Sunteți pe pagina 1din 310

University of Wollongong Thesis Collections

University of Wollongong Thesis Collection


University of Wollongong Year 

Contributions towards the development


of the Technical Report IEC/TR
61000-3-13 on voltage unbalance emission
allocation
Prabodha Paranavithana
University of Wollongong

Paranavithana, Prabodha, Contributions towards the development of the Technical Report


IEC/TR 61000-3-13 on voltage unbalance emission allocation, PhD thesis, School of Elec-
trical, Computer and Telecommunications Engineering, University of Wollongong, 2009.
http://ro.uow.edu.au/theses/834

This paper is posted at Research Online.


http://ro.uow.edu.au/theses/834
Contributions Towards the Development of the
Technical Report IEC/TR 61000-3-13 on Voltage
Unbalance Emission Allocation

A thesis submitted in fulfilment of the

requirements for the award of the degree

Doctor of Philosophy

from

University of Wollongong

by

Prabodha Paranavithana, BSc(Eng)

School of Electrical, Computer and Telecommunications

Engineering

March 2009
Dedicated to my parents...
Acknowledgements

It is a pleasure to be able to thank many people to whom I am indebted for the

development of this thesis.

First and foremost, I wish to express my utmost gratitude to my principal super-

visor, Associate Professor Sarath Perera of the University of Wollongong (UoW), for

enabling me to pursue postgraduate studies at the University of Wollongong and the

support given throughout the study period in many ways. Your dedication, patience,

knowledge and experience could not have been surpassed. I admire your guidance

towards growing me up academically and personally over last few years.

Thanks to my co-supervisor, Professor Danny Sutanto of the UoW, for the assis-

tance provided. I would also like to offer many appreciations to Dr. Duane Robinson

of Beca, Australia for proofreading this thesis. To Mr. Robert Koch of Eskom

Holdings Limited, South Africa and Dr. Zia Emin of National Grid Electricity Trans-

mission, United Kingdom go many thanks for their insightful technical contributions

and helpful attitude. LATEX assistance received from Dr. Timothy Browne, previ-

ously with the Integral Energy Power Quality and Reliability Centre (IEPQRC) at

the UoW, is much appreciated.

Funding for this project was provided by SP AusNet, Victoria and the IEPQRC.

I am grateful to Mr. Dhammika Adihetti, Mr. Shiva Bellur and Mr. Sanath Peiris of

SP AusNet for arraigning this. Many thanks to Mr. Jeff Sultana, Mr. Shem Cardosa

and Mr. Mahinda Wickramasuriya of SP AusNet for the support given in collecting

the required data for Chapter 7 of this thesis.

Thanks to Dr. Vic Smith and Sean Elphick of the IEPQRC who have graciously

responded to many administrative and software related requests. My thanks also go to

Roslyn Causer-Temby of the School of Electrical, Computer and Telecommunications


iii
iv

Engineering (SECTE) at the UoW, Tracey O’Keefe and Maree Burnett who are

former members of the SECTE staff, and Esperanza Riley of the IEPQRC for solving

many administrative problems and providing perspective. The SECTE workshop

staff have cheerfully provided the technical assistance.

Very special thanks go to my friend Dr. Sankika Tennakoon, previously with the

IEPQRC, for being generously supportive especially during hard times along the way.

Your contribution to my PhD experience is also appreciated.

My heartiest gratitude goes to my parents Mithrananda and Manike for all encour-

agements, guidance and sacrifices made on behalf of me to come this far. Finally, my

thanks go to the rest of my family and friends particularly Pinky, Dimuthu, Radley,

Matthew and Nishad for being supportive in many ways.


Certification

I, Prabodha Paranavithana, declare that this thesis, submitted in fulfilment of the

requirements for the award of Doctor of Philosophy, in the School of Electrical, Com-

puter and Telecommunications Engineering, University of Wollongong, is entirely my

own work unless otherwise referenced or acknowledged. This manuscript has not been

submitted for qualifications at any other academic institute.

Prabodha Paranavithana

Date: 31 March 2009

v
Abstract

Although voltage unbalance is a well understood concept, its presence as a power

quality problem in electricity transmission and distribution networks has continued

to be an issue of concerns primarily due to difficulties found by some network service

providers in maintaining acceptable levels. This emphasises the lack of recommenda-

tions on engineering practices governing voltage unbalance that would facilitate the

provision of adequate supply quality to connected customers.

The International Electrotechnical Commission (IEC) has recently released the

Technical Report IEC/TR 61000-3-13 which provides guiding principles for coordi-

nating voltage unbalance between various voltage levels of a power system through

the allocation of emission limits to installations. Although the IEC report is based

on widely accepted basic concepts and principles, it requires refinements and original

developments in relation to some of the key aspects. This thesis primarily focuses on

making contributions for further improvements to the IEC report so as to present a

more comprehensive voltage unbalance allocation procedure.

Similar to the counterpart IEC guidelines for harmonics (IEC 61000-3-6) and

flicker (IEC 61000-3-7) allocation, IEC/TR 61000-3-13 also apportions the global

emission allowance to an installation in proportion to the ratio between the agreed

apparent power, and the total available apparent power of the system seen at the

busbar where it is connected. However, noting that voltage unbalance at a busbar

can arise as a result of both load and system (essentially lines) asymmetries, IEC/TR

61000-3-13 applies an additional factor which is referred to as ‘Kue’ to the appor-

tioned allowance. This factor Kue represents the fraction of the global emission

allowance that can be allocated to customers, whereas the factor K 0 ue (= 1 − Kue)

accounts for voltage unbalance which arises as a result of line asymmetries. Although

vi
vii

IEC/TR 61000-3-13 recommends system operators to assess the factors Kue and

K 0 ue for prevailing system conditions, a systematic method for its evaluation is not

provided other than a rudimentary direction. This thesis initially examines, employ-

ing radial systems, the influence of line asymmetries on the global emission levels

in medium voltage (MV) and high voltage (HV) power systems in the presence of

various load types/bases including three-phase induction motors. It is shown that

the factor K 0 ue is seen to be dependant not only on line parameters as evident from

IEC/TR 61000-3-13, but also on the downstream load composition. In essence, the

global emission levels in HV power systems is seen to arise as a result of both the local

HV lines and the downstream MV lines in the presence of considerable proportions of

induction motor loads. Eventually, generalised methodologies, covering both radial

and interconnected networks, for the assessment of the global emission in MV and

HV power systems which arises due to line asymmetries are proposed.

