit without earth contact. The
circuit in figure 7.18 shows the asymmetrical fault with resistance R between
the phases S and T. Due to this fault unbalanced currents and voltages will
arise in the system.

E1
E2 = E0 = 0
Z 1, Z 2, Z 0

IR

UR

IS

US

IT

UT

If
R

Figure 7.18. RST equivalent circuit of a two-phase short-circuit without ground contact.

The assumption that the fault takes place between the phases S and T
simplifies the calculation compared with the other two possibilities. They

156

7. Symmetrical components in three-phase systems

differ however only in phase shifts a and a2 , respectively. As in a singlephase ground fault, in case of a short-circuit between the phases R and S or
R and T, the phases can also here be renamed.
With a resistance between the phases S and T the unsymmetrical conditions in the RST system are given by:
IR = 0

(7.64a)

I S = I T

(7.64b)

U S U T = R IS

(7.64c)

After the transformation into symmetrical components we obtain


I1 + I2 + I0 = 0
2

a I 1 + aI 2 + I 0 = aI 1 + a I 2 + I 0


a U 1 + aU 2 + U 0 aU 1 + a2 U 2 + U 0 = R a2 I 1 + aI 2 + I 0
2

(7.65a)
(7.65b)
(7.65c)

From equation (7.65a) follows

I 0 = I 1 I 2

(7.66)

which inserted in equation (7.65b) yields


I 1 = I 2

(7.67)

Together with equation (7.65a) follows


I0 = 0

(7.68)

One could have written the latter equation directly, because a zero sequence
current always flows through a possible connection between the neutral point
and the ground. In the case of a two-phase short-circuit there is no closed
current loop through the neutral point and the fault impedance. The current
paths I R , I S and I T do not have any possibility (in contrast with the singlephase failure case) to get closed through the ground and therefore they
do not contain any zero sequence component. The above properties of the
symmetrical component currents can now be used in eq. (7.65c) together
with the properties of the constant a giving:
I0 = 0

(7.69a)

I 1 = I 2

(7.69b)

U 1 U 2 = R I1

(7.69c)

As in the case of single-phase ground fault the above conditions can be


represented by suitable connection of the symmetrical component systems.
The equivalent circuits of the positive and negative sequence systems must

157

7.5. Fault analysis using symmetrical components


Z1

Positive sequence E 1

I1

U1
R

Z2

I2

Negative sequence

U2

Z0

I0

Zero sequence

U0

Figure 7.19. Equivalent circuit of the symmetrical components for a


two-phase short-circuit without ground contact.

now be connected crosswise, so that the current flows in opposite directions


in the two systems. Between them the resistance R should be connected
(this time without factor 3), so that the Ohms law according to equation
(7.69c) is satisfied. The zero sequence system is not connected at all in this
case, and is therefore completely insignificant for the further considerations.
Also in this case the voltages and currents could be calculated in symmetrical components, which with the aid of the matrix T can be transformed
back again into the RST system. For the currents in positive and negative
sequence systems applies:

I 1 = I 2 =

E1
Z1 + Z2 + R

(7.70)

with
1
E 1 = (E R + aE S + a2 E T ) = E R
3

(7.71)

158

7. Symmetrical components in three-phase systems

and (7.70) the fault current I f can be calculated as


If

= I S = a2 I 1 + a I 2 + I 0
= (a2 a) I 1 = (a2 a)

= j 3

7.5.3

(7.72)
E1
Z1 + Z2 + R

ER
Z1 + Z2 + R

(7.73)
(7.74)

Three-phase short-circuit

Contrary to the fault situations described above, the three-phase shortcircuit is symmetrical in all phases, if we assume the same fault resistance
in the three phases. The figure 7.20 shows the faulted circuit in the RSTsystem.

E1
E2 = E0 = 0
Z 1, Z 2, Z 0

IR

UR

IS

US

IT

UT

Figure 7.20. Three-phase system of a three phase short-circuit.

This kind of short-circuit arises very rarely, but normally causes the
highest short circuit currents (and thus the highest thermal and mechanical
stresses) and is used as worst case scenario in short-circuit computations.
In the here considered example, it is assumed that a fault resistance R exists
between the phases and the ground. As symmetrical conditions do prevail,
the currents and the voltages in case of the three-phase short-circuit with
ground contact do not differ from the case without earth contact. (The
point, at which the resistances R in figure 7.20 are connected together and
grounded, would have zero voltage even without connection to the ground.)
For the voltages and currents in the RST system applies
U R = R IR

(7.75)

U S = R IS

(7.76)

U T = R IT

(7.77)

159

7.5. Fault analysis using symmetrical components


Z1

Positive sequence E 1

I1

Z2

U2

U0

I2

Negative sequence

Z0

U1

I0

Zero sequence

Figure 7.21. Equivalent circuit in symmetrical components of a three


phase short-circuit.

Transformation into symmetrical components and some algebra yields:


U 1 = R I1

(7.78)

U 2 = R I2

(7.79)

U 0 = R I0

(7.80)

The equivalent circuit in symmetrical component is simple, since the


symmetrical component systems are not connected in the equivalent circuit and only the positive sequence system is activated. The corresponding
circuits in symmetrical components is given in figure 7.21.
The fault currents in the RST system are
IR =
IS =
IT =

ER
Z1 + R
a2 E R
Z1 + R
aE R
Z1 + R

(7.81)
(7.82)
(7.83)

Comparing with (7.57) one can see that if the positive, negative and zero
sequence impedances all are equal, the same fault current as in the case of

160

7. Symmetrical components in three-phase systems

the single-phase ground fault results, i.e. I f = I R . Since the system in this
case is balanced also during the fault, no zero sequence current flows through
the neutral points.

7.6

Resonant Grounding

There are certain advantages with an electric power system , which has
ungrounded neutral points when single-phase to ground faults occur: Since
no return path for the short-circuit current exists no fault current will flow
(see equation (7.57)). This can also be seen in the equivalent circuit in figure
7.16, whose zero sequence system would be open if the neutral points are
left ungrounded. Thus, besides I 0 also I 2 vanishes, i.e. the currents remain
symmetrical since only the positive sequence system is active. However, the
phase voltages become unsymmetric. Due to the fact that one phase is forced
to ground potential by the fault,
the voltages of the two other phases increase
relative to ground by the factor 3 (see equations (7.62a)-(7.62c) and figure
7.17). Medium voltage grids are usually operated ungrounded in Switzerland.
If the grid has a small geographic size the short-circuit current is also small
and the corresponding arc at the overhead line insulator is often extinguished
automatically. The voltage increases for the healthy phases are acceptable
and are also used as a detection of single-phase to ground faults. Different
countries have different philosophies and practices concerning grounding of
low and medium voltage grids.
In an electric power system with grounded or low-ohmic grounded neutrals a single-phase to ground fault implies a large short-circuit current,
possibly causing damages, and requires isolation of the fault e.g. through
line tripping. All electric power systems with power lines or cables are actually grounded, if not over a direct connection (i.e. grounded neutrals) then
over the shunt capacities between the power lines and ground. Since high
and extra-high voltage grids usually cannot withstand the voltage increases
incurred by single-line to ground faults in ungrounded systems, the neutrals
in these grids must be grounded through low-ohmic impedances. As a consequence, the high fault currents caused by a single-phase to ground fault
has to be extinguished by action of protections and circuit breakers through
tripping of lines or other pieces of equipment where the fault actually occurs.
In circuits with low ohmic neutral grounding connections (effectively
grounded) the neutrals of individual or all transformers are in practice
grounded either directly or over a low ohmic impedance. The cost for the
lower short-circuit currents in ungrounded or high impedance grounded grids
is the voltage increase in the unfaulted phases in connection with single-line
to ground faults. The insulation has to withstand these higher voltages. The
following three options are the ones used in practice in low and medium
voltage grids:

161

7.6. Resonant Grounding

The neutrals of all transformers are grounded directly. In this case the
highest short-circuit currents occur.
Only selected neutrals are grounded directly. The other neutrals are
ungrounded.
In order to decrease the short-circuit current in the case of singlephase to ground faults the neutral points are not grounded directly,
but through impedances.
Figure 7.22 shows a symmetrical three-phase power transmission system
with ungrounded neutral. The capacitances represent the (identical) capacitances between the conductors of the power line and ground. A single-phase
UR

R
IR

IS
S

IS

IT

IR

N (G)

UT

IT

US

CE

Figure 7.22. The power system in balanced operation; The capacitances CE corresponds to the capacitances between the conductors and
ground.

to ground fault on any phase will result in an unbalanced situation. For the
following considerations it is assumed that the single-phase to ground fault
appears on the phase R, as shown in figure 7.23. Additionally it is supposed
that the fault resistance is zero, so that the faulty phase is at ground potential at the fault location. This shift of the ungrounded neutral affects of
course also the two other phases S and T (see the voltage phasor
diagram in

figure 7.23). The voltages are approximately by a factor of 3 larger than in


the stationary conditions, and they are also phase shifted. The fault current
in figure 7.23 is often sufficient to maintain an arc at the fault position.
In order to further reduce the fault current, a coil (inductance) can be
inserted between the neutral point and ground, which is known as a Petersen
coil (figure 7.24).
If a single-phase ground fault occurs, part of the short-circuit current
flows through the Petersen coil. The current through the coil lags the voltage
with 2 , but the current through the earth capacitances leads the voltage

162

7. Symmetrical components in three-phase systems

If

R
G
S
IS

IS

N
T

IT

If
IT

U TR

U SR

CE
IS + IT

Figure 7.23. The unbalanced situation of the power system during a


single-line to ground fault.
ET

~
ES

~
ER
LP

~
CE

Figure 7.24. Petersen coil in a three-phase grid.

with 2 . Thus the two current contributions in the arc are phase shifted by
with respect to each other and by a suitable choice of the inductance
of the Petersen coil they will cancel out, resulting in a zero fault current.
The size of the Petersen coil must thus be selected in such a way that it
forms a parallel circuit together with the ground capacitances, which is in
resonance at the grid frequency. In this case no current can flow through the
zero sequence system and, consequently, not through the fault resistance R
in the ideal case, either.
In reality a small current will flow through the ground fault despite the
Petersen coil, since the real, not ideal coil has a resistance in addition to
the inductance. Furthermore, shunt resistances do exist in the grid, whereby
both the inductive and the capacitive currents are not purely reactive cur-

163

7.6. Resonant Grounding

rents and the capacitive and inductive currents are not exactly phase-shifted
by . Thus, the resonant circuit does not block the current completely, but
a small resistive current remains. This resistive current lies in real systems
in the order of magnitude of 3 to 10% of the uncompensated ground fault
current. This current is usually too small to maintain an arc for single-phase
to ground faults, which are as already mentioned the most common type of
faults in power systems.
The inductance of the neutral point connection LP , as will be derived
in the next section, must fulfill
LP =

1
3 2 CE

(7.84)

where CE is the capacitance between the conductors and ground in order


to achieve the desired resonant conditions. If a network is grounded with a
coil with an inductance according to eq. (7.84), this network is said to have
resonant grounding.
Besides the resistive elements in the resonant circuit, there is a further
complication and that is the determination of the effective ground capacitances. The larger a geographical area the system covers, the more the
distributed structure of the lines must be taken into account and the approximation with one single capacitance gets less accurate. Hence, in large
systems the Petersen coil will still limit the short-circuit current, but to a
value which can be detected by the line protection system and the fault can
be isolated quickly.
Derivation of resonant grounding conditions The left part of figure 7.25
(RST) shows a single-phase to ground fault in a system with a solidly
grounded neutral point. The flow of the short-circuit current is shown by
the bold lines in the figure. As explained above a fairly high short-circuit
current flows in such a system and the arc at the fault location is fully
developed. The aim of the Petersen coil (and thus of the resonant grounding) consists in bringing the fault current to zero and consequently prevent
the development of an arc. In practice however, the current will never be
completely extinguished due to various reasons as explained above.2
The right part of figure 7.25 shows the grid transformed into symmetrical
components during a single-phase to ground fault and Petersen coil in the
zero sequence system, and from this the conditions for resonant grounding
can be derived. (Note that the inductance of the Peterson coil in the zero
sequence system is 3LP .)
The total reactance in the zero sequence system consists of the sum of
the zero sequence impedance of the grid and the impedance of the Petersen
2

Since only a small short-circuit current flows through the fault, the power grid can thus
be operated for longer time with the ground fault without incurring any severe damages.