In allocating voltage unbalance based on the IEC/TR 61000-3-13 recommenda-

tions, quantitative measures of its propagation from higher voltage to lower voltage

levels in terms of transfer coefficients, and from one busbar to other neighbouring bus-

bar of a sub-system in terms of influence coefficients are required. IEC/TR 61000-3-13

gives a method for evaluating the MV to LV transfer coefficient suggesting a value

less than unity for industrial load bases containing large proportions of mains con-

nected three-phase induction motors, and a value of unity for passive loads in general.

Upon detailed examination, it is noted that a transfer coefficient > 1 can arise in the

presence of commonly prevailing constant power loads. Incorporating these different

influences exhibited by various load types under unbalanced supply conditions on the

propagation, comprehensive methods for assessing the MV to LV and HV to MV

transfer coefficients are proposed. A systematic approach for estimating influence

coefficients for interconnected network environments taking their dependency on the


viii

downstream load composition into account is developed.

The IEC allocation policy with regard to harmonics and flicker has been found not

to guarantee that the emission limits allocated to customers ensure non-exceedance

of the set planning levels. This thesis reports that the above is an issue with voltage

unbalance as well. Overcoming this problem, an alternative allocation technique

referred to as ‘constraint bus voltage’ (CBV) method which closely aligns with the

IEC approach has been suggested for harmonics and flicker. The work presented in

this thesis extends the suggested CBV method to voltage unbalance allocation adding

appropriate revisions to address the additional aspect of the emission which arises as

a result of line asymmetries.

In the application of the IEC/TR 61000-3-13 principles to better manage existing

networks already experiencing excessive voltage unbalance levels, the initial develop-

ment of insights into the influences made by various sources of unbalance is required.

Employing an existing 66kV interconnected sub-transmission system as the study

case, deterministic studies are carried out in a systematic manner considering each of

the asymmetrical elements. Approaches for studying the voltage unbalance behaviour

exhibited by various sources which exist in interconnected network environments are

established. These are employed to identify the most favourable line transposition

options for the study system. Further, this knowledge that facilitates the identifi-

cation of contributions made by individual unbalanced sources forms a platform for

developing techniques to assess the compliance with emission limits, which is another

subject of relevance to future editions of IEC/TR 61000-3-13.

As an essential tool for carrying out the studies, an unbalanced load flow program

based on the phase coordinate reference frame incorporating the component level load

flow constraints and the three-phase modelling of system components is developed.


List of Principal Symbols and Abbreviations

a, b, c refer to the three phases


α summation law exponent
CBV constraint bus voltage
CIGRE International Council on Large Electric Systems
CIRED International Conference on Electricity Distribution
Es:x emission limit of any busbar x of any sub-system S [VUF]
Es:x−j emission limit of any installation j to be connected at any
busbar x of any sub-system S [VUF]
EHV extra high voltage
hm refers to a HV-MV coupling transformer
HV high voltage
I refers to a constant current load
[I] matrix of nodal currents
Iλ:t λ (= 0, +, −) sequence current in any line t [A]
Iλ:x λ (= 0, +, −) sequence component of Ix [A]
Ix nodal current at any busbar x [A]
I−:c/e negative sequence current in any system element e (e = t, tf, busbar x)
caused by any source of unbalance c (c = t, td , lines, Ux ) [A]
IEC International Electrotechnical Commission
IEEE Institute of Electrical and Electronics Engineers
IM refers to a three-phase induction motor load
ka allocation constant
ki−x influence coefficient from any busbar i to any other busbar x

ix
x

klv fraction of LV loads supplied by any higher voltage (MV, HV) busbar
km ratio between the rated motor load (in MVA) and the total
load (in MVA) supplied by an LV system
kmmv ratio between the rated motor load (in MVA) and the total
load (in MVA) supplied by an MV system
kpq ratio between the constant power load (in MVA) and the total
load (in MVA) supplied by an LV system
kpqmv ratio between the constant power load (in MVA) and the total
load (in MVA) supplied by an MV system
ks ratio between the positive and negative sequence impedances of the
aggregated motor load supplied by an LV system
ksc−s ratio between the short-circuit capacity (in MVA) at any busbar S
and the total load (in MVA) supplied by the busbar S
kz ratio between the constant impedance load (in MVA) and the total
load (in MVA) supplied by an LV system
kzmv ratio between the constant impedance load (in MVA) and the total
load (in MVA) supplied by an MV system
Kues:x fraction of the busbar emission allowance at any busbar x of any
sub-system S that can be allocated to installations
K 0 ues:x fraction of the busbar emission allowance at any busbar x of any
sub-system S that accounts for the emission arising as a result of
system inherent asymmetries
LF load flow
LV low voltage
ml refers to a MV-LV coupling transformer
MV medium voltage
NECA National Electricity Code Australia
NEMA National Equipment Manufacturer’s Association
PCC point of common coupling
PQ refers to a constant power load
PS refers to a passive load
xi

rec receiving end busbar of any line t


S represents any sub-system (S = HV, MV, LV)
Ssc:s short-circuit capacity at any busbar S [MVA]
Ss:x total apparent power to be supplied by any busbar x of any
sub-system S [MVA]
Ss:x−ds part of Ss:x supplied at the downstream (DS) [MVA]
Ss:x−j agreed apparent power of any installation j to be connected
at any busbar x of any sub-system S [MVA]
Ss:x−local part of Ss:x supplied locally [MVA]
Ss:x−total total apparent power, as seen at any busbar x of any
sub-system S, to be supplied by the sub-system S [MVA]
send sending end busbar of any line t
t any radial local line of any sub-system under evaluation
td any radial downstream line of any sub-system under evaluation
tij any line between busbars i and j of any sub-system
under evaluation
tf refers to a coupling transformer
Tus−s US to S transfer coefficient
θpf :x power factor angle at any busbar x [deg.]
θpf :z , θpf :pq power factor angle of the constant impedance and constant
power loads respectively supplied by an LV system [deg.]
θpf :zmv , θpf :pqmv power factor angle of the constant impedance and constant
power loads respectively supplied by an MV system [deg.]
θY−+:x phase angle of the admittance Y−+:x [deg.]
θZ−+:td phase angle of the impedance Z−+:td [deg.]
θZλ∆:t phase angle of the impedance Zλ∆:t [deg.]
θIλ:t phase angle of the current Iλ:t [deg.]
Ug/s global emission allowance of any sub-system S [VUF]
Ug/s:x emission allowance of any busbar x of any sub-system S [VUF]
loads
Ug/s:x global emission arising as a result of unbalanced installations
at any busbar x of any sub-system S [VUF]
xii