164

7. Symmetrical components in three-phase systems


I 1f

Z1

ET
E1

1
jCE

ES

I 2f

Z2

~
1
jCE

ER
LP

Z0

I 0f

CE
3jLP

U 0f

1
jCE

120

RST

Figure 7.25. Derivation of the inductance of the Petersen coil for


resonant grounding conditions.

coil in parallel with the impedance of the line capacitance:


Z total = (Z 0 + 3jLP )

1
jCE

(7.85)

Assuming that 3jLP Z 0 , an approximative value of LP can easily be


calculated:
Z total 3jLP

1
3jLP jC
1
3jLP
E
=
=
1
jCE
3 2 LP CE + 1
3jLP + jCE

(7.86)

The aim is to eliminate the fault current, which means


!

I 1f = I 2f = I 0f = 0

(7.87)

which means that3


I 0f =

U 0f
1 3 2 LP CE !
= U 0f
=0
Z total
3jLP

(7.88)

The fault current vanishes if the numerator in equation (7.88) is zero. Consequently the inductance of the Petersen coil should be
LP =
3

1
3 2 CE

(7.89)

Of course one could formulate a condition for the blocking fault current: In resonance
the total impedance the fault current sees becomes infinitely large, i.e. the denominator
in equation (7.86) should be zero. From this one obtains the same result.

165

7.6. Resonant Grounding


2

U1

Linear network

U2

Figure 7.26. Linear model of passive network elements of a power grid.

Finally, it must be verified whether the assumption 3jLP Z 0 is fulfilled.


For a line 100 km long the ground capacity amounts to approximately 2 F
and the zero sequence impedance to (30+120j) . Since the Petersen coil can
compensate only the capacitive part of the fault current, only the reactance
of the zero sequence impedance is of interest, i.e. 2122j. It is clear that
2122j 120j, and the above assumption made is consequently justified.
For this derivation we have modelled the ground capacitances of the
conductors as concentrated elements CE . In reality the capacitances are distributed over the whole length of the power lines and cables, the size of
the lumped elements corresponds to the sum of all capacitances over of the
line(s). A relevant question is, how the impedance between a conductor and
the ground varies, in case that one moves along the line. In the following it
will be shown that this impedance, if it is infinitely large at one point, then
for other points in the grid it will not change.
Consider the linear circuit in figure 7.26. The points marked 1 and 2
denote two different locations on a certain phase, e.g. the R phase, in the
grid. This linear circuit can be regarded as a two-port equivalent circuit and
can thus be modelled by a -equivalent (see figure 7.27). We assume that the
linear circuit is grounded through a Petersen coil and thus for the location
1 resonant conditions prevail, whereby the admittance between phase and
earth at this location (Y 1 in figure 7.27) is equal to zero. It will now be
shown that, provided two conditions are fulfilled, as will be explained below,
the admittance seen at location 2, i.e. Y 2 , will also be zero. The conditions
z

Y1

y1

y2

Y2

Figure 7.27. Linear network model of the system in figure 7.26 as a


two-port (-equivalent).