lines
Ug/s:x global emission arising as a result of system inherent asymmetries
at any busbar x of any sub-system S [VUF]
Uj/s:x emission level caused by any source of unbalance j
at any busbar x of any sub-system S [VUF]
result
Us:x resultant emission level at any busbar x of any sub-system S [VUF]
Ux voltage unbalance at any busbar x [VUF]
UIE International Union for Electricity Applications
US represents any upstream system of any sub-system S
(US = EHV, HV, MV)
[V ] matrix of nodal voltages
Vλ:x λ (= 0, +, −) sequence component of Vx [V]
Vλ:s−us λ (= 0, +, −) sequence voltage, referred to US, at any busbar S [V]
Vn−s nominal line-line voltage of any sub-system S [V]
Vx voltage at any busbar x [V]
lines
V−:g/s:x global negative sequence voltage arising as a result of line
asymmetries at any busbar x of any sub-system S [V]
V−:Ui /x negative sequence voltage at any busbar x caused by
the voltage unbalance Ui that exists at any other busbar i
V Rt voltage regulation of any line t
V Rtd voltage regulation of any line td
VUF voltage unbalance factor [%]
[Y ] matrix of nodal admittances
Yλ∆:xy λ − ∆ (λ, ∆ = 0, +, −) sequence coupling admittance
component of Yxy [S]
Yxy nodal admittance between any busbar x and any
other busbar y [S]
Y−−:x−im downstream negative sequence admittance seen at any
busbar x taking only induction motors into account [S]
Y−+:x downstream negative-positive sequence coupling
admittance seen at any busbar x [S]
xiii

Z refers to a constant impedance load


Zλ∆:t λ − ∆ (λ, ∆ = 0, +, −) sequence coupling impedance
of any line t [Ω]
Zλλ:x downstream λ (λ = 0, +, −) sequence impedance seen
at any busbar x [Ω]
Zλλ:tf −s λ (λ = 0, +, −) sequence impedance, referred to S, of any
coupling transformer [Ω]
Z−−:x−im downstream negative sequence impedance seen at any
busbar x taking only induction motors into account [Ω]
Z−+:td negative-positive sequence coupling impedance
of any line td [Ω]
Z−+:td −us negative-positive sequence coupling impedance, referred
to US, of any line td [Ω]
0, +, − refer to zero, positive and negative sequences respectively
Publications Arising from the Thesis

1. Prabodha Paranavithana, Sarath Perera, and Danny Sutanto. Impact of Un-

transposed 66kV Sub-transmission Lines on Voltage Unbalance. In Proc. Aus-

tralasian Universities Power Engineering Conference (AUPEC 2006), paper 28,

Melbourne, Australia, December 2006.

2. P. Paranavithana, S. Perera, and D. Sutanto. Analysis of System Asymmetry

of Interconnected 66kV Sub-transmission Systems in relation to Voltage Unbal-

ance. In Proc. IEEE Power Engineering Society Conference and Exposition in

Africa (PowerAfrica ’07), Johannesburg, South Africa, July 2007.

3. Prabodha Paranavithana, Sarath Perera, Danny Sutanto, and Robert Koch.

A Systematic Approach Towards Evaluating Voltage Unbalance Problem in In-

terconnected Sub-transmission Networks: Separation of Contribution by Lines,

Loads And Mitigation. In Proc. 13th IEEE International Conference on Har-

monics and Quality of Power (ICHQP 2008), Wollongong, Australia, September-

October 2008.

4. Prabodha Paranavithana, Sarath Perera, and Robert Koch. An Improved

Methodology for Determining MV to LV Voltage Unbalance Transfer Coeffi-

cient. In Proc. 13th IEEE International Conference on Harmonics and Quality

of Power (ICHQP 2008), Wollongong, Australia, September-October 2008.

5. Robert Koch, Alex Baith, Sarath Perera, and Prabodha Paranavithana. Volt-

age Unbalance Emission Limits for Installations - General Guidelines and Sys-

tem Specific Considerations. In Proc. 13th IEEE International Conference

on Harmonics and Quality of Power (ICHQP 2008), Wollongong, Australia,

September-October 2008.

xiv
xv

6. Prabodha Paranavithana, Sarath Perera, and Danny Sutanto. Management of

Voltage Unbalance Through Allocation of Emission Limits to Installations. In

Proc. Australasian Universities Power Engineering Conference (AUPEC 2008),

paper 017, Sydney, Australia, December 2008.

7. Prabodha Paranavithana, Sarath Perera, and Robert Koch. Propagation of

Voltage Unbalance from HV to MV Power Systems. In Proc. 21st International

Conference on Electricity Distribution (CIRED 2009), paper 0497, Prague, June

2009.

8. Prabodha Paranavithana, Sarath Perera, and Robert Koch. A Generalised

Methodology for Evaluating Voltage Unbalance Influence Coefficients. In Proc.

21st International Conference on Electricity Distribution (CIRED 2009), paper

0500, Prague, June 2009.

9. Prabodha Paranavithana and Sarath Perera. Location of Sources of Voltage Un-

balance in an Interconnected Network. In Proc. IEEE Power Engineering So-

ciety General Meeting (panel session on “Developments in Determining Power

Quality Disturbance Sources and Harmonic Source Contributions”) , Calgary,

Alberta, Canada, July 2009.

10. Prabodha Paranavithana and Sarath Perera. A Robust Voltage Unbalance

Allocation Methodology Based on the IEC/TR 61000-3-13 Guidelines. In Proc.

IEEE Power Engineering Society General Meeting , Calgary, Alberta, Canada,

July 2009.

11. P. Paranavithana, S. Perera, R. Koch, and Z. Emin. Global Voltage Unbalance

in MV Power Systems due to Line Asymmetries. Accepted for publication in

IEEE Trans. on Power Delivery.


xvi

12. P. Paranavithana, S. Perera, R. Koch, and Z. Emin. Global Voltage Unbalance

in HV Power Systems due to Line Asymmetries: Dependency on Loads And an

Evaluation Methodology. Accepted for publication in IEEE Trans. on Power

Delivery.

13. Prabodha Paranavithana, Sarath Perera, and Danny Sutanto. Management

of Voltage Unbalance Through Allocation of Emission Limits to Installations.

Accepted for publication in Australian Journal of Electrical and Electronics

Engineering (reproduction of Proc. AUPEC 2008 ).


Table of Contents

1 Introduction 1
1.1 Statement of the Problem . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Research Objectives and Methodologies . . . . . . . . . . . . . . . . . 4
1.3 Outline of the Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2 Literature Review 10
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2 Definition of Voltage Unbalance . . . . . . . . . . . . . . . . . . . . . 11
2.3 Sources of Voltage Unbalance . . . . . . . . . . . . . . . . . . . . . . 13
2.4 Effects of Voltage Unbalance . . . . . . . . . . . . . . . . . . . . . . . 14
2.5 Mitigation Techniques of Voltage Unbalance . . . . . . . . . . . . . . 17
2.6 Measurement and Indices of Voltage Unbalance . . . . . . . . . . . . 18
2.7 Limits of Voltage Unbalance . . . . . . . . . . . . . . . . . . . . . . . 21
2.7.1 Compatibility Levels . . . . . . . . . . . . . . . . . . . . . . . 21
2.7.2 Voltage Characteristics . . . . . . . . . . . . . . . . . . . . . . 22
2.7.3 Planning Levels . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.7.4 Customer Emission Limits . . . . . . . . . . . . . . . . . . . . 26
2.8 Guiding Principles of IEC/TR 61000-3-13 [1] for Voltage Unbalance
Emission Allocation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.8.1 Basic Concepts Used in IEC/TR 61000-3-13 . . . . . . . . . . 28
2.8.2 Emission Limits: Stages 1, 2 and 3 . . . . . . . . . . . . . . . 30
2.8.3 Development of Stage 2 Emission Limits . . . . . . . . . . . . 31
2.8.4 Voltage Unbalance Transfer Coefficients . . . . . . . . . . . . 39
2.8.5 Factor K 0 ue . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
2.9 A Revised Harmonics/Flicker Allocation Technique Based on the IEC
Guidelines - A Preamble to Voltage Unbalance Allocation . . . . . . . 43
2.10 Chapter Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

3 Global Voltage Unbalance in MV Power Systems due to System Inherent


Asymmetries 49
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.2 Influence of Line Asymmetries on the Global Emission and its Depen-
dency on Load Types/Bases . . . . . . . . . . . . . . . . . . . . . . . 52
3.2.1 Constant Impedance (Z) Loads . . . . . . . . . . . . . . . . . 54
3.2.2 Constant Current (I) Loads . . . . . . . . . . . . . . . . . . . 55
3.2.3 Constant Power (P Q) Loads . . . . . . . . . . . . . . . . . . 55
3.2.4 Induction Motor (IM ) Loads . . . . . . . . . . . . . . . . . . 56
3.2.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
3.2.6 Mixes of Passive and Induction Motor Loads . . . . . . . . . . 58
3.3 Methodology for Evaluating the Global Emission Arising Due to Line
Asymmetries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

xvii
xviii

3.4 Verification of the Methodology . . . . . . . . . . . . . . . . . . . . . 66


3.5 Chapter Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

4 Global Voltage Unbalance in HV Power Systems due to System Inherent Asym-


metries 70
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
4.2 Influence of Line Asymmetries on the Global Emission in the Presence
of Induction Motor Loads . . . . . . . . . . . . . . . . . . . . . . . . 74
4.3 Methodology for Evaluating the Global Emission Arising Due to Line
Asymmetries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
4.4 Verification of the Methodology Using a Three-bus Test System . . . 85
4.5 Verification of the Methodology Using the IEEE 14-bus Test System . 89
4.6 Chapter Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

5 Propagation of Voltage Unbalance 94


5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
5.2 Voltage Unbalance Transfer Coefficients . . . . . . . . . . . . . . . . . 97
5.2.1 MV to LV Transfer Coefficient, Tmv−lv . . . . . . . . . . . . . 103
5.2.2 HV to MV Transfer Coefficient, Thv−mv . . . . . . . . . . . . . 110
5.3 Voltage Unbalance Influence Coefficients . . . . . . . . . . . . . . . . 117
5.3.1 Preliminary Investigations - Dependency of Influence Coeffi-
cients on Load Types/Bases . . . . . . . . . . . . . . . . . . . 117
5.3.2 Methodology for Evaluating Influence Coefficients . . . . . . . 121
5.3.3 Verification of the Methodology Using a Three-bus MV Test
System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
5.3.4 Verification of the Methodology Using the IEEE 14-bus Test
System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
5.4 Chapter Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129

6 A Revised Voltage Unbalance Allocation Technique Based on the IEC/TR


61000-3-13 Guidelines 131
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
6.2 Examination of the IEC/TR 61000-3-13 Approach . . . . . . . . . . 132
6.2.1 Calculation of Individual Emission Limits . . . . . . . . . . . 134
6.2.2 Resulting Busbar Emission Levels and Examination Remarks . 138
6.3 A Revised Voltage Unbalance Allocation Technique Based on the CBV
Allocation Principles . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
6.4 Examination of the Revised Voltage Unbalance Allocation Technique 142
6.4.1 Calculation of Individual Emission Limits . . . . . . . . . . . 142
6.4.2 Resulting Busbar Emission Levels and Examination Remarks . 144
6.5 Chapter Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
xix