166

7. Symmetrical components in three-phase systems

that must be fulfilled are:


1. A voltage E 6= 0 applied at location 1 will give rise to a non-zero
voltage at location 2.
2. A voltage E 6= 0 applied at location 2 will give rise to a non-zero
voltage at location 1.
In a power system this is fulfilled since a voltage applied between one phase
and the ground in any point of the network will of course result in non-zero
voltage between this phase and ground. (It is assumed that the network is
passive, i.e. the generators are not connected.) The theorem we intend to
prove can thus be summarized as

Y1 = 0

U 1 = E(6= 0) U 2 6= 0
Y2 =0
(7.90)

U 2 = E(6= 0) U 1 6= 0

This means that, if the admittance is zero for one location, it will also be
zero for all other locations fulfilling the above, i.e. in practice at all locations
in the system.
To prove this we start with the calculation of the admittances Y 1 and
Y 2 of the circuit in figure 7.27:
y 1 + y 2 + z y 1 y2
1
=
z + 1/y 2
1 + z y2
y 1 + y 2 + z y 1 y2
1
Y 2 = y2 +
=
z + 1/y 1
1 + z y1
Y 1 = y1 +

(7.91a)
(7.91b)

A voltage applied at port 1 and port 2, respectively, will then give rise to
the following voltages at the other port, which are non-zero according to the
above stated conditions. This yields
E
6 0
=
1 + z y2
1
6 0
=

1 + z y2

U1 = E U2 =

E
6 0
=
1 + z y1
1

6 0
=
1 + z y1

(7.92a)

U2 = E U1 =

(7.92b)

Obviously the fractions in equations (7.92a) and (7.92b) are non-zero. As


we assumed that we had resonant grounding conditions as seen in location,

167

7.6. Resonant Grounding

this means that admittance in equation (7.91a), i.e. the admittance between
a phase and ground, which can be written as
Y1 =

y1 + y 2 + z y 1 y 2
=0
1 + z y2

(7.93)

Equation (7.93) together with equation (7.92a) implies


y1 + y2 + z y1 y2 = 0

(7.94)

Comparing this result with the expression for the admittance at location 2
in equation (7.91b), we see directly that
Y2 =

y1 + y 2 + z y 1 y 2
=0
1 + z y1

since
y 1 + y 2 + z y1 y 2 = 0 and

(7.95)

1
6= 0
1 + z y1

This means that the admittance at location 2 also is also equal to zero,
and thus for a linear passive power system with resonant grounding the
impedances between a phase and ground at all locations of system are the
same and consequently infinite.
In a real system where the losses are considered as well, the admittance
will not be identically zero. Furthermore, switching actions in the grid will
change the total capacitance, and the resonant circuit will in practice be
detuned. The compensation is sufficient, however, to keep the ground fault
currents to relative small values (e.g. 50 A for medium voltage level) so that
equipment will not be destroyed and system operation could be maintained
during single-phase to ground faults.

168

7. Symmetrical components in three-phase systems

Appendix A
Zero sequence impedance of transformers
The zero sequence impedance in a three-phase system is generally defined as
the ratio between phase voltages and phase currents when all three phases
are connected, see figure A.1. For the system in figure A.1 the zero sequence
impedance is thus defined as
Z0 =

U0
I0

(A.1)

In transformers the zero sequence impedance is finite only if the windings


are wye connected1 . For delta connected windings there is no neutral point
and thus no possible return path for a zero sequence current, which means
that Z 0 .
If one wants to measure the zero sequence impedance of a transformer
winding, then a voltage will be applied between the short-circuited phases
R, S and T and the neutral conductor as shown in figure A.1. The other
winding(s) (primary, secondary, tertiary) should be open. Depending on the
design of the transformer a zero sequence current will flow, from which, together with the applied voltage, one can calculate the zero sequence impedance
according to the equation (A.1).
The following considerations refer to three-phase transformers with wye
connected windings. First a transformer where both windings (primary and
1
Transformers with Z-connected windings can also have non-zero zero sequence
impedances, but they are not considered here.

I0

Phase R

I0

Phase S

I0

Phase T

3ph. System

U0
3I 0 neutral conductor

Figure A.1. Equivalent for the zero sequence impedance of a threephase system.