7 Analysis of the Problem of Voltage Unbalance in Interconnected Power Sys-


tems 147
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
7.2 Voltage Unbalance Behaviour of Line Asymmetries . . . . . . . . . . 150
7.2.1 Impact of the Line Asymmetries of the Study System on the
Voltage Unbalance Problem . . . . . . . . . . . . . . . . . . . 150
7.2.2 Voltage Unbalance Behaviour of the Individual Lines of the
Study System - as Standalone Lines . . . . . . . . . . . . . . . 152
7.2.3 Voltage Unbalance Behaviour of the Individual Lines of the
Study System - as Elements in the Interconnected Network . . 155
7.2.4 General Outcomes - Representation of the Voltage Unbalance
Behaviour of an Asymmetrical Line as an Element in an Inter-
connected Network . . . . . . . . . . . . . . . . . . . . . . . . 160
7.2.5 General Outcomes - Representation of the Interaction of All
Asymmetrical Lines . . . . . . . . . . . . . . . . . . . . . . . . 160
7.3 Voltage Unbalance Behaviour of Load Asymmetries . . . . . . . . . . 167
7.3.1 Impact of the Load Asymmetries of the Study System on the
Voltage Unbalance Problem . . . . . . . . . . . . . . . . . . . 167
7.3.2 Voltage Unbalance Behaviour of the Individual Loads of the
Study System - as Elements in the Interconnected Network . . 169
7.3.3 General Outcomes . . . . . . . . . . . . . . . . . . . . . . . . 174
7.4 Combined Voltage Unbalance Behaviour of Line and Load Asymmetries176
7.4.1 Combined Impact of the Line and Load Asymmetries of the
Study System on the Voltage Unbalance Problem . . . . . . . 176
7.4.2 Representation of the Voltage Unbalance Behaviour of the En-
tire System . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
7.5 Chapter Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181

8 Conclusions and Recommendations for Future Work 184


8.1 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
8.2 Recommendations for Future Work . . . . . . . . . . . . . . . . . . . 191

Appendices

A Derivation of (3.5) 204

B Radial MV-LV Test System


(Fig. 3.2) 207

C Derivation of (3.14) 209

D Y−−:x−im for an MV Network 212

E Application of the Methodology Given by (3.25) to the Three-bus MV Test


System (Fig. 3.7) 214
xx

F Derivation of (4.7) 218

G Derivation of (4.9) 221

H Test Case Description of the Radial HV-MV-LV System (Fig. 4.2) 224

I Y−+:x for an HV Network 227

J Application of the Methodology Given by (3.22) to the Three-bus HV Test


System (Fig. 4.6) 229

K Data of the IEEE 14-bus Test System (Fig. 4.9) 233

L Derivation of (5.18) 237

M Application of the Methodology Given by (5.37) to the Three-bus MV Test


System (Fig. 5.16) 240

N 66kV Sub-transmission Interconnected Study System (Fig. 7.1) - Additional


Data/Information 243
N.1 Operating Conditions at the Considered Time Stamp . . . . . . . . . 243
N.2 Line Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
N.3 An Explanation on the Influence of the Location of an Asymmetri-
cal Line of an Interconnected Network on the Voltage Unbalance Be-
haviour of the Line . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
N.4 A Demonstration of the Linearity of Negative Sequence Voltages . . . 247

O Development of a Method for Unbalanced Load Flow Analysis 249


O.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
O.2 Symmetrical Component Versus Phase Coordinate Reference Frames
for Unbalanced Load Flow Analysis . . . . . . . . . . . . . . . . . . . 250
O.3 Special Considerations in Developing an Unbalanced Load Flow Program250
O.4 Representation of System Components . . . . . . . . . . . . . . . . . 251
O.4.1 Synchronous Generators . . . . . . . . . . . . . . . . . . . . . 251
O.4.2 Passive Loads . . . . . . . . . . . . . . . . . . . . . . . . . . . 254
O.4.3 Overhead Lines . . . . . . . . . . . . . . . . . . . . . . . . . . 255
O.4.4 Capacitor Banks . . . . . . . . . . . . . . . . . . . . . . . . . 256
O.4.5 Three-phase Voltage Regulators/Transformers . . . . . . . . . 256
O.4.6 Three-phase Induction Motors . . . . . . . . . . . . . . . . . . 256
O.4.7 Network Interactions . . . . . . . . . . . . . . . . . . . . . . . 280
O.5 Load Flow Solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280
O.6 Related References . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
List of Figures
2.1 Derating of three-phase induction motors (UIE) . . . . . . . . . . . . 15
2.2 Statistical interpretation of the compatibility level (IEC 61000-2-2,
IEC 61000-2-12) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.3 Statistical interpretation of the planning level (IEC 61000-2-2, IEC 61000-
2-12) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.4 Interpretation of the emission level (IEC/TR 61000-3-13) . . . . . . . 30
2.5 Illustration of the global emission allowance (IEC/TR 61000-3-13) . . 35
2.6 Interconnected sub-system S . . . . . . . . . . . . . . . . . . . . . . . 37
2.7 System representation of any busbar x of the system S shown in Fig. 2.6 37
2.8 Variation of Tmv−lv with km established using (2.17) for various com-
binations of ks and ksc−lv values . . . . . . . . . . . . . . . . . . . . . 40

3.1 Simple MV network . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51


3.2 Radial MV-LV system . . . . . . . . . . . . . . . . . . . . . . . . . . 53
t
3.3 Variation of |V−:g/mv:rec | with |I+:t | (V Rt values corresponding to vari-
ous |I+:t | are also indicated) for the four basic load types . . . . . . . 57
t
3.4 Variation of Ug/mv:rec with km for the cases where klv = 1, klv = 0.5
and klv = 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
3.5 Interconnected MV sub-system . . . . . . . . . . . . . . . . . . . . . 61
3.6 System representation of any busbar x of the MV system shown in
Fig. 3.5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
3.7 Three-bus MV test system considered for applying the proposed method-
ology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
lines
3.8 Emissions Ug/mv:x for the three-bus MV test system for the two cases
where km:2 = 0 and km:2 = 1 . . . . . . . . . . . . . . . . . . . . . . . 68