169

170

A. Zero sequence impedance of transformers


Primary winding
Yoke

Limbs

Secondary winding

Figure A.2. Three-phase transformer with five limbs.

secondary) are wye connected is considered. The three equal voltages in the
three phases will cause three equal magnetic fluxes in all three limbs of the
transformer. These flows add to non-zero in the yoke. As return path this
resulting flux needs a magnetic connection outside the windings of the three
phases. In transformers with additional limbs, these will provide the return
path, and an example is given in figure A.2 where a five limb transformer is
shown.
In a three limb transformer there exists no return path in the iron core
for the magnetic flux. The return flux must thus flow outside the iron core,
i.e. through the air and transformer tank. Such a return path has a high
reluctance, which implies that a high magnetizing current is needed to drive
the flux. Hence it follows that the zero impedance of a three limb threephase transformer is relatively low. If a three limb transformer is fed by a
zero current of the amount of the nominal current, then the zero voltage lies
normally in the range 60. . . 100 % of the nominal voltage.
In a five limb transformer a return path is possible via the side limbs.
As these side limbs are practically reluctance free in unsaturated conditions,
even small currents will create a considerable voltage between the phase
conductors and the neutral. The voltage and current quotient reaches very
high values, i.e. the zero sequence impedance of the transformer is very
large. For large zero currents, saturation appears in the side limbs whereby
the reluctance decreases. Thus the voltage and current quotient decreases,
the zero sequence impedance is thus not constant.
If the transformer has delta connected windings, then these represent a
short-circuit path for the induced zero voltage (see figure A.3). A counter
current will be induced which will correspond to the current in the wye connected windings. The total zero impedance of such a transformer connection
consists of two components:
1. The zero impedance of the wye connected winding without the delta

171

U0

U0

Figure A.3. Wye and delta connected windings. For the wye connection the zero voltages in the three phases add themselves to zero and
can be regarded as short-circuited.

winding, plus
2. The mutual impedance between the two windings.
Figure 7.12 shows the zero impedances for different transformer connections. The value of the zero impedance Z 0 is not the same for all connections
no. 13. For wye-wye configurations like in no. 1 the zero impedance is essentially dependent on the existence of the magnetic return path and on the
grounding impedances. For the case no. 1 the zero impedance consists of the
grounding impedances together plus the impedance between the two windings (the leakage reactance Z ). Since the triple zero current flows through
the neutral points impedances, these must be multiplied by three in the zero
sequence scheme. In case no. 2 no zero currents flow in the windings because
all three connection points R, S and T have the same zero voltage U 0 . Thus
the zero impedance becomes infinitely high. For connection no. 3 the zero
impedance results again from the coupling of the leakage fluxes and the neutral point impedance of the primary winding. This applies also to the case
when the primary and secondary windings are reversed.

172

A. Zero sequence impedance of transformers

Appendix B
Zero sequence impedance of overhead
lines
The basic equations for the calculation of the zero impedance of overhead
lines can be found in [20, 21, 22]. The derivation of these equations is relatively complex and will not be discussed here in detail. Instead the fundamental effects will be qualitatively discussed and some important results
will be highlighted.

B.1

Impedance of a single line with return current


through ground

We consider the circuit with ground return shown in the figure B.1. The
conductor with diameter 2r runs parallel to the earths surface level (y = 0)
at the height of y = h. For the resistivity of the ground an effective average
value can be assumed. The zero current I 0 in the conductor returns through
the ground, and accordingly a certain current density distribution arises.
Now one can assume that the ground return current is a concentrated current
at the depth y = . This equivalent depth of the return current depends
on the frequency and on the resistivity of the ground:
=
y

3.42
0

I0

(B.1)

2r
h

U0
0

Figure B.1. A conductor with return current through the ground.

173

174

B. Zero sequence impedance of overhead lines


Table B.1. Specific earth resistance and earth current depth for different soil types. [23].