4.1 Simple HV network . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72


4.2 Radial HV-MV-LV system . . . . . . . . . . . . . . . . . . . . . . . . 75
t+td
4.3 Variation of Ug/hv:rec with klvr for the two cases where kmr = 0 and
kmr = 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
4.4 Interconnected HV sub-system . . . . . . . . . . . . . . . . . . . . . . 80
4.5 System representation of any busbar x of the HV system shown in
Fig. 4.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
4.6 Three-bus HV test system considered for applying the proposed method-
ology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
lines
4.7 Emissions Ug/hv:x for the three-bus HV test system for the cases where
km:2 = 0 and km:2 = 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
lines
4.8 Emissions Ug/hv:x for the three-bus HV test system for the case where
km:2 = 1 in relation to the Phase arrangements I and II of the MV lines 89
4.9 IEEE 14-bus test system . . . . . . . . . . . . . . . . . . . . . . . . . 91
lines
4.10 Emissions Ug/hv:x for the IEEE 14-bus test system . . . . . . . . . . . 91

xxi
xxii

5.1 Variation of Tmv−lv with ksc−lv obtained for constant power loads using
unbalanced load flow analysis . . . . . . . . . . . . . . . . . . . . . . 95
5.2 Radial system considered for the illustration of transfer coefficients . 97
5.3 Variation of Tmv−lv with ksc−lv for constant current loads: I - 0.99
lagging pf, II - 0.9 lagging pf . . . . . . . . . . . . . . . . . . . . . . . 104
5.4 Variation of Tmv−lv with ksc−lv for constant power loads: I - 0.99 lagging
pf, II - 0.9 lagging pf . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
5.5 Variation of Tmv−lv with ksc−lv for induction motor loads with ks = 6.7
and pf = 0.9 lagging . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
5.6 Variation of Tmv−lv with ksc−lv : I - for a load base dominated by in-
duction motors, II - for a load base dominated by passive elements . . 108
5.7 Variation of Tmv−lv with km for ksc−lv ≈ 25 and ksc−lv ≈ 10: I - for
load mixes of Z and IM loads, II - for load mixes of P Q and IM loads 109
5.8 Variation of Tmv−lv with km established using the IEC method, (5.19),
(5.20) and unbalanced load flow analysis . . . . . . . . . . . . . . . . 110
5.9 Variation of Thv−mv with klv for ksc−mv = 12 (loads are supplied directly
at the MV busbar): I - for load mixes of Z and IM loads, II - for load
mixes of PQ and IM loads . . . . . . . . . . . . . . . . . . . . . . . . 115
5.10 Variation of Thv−mv with klv for ksc−mv = 4 (loads are supplied directly
at the MV busbar): I - for load mixes of Z and IM loads, II - for load
mixes of PQ and IM loads . . . . . . . . . . . . . . . . . . . . . . . . 116
5.11 Variation of Thv−mv with klv (LV loads are supplied through MV lines):
I - for ksc−mv = 12, II - for ksc−mv = 4 . . . . . . . . . . . . . . . . . . 116
5.12 Radial MV-LV system (reproduction of Fig. 3.2) . . . . . . . . . . . . 117
5.13 Variation of ksend−rec with km for the cases where klv = 1, klv = 0.5
and klv = 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
5.14 Interconnected sub-system S (reproduction of Fig. 2.6) . . . . . . . . 122
5.15 System representation of any busbar x of the MV system shown in
Fig. 5.14 (reproduction of Fig. 3.6) . . . . . . . . . . . . . . . . . . . 124
5.16 Three-bus MV test system considered for applying the proposed method-
ology (reproduction of Fig. 3.7) . . . . . . . . . . . . . . . . . . . . . 127
5.17 Variations of k1−2 and k1−3 with km:2 for the three-bus MV test system 127
5.18 IEEE 14-bus test system (reproduction of Fig. 4.9) . . . . . . . . . . 128
5.19 Influence coefficients k4−x (x = 1 − 14, x 6= 4) for the IEEE 14-bus
test system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128

6.1 Three-bus HV test system considered for examining the IEC/TR 61000-
3-13 approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
6.2 A comparison of the influence coefficients for the test system derived
using the proposed method: (5.37), and unbalanced load flow analysis 135
6.3 A comparison of the K 0 uex factors for the test system derived using
the proposed method: (4.16), and unbalanced load flow analysis . . . 138
xxiii

6.4 Comparison of the busbar emission limits Ehv:x derived according to


IEC/TR 61000-3-13 and the revised method for the test system: I -
for Case 1, II - for Case 2 . . . . . . . . . . . . . . . . . . . . . . . . 144
reult
6.5 Comparison of the resulting emission levels Ug/hv:x derived according
to IEC/TR 61000-3-13 and the revised method for the test system: I
- for Case 1, II - for Case 2 . . . . . . . . . . . . . . . . . . . . . . . . 145

7.1 66kV sub-transmission interconnected system under study . . . . . . 148


7.2 Measured nodal VUF values for the study system . . . . . . . . . . . 149
7.3 Nodal VUF values (load flow results) which arise as a result of the line
asymmetries, in comparison to the measured values . . . . . . . . . . 151
t
7.4 Variation of |V−:rec | with |I+:t | for the individual lines . . . . . . . . . 153
7.5 Variation of θV−:rec
t with |I+:t | for the individual lines . . . . . . . . . . 154
7.6 Nodal VUF values arising as a result of the individual lines . . . . . . 157
7.7 Phase angles of the nodal negative sequence voltages introduced by the
individual lines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
7.8 Global emission vectors of the individual lines (drawn approximately
to a scale) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
7.9 Resultant influence of the interaction of all asymmetrical lines (drawn
approximately to a scale) . . . . . . . . . . . . . . . . . . . . . . . . . 162
7.10 Nodal contributions made by the individual lines to the resultant volt-
age unbalance levels . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
7.11 (I) Deduced from Fig. 7.8 (II) Effect of the transposition of line F
only (III) Effect of the transposition of lines A and F together (drawn
approximately to a scale) . . . . . . . . . . . . . . . . . . . . . . . . . 165
7.12 Effects, obtained using unbalanced load flow analysis, of the transpo-
sition of line F only, and lines A and F together . . . . . . . . . . . . 166
7.13 Nodal VUF values which arise as a result of the load asymmetries, in
comparison to that of the line asymmetries . . . . . . . . . . . . . . . 168
7.14 Nodal VUF values which arise as a result of the individual loads . . . 170
7.15 Phase angles of the nodal negative sequence voltages introduced by the
individual loads . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
7.16 Global emission vectors of the individual loads (drawn approximately
to a scale) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
7.17 Resultant influence of the interaction of all unbalanced loads (drawn
approximately to a scale) . . . . . . . . . . . . . . . . . . . . . . . . . 175
7.18 Nodal VUF values which arise as a result of both the line and load
asymmetries, in comparison to that of the line asymmetries alone, and
the load asymmetries alone, and also to the measured values . . . . . 177
7.19 Resultant influence of the interaction of all lines and loads (drawn
approximately to a scale) . . . . . . . . . . . . . . . . . . . . . . . . . 178
7.20 Nodal contributions made by the line and load asymmetries to the
overall voltage unbalance levels . . . . . . . . . . . . . . . . . . . . . 179
xxiv