Type of
ground
in m
in m

Moor
land
30
510

Clay and
farm land
100
930

Wet
sand
200
1320

Dry
sand or gravel
1000
2940

Stony
land
3000
5100

With = 2f as the frequency of the alternating current in rad/s and


0 = 4 107 Vs/Am is the permeability of the vacuum. With in m,
is given in m. It is seen that the earth current depth
decreases with increasing frequency and
increases with increasing ground resistivity.
The table B.1 indicates some typical values for the earth current depth as
function of the specific earth resistance at f = 50 Hz. In the northern preAlps the specific earth resistance is typically between 100 and 200 m. It
should be noted that metallic pipelines, fences, and other structures in the
ground can substantially affect the effective earth resistance as well as the
earth current depth.
Assuming h (see table B.1) a simplified expression for the impedance
per length unit can be given as [23]:
Z LE = R +

0
0
+j
8
2

ln

r
+
r
4

(B.2)

This equation is valid in the range 50 Hz to 20 kHz with normally sufficiently


good accuracy. The individual terms in (B.2) have the following meaning:
R is the effective resistance of the conductor per length unit.
0
8

corresponds to the effective resistance of the ground return path


per length unit. It should be noted that this expression is independent
of the ground resistivity. This is a consequence of the assumption h
, which according to (B.1) implies a high or a low conductivity of
the ground.

0
2

ln r corresponds to the outer reactance of the circuit per unit of


length.

0
2

r
4

is the inner reactance of the conductor per unit of length.

175

B.2. Zero sequence impedance of three-phase overhead lines

B.2

Zero sequence impedance of three-phase overhead lines

The zero impedance of three-phase overhead lines is strongly dependent on


whether the line has ground wire(s) or not1 . As shown in figure B.2 (lower
part) the return current is divided between the ground wire (I S ) and to the
ground (I E ) when ground wire(s) is (are) installed.
Zero sequence impedance without ground wire Figure B.2 (upper part)
shows a three-phase overhead line without ground wire. Each of the currents produces a magnetic flux linkage, which (partly) penetrates also the
other two conducting circuits. With consideration of this mutual inductive
coupling we obtain
U 0 = I 0 Z 0RR + I 0 Z 0RS + I 0 Z 0RT

(B.3a)

U 0 = I 0 Z 0SR + I 0 Z 0SS + I 0 Z 0ST

(B.3b)

U 0 = I 0 Z 0T R + I 0 Z 0T S + I 0 Z 0T T

(B.3c)
(B.3d)

Here the Z 0RR = Z 0SS = Z 0T T are the circuit impedances according to


equation B.2. The impedances Z 0RS , Z 0RT , . . . , Z 0T S represent the mutual
impedances between two conductors with return current through ground
and can be approximately calculated as follows [23]:
Z nm =

0
0

+ j
ln
8
2 dnm

(B.4)

Here Z nm is the mutual impedance per length unit between the phases n and
m; dnm is the distance between the conductors. Thus the zero impedance
(per length unit) of a three-phase overhead line without ground wire results
to


0
0

Z 0 = R0 + jX0 = R + 3
+ j
3 ln
+
(B.5)
3
8
2
4
rD 2
where D is the geometric average distance between the three phases.
Zero sequence impedance with ground wire If the power line has ground
wire(s), then the coupling impedance between the phase conductors and the
ground wire needs also to be taken into account. Appropriate equations are
given in [23].
1
Primarily ground wires serve as shielding for lightning. Apart from that, however,
they still perform other tasks: Through the ground wire a parallel connection of the tower
grounding is established, which can lead to improved grounding conditions. In addition a
ground wire reduces the magnetic field from the line.

176

B. Zero sequence impedance of overhead lines

I0

I0

I0

U0

3I 0

IS
I0

I0

I0

U0

3I 0

IE
Figure B.2. Three-phase overhead line without (upper part) and with
(lower part) ground wire.

B.2. Zero sequence impedance of three-phase overhead lines

177

Table B.2. Typical values for the positive and zero sequence
impedances of high voltage overhead power lines, one three phase system per tower, one ground wire [23].

Voltage
(kV)
110
220
380

Bundle
conductor
no
2
4

Positive sequence
Z 1 in /km
0.12 + j0.39
0.06 + j0.30
0.03 + j0.26

Zero sequence
Z 0 in /km
0.31 + j1.38
0.16 + j0.98
0.13 + j0.91

Apart from the existence of a ground wire, the zero impedance of a line
is determined particularly by the tower design, i.e. geometrical configuration
of the phase conductors and the ground wires at the tower and number of
three-phase lines on the tower, and the conductivity of the ground. Table B.2
gives some typical values for zero sequence impedances of overhead lines.

178

B. Zero sequence impedance of overhead lines

Bibliography
[1] Bergen, A. R. ; Vittal, V.: Power Systems Analysis. 2nd edition.
Prentice-Hall, 2000. ISBN 0-13-691990-1
[2] Oeding, D. ; Oswald, B. R.: Elektrische Kraftwerke und Netze. 6.
Aufl. Springer, 2004. ISBN 3-540-00863-2
[3] Crastan, V.: Elektrische Energieversorgung 1. Bd. 1. Springer, 2000.
ISBN 3-540-64193-9
[4] International Energy Agency. Key World Energy Statistics 2007
[5] Chow, J. ; Kopp, R. J. ; Portney, P. R.: Energy Resources and
Global Development. In: Science Vol. 302 (2003), S. 15281531
[6] Organisation for Nordic Electrical Cooperation (Nordel).
Nordel Annual Statistics 2007
[7] Weingartner, H.: Wasserkraft Energie mit Zukunft. In: VA Tech
Direkt 01/2005 (2005)
[8] Schweizerisches Bundesamt f
ur Energie. Elektrizit
atsstatistik
2007
[9] International Energy Agency. Renewables Information 2003
[10] Kundur, P.: Power system stability and control. McGraw-Hill, 1994.
ISBN 0-07-035958-X
[11] Glavitsch, H.: Leitungstheorie. AMIV-Verlag, 1988
[12] Haubrich, H.-J.: Aachener Beitr
age zur Energieversorgung. Bd. 13:
Elektrische Energieversorgungssysteme technische und wirtschaftliche
Zusammenh
ange. Verlag der Augustiner Buchhandlung, 1993
[13] Weedy, B. M. ; Cory, B. J.: Electric power systems. 4th edition.
Wiley, 1998. ISBN 0-471-97677-6
[14] StClaire, H. P.: Practical concepts in capability and performance of
transmission lines. In: AIEE Transactions 72 (1953), S. 11521157
179

180

BIBLIOGRAPHY

[15] Handschin, E.: Elektrische Energie


ubertragungssysteme. 2. Auflage.
H
uthig, 1987. ISBN 3-7785-1401-6
[16] Heuck, K. ; Dettmann, K.-D. ; Reuter, E.: Elektrische Energieversorgung; Erzeugung, Transport und Verteilung elektrischer Energie f
ur
Studium und Praxis. 5. Auflage. Vieweg, 2002. ISBN 3-528-48547-7
[17] Flosdorff, R. ; Hilgarth, G.: Elektrische Energieverteilung. 8. Auflage. Teubner, 2003. ISBN 3-519-26424-2
[18] Fortesque, F. L.: Method of symmetrical co-ordinates applied to the
solution of polyphase networks. In: 34th Annual Convention of the
AIEE. Atlantic City, NJ, 1918, S. 10271115
[19] Anderson, P. M.: Analysis of faulted power systems. IEEE Press,
1995. ISBN 0-7803-1145-0

[20] Pollaczek, F.: Uber


das Feld einer unendlich langen wechselstromdurchflossenen Einfachleitung. In: Elektische Nachrichtentechnik 9
(1926), Nr. 3, S. 339359
[21] Carson, J. R.: Wave Propagation in Overhead Wires with Ground
Return. In: Bell System Technical Journal 5 (1926), S. 539554
[22] Horak, J.: Zero Sequence Impedance of Overhead Transmission
Lines. In: 32nd Annual Western Protective Relay Conference (WPRC).
Spokane, USA, 2005
[23] Happoldt, H. ; Oeding, D.: Elektrische Kraftwerke und Netze. 5.
Auflage. Springer, 1978. ISBN 3-540-08305-7
[24] Balzer, G. [u. a.]: Switchgear Manual. 9th edition. ABB Calor Emag
Schaltanlagen AG, Mannheim, Germany; Cornelsen Verlag, Berlin,
1995. ISBN 3-464-48234-0

You might also like