7.21 Nodal contributions made by the individual sources of unbalance to


the overall voltage unbalance levels . . . . . . . . . . . . . . . . . . . 181

O.1 Synchronous generator model . . . . . . . . . . . . . . . . . . . . . . 252


O.2 Load model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
O.3 Equivalent circuit of a voltage regulator/transformer . . . . . . . . . 257
O.4 Three-phase induction motor model proposed in [4, 5] . . . . . . . . . 257
O.5 Variation of the real (P) and reactive (Q) power with the supply voltage
level for a typical three-phase induction motor . . . . . . . . . . . . . 259
O.6 Variation of the real (P) and reactive (Q) power with kp (motor loading
levels corresponding to various kp is also given as a percentage to the
rated output power) for a 2250hp induction motor . . . . . . . . . . . 260
O.7 Variation of the speed with kp (motor loading levels corresponding to
various kp is also given as a percentage to the rated output power) for
a 2250hp induction motor . . . . . . . . . . . . . . . . . . . . . . . . 261
O.8 Impedance type induction motor model . . . . . . . . . . . . . . . . . 261
O.9 PQ type induction motor model . . . . . . . . . . . . . . . . . . . . . 262
O.10 Sequence equivalent circuits of a three-phase induction motor: I - pos-
itive sequence, II - negative sequence . . . . . . . . . . . . . . . . . . 263
|Yim:s | cos(−θim:s ) 0
O.11 Variation of |Y n | cos(−θ n )
of Px−xx with ωωrt
n for the 3hp, 220V motor 270
im:s im:s rt
0
O.12 Variation of |Yim:m2 | sin(−θim:m2 −120 ) 0
n
|Yim:m2 n
| sin(−θim:m2 −1200 )
of Qx−xz with ωωnt for the 3hp, 220V
rt
motor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
O.13 Variation of ηim with ωrt for the 3hp, 220V motor . . . . . . . . . . . 275
O.14 Variation of the per phase input active and reactive power with the
motor loading level for the 3hp, 220V motor excited at the rated voltage
(balanced) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 277
O.15 Variation of the per phase input active and reactive power components
with the motor loading level for the 3hp, 220V motor excited at reduced
and unbalanced voltages . . . . . . . . . . . . . . . . . . . . . . . . . 277
O.16 Variation of the per phase input active and reactive power components
with the motor loading level for a 2250hp, 2.3kV motor excited at
reduced and unbalanced voltages . . . . . . . . . . . . . . . . . . . . 278
O.17 Variation of Pim:a with kp for the existing and proposed induction motor
models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 278
O.18 Variation of Qim:a with kp for the existing and proposed induction
motor models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
List of Tables
2.1 Requirements of background disturbances in assessing the uncertainty
of Class A instruments for the measurement of voltage unbalance (IEC
61000-4-30) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.2 Indicative planning levels given in IEC/TR 61000-3-13 . . . . . . . . 29
2.3 Indicative values for the factor K 0 ue given in IEC/TR 61000-3-13 . . 42

6.1 Influence coefficients for the test system shown in Fig. 6.1 . . . . . . 135
6.2 Shv:x , Shv:x−total and Ug/hv:x for the test system shown in Fig. 6.1 . . . 135
6.3 lines
Ug/hv:x , K 0 uex and Kuex for Case 2 of the test system shown in Fig. 6.1137
6.4 Ehv:x according to IEC/TR 61000-3-13 for the test system shown in
Fig. 6.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
reult
6.5 Ug/hv:x arising as a result of the IEC/TR 61000-3-13 allocation proce-
dure for the test system shown in Fig. 6.1 . . . . . . . . . . . . . . . 139
6.6 Values of the RHS of (6.8) in relation to the test system shown in Fig.
6.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
6.7 ka for the test system shown in Fig. 6.1 . . . . . . . . . . . . . . . . . 143
6.8 Kuex and Ehv:x according to the revised allocation method for the test
system shown in Fig. 6.1 . . . . . . . . . . . . . . . . . . . . . . . . . 143
reult
6.9 Ug/hv:x arising as a result of the revised allocation procedure for the
test system shown in Fig. 6.1 . . . . . . . . . . . . . . . . . . . . . . 144

7.1 Ranking of the sub-transmission lines based on the associated degree


of asymmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
7.2 Parameters, operating features and emission levels of the individual
lines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
7.3 Distribution of the active and reactive power across the three phases
at each of the load busbars of the study system . . . . . . . . . . . . 167
7.4 Operating features and emission levels of the individual loads of the
study system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171

H.1 Values of ksc−lvragg and σ for various klvr . . . . . . . . . . . . . . . . 226

K.1 Voltage controlled bus data . . . . . . . . . . . . . . . . . . . . . . . 233


K.2 Static capacitor data: susceptances . . . . . . . . . . . . . . . . . . . 233
K.3 Generator and load bus data: three-phase MW and MVAr values . . 234
K.4 Transformer data: impedances and secondary tap settings (1st and 2nd
bus numbers refer to the primary and the secondary respectively) . . 234
K.5 Nodal positive sequence voltages . . . . . . . . . . . . . . . . . . . . . 235
K.6 Transmission line data: lengths and impedances . . . . . . . . . . . . 236

L.1 Replacement factors for a mix of various load types . . . . . . . . . . 239

N.1 System details . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243

xxv
xxvi

N.2 Voltage controlled bus data . . . . . . . . . . . . . . . . . . . . . . . 244


N.3 Generator and load bus data: three-phase MW and MVAr values . . 244
N.4 Voltage regulator data: impedances and secondary tap settings . . . . 245
N.5 Static capacitor data: susceptances . . . . . . . . . . . . . . . . . . . 245
N.6 Generator impedance data . . . . . . . . . . . . . . . . . . . . . . . . 245
N.7 Lengths and impedances (Z−+ and Z−+ ) of the sub-transmission lines 246
t
N.8 Negative sequence voltages V−:S2 caused by the individual lines A - N
at the busbar S2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
lines
N.9 Resultant negative sequence voltage V−:S2 at the busbar S2 . . . . . . 248

O.1 Parameters of a 60Hz, 3hp, 220V induction motor . . . . . . . . . . . 270


n
O.2 Power components Px−xx - Qnx−xz for the 3hp, 220V motor . . . . . . 270
O.3 Speed coefficients corresponding to the power components Px−xx -
Qx−xz for a range of induction motors . . . . . . . . . . . . . . . . . 272
O.4 Efficiency coefficients for a range of induction motors . . . . . . . . . 275
Chapter 1

Introduction

1.1 Statement of the Problem

Excessive voltage unbalance1 levels in electrical power systems arising as a result of

unbalanced installations and system inherent asymmetries can cause damage to, and

degradation and maloperation of, customer and utility equipment. Despite the exis-

tence of voltage unbalance regulatory codes, some network service providers are facing

difficulties in complying with stipulated levels. This emphasises the need for recom-

mendations based on well researched engineering practices governing the management

of the problem of voltage unbalance, which this thesis aims to fulfil.

The IEC, one of the world’s leading organisation for standardisation on power

quality, has recently released the Technical Report IEC/TR 61000-3-13 [1] which

provides guiding principles for coordinating voltage unbalance between various voltage

levels of a power system through the allocation of emission limits to installations.

The philosophy of this voltage unbalance allocation process is similar to that of the

counterpart IEC approaches to harmonics (IEC 61000-3-6 [2]) and flicker (IEC 61000-

3-7 [3]) allocation. The absorption capacity or the allowed global emission of a sub-
1
In the context of the thesis, this is limited to negative sequence unbalance.

1
2

system of a power system is established such that the total emission level derived using

the general summation law, taking the upstream contribution in terms of a transfer

coefficient into account, at any point is maintained at or below the set planning

level. The global emission allowance of the sub-system is allocated to its busbars in

proportion to the ratio between the total apparent power to be supplied by the busbar

under evaluation, and the total available apparent power of the sub-system as seen

at the busbar. Voltage unbalance contributions from neighbouring busbars are taken

into account using influence coefficients in determining the total available apparent

power of the sub-system as seen at the busbar. This busbar emission allowance is then

apportioned to individual customers in proportion to the ratio between the agreed

apparent power, and the total apparent power supplied by the busbar.

In the case of voltage unbalance, the global emission at a busbar generally arises

not only as a result of unbalanced installations but also as a result of system inherent

asymmetries (essentially lines). Thus, the apportioning of the total headroom to

installations as in the case of harmonics and flicker can lead to exceedances of the

set planning levels. Hence, IEC/TR-61000-3-13 applies an additional factor which is

referred to as ‘Kue’ to the apportioned allowance. This factor Kue represents the

fraction of the emission allowance that can be allocated to customers, whereas the

factor K 0 ue (= 1 − Kue) accounts for the emission which arises as a result of system

inherent asymmetries. It is recommended that system operators assess the factors

Kue and K 0 ue for prevailing system conditions in their specific networks. However,

a systematic method for its evaluation is not provided other than a rudimentary

direction together with a set of indicative values.

The Technical Report IEC/TR 61000-3-13 gives a method for estimating the MV

to LV transfer coefficient considering the system and load characteristics and the

downstream load composition. This suggests a value less than unity for the trans-
3

fer coefficient in the presence of industrial load bases containing large proportions

of mains connected three-phase induction motors, and a unity transfer coefficient in

relation to passive loads in general. Although a transfer coefficient of unity is math-

ematically trivial for constant impedance loads, its validity has not been cautiously

examined in relation to constant current and constant power loads which may exhibit

different behaviours under unbalanced supply conditions. Further, systematic meth-

ods for assessing the HV to MV and EHV to HV transfer coefficients and influence

coefficients are yet to be developed.

The IEC allocation policy with regard to harmonics and flicker has been found not

to guarantee that the emission limits allocated to individual customer installations

ensure non-exceedance of the set planning levels [4, 5] 2 . Overcoming this problem, an

alternative allocation technique that is referred to as ‘constraint bus voltage’ (CBV)

method which closely aligns with the IEC approach has been suggested for harmon-

ics and flicker [4, 5]. Being based on a common philosophy, the above problem is

anticipated to be experienced also by the recently introduced voltage unbalance allo-

cation approach of IEC/TR 61000-3-13. Thus, it is vital to examine the application

of IEC/TR 61000-3-13 which also involves an additional aspect, i.e. the emission aris-

ing due to system inherent asymmetries. Extension of the CBV method to voltage

unbalance allocation requires revisions addressing this new aspect.

In the application of the IEC/TR 61000-3-13 principles to better manage existing

networks already experiencing excessive voltage unbalance levels, the initial develop-

ment of insights into the influences made by various sources of unbalance is required.

In some circumstances, especially in sub-transmission networks where line transposi-

tion is not a usual practice, the emission which arises as a result of system inherent

asymmetries would not allow an equitable share of busbar emission allowances to in-
2
References [4, 5] are the only sources which provide evidence in support of this statement.
4

stallations. The present knowledge which describes the voltage unbalance behaviour

in interconnected network environments is seen to be limited. Fundamental theories

on a standalone asymmetrical line/load would not provide a comprehensive basis for

understanding the interactive behaviour of various sources of unbalance that exist

in interconnected networks. As an example, in cases where voltage unbalance lev-

els arising as a result of line asymmetries themselves are excessive, making decisions

on line transposition options which effectively correct these networks is seen to be a

challenge.

1.2 Research Objectives and Methodologies

The aim of the work presented in this thesis is to make contributions for further

improvements to the current Technical Report IEC/TR 61000-3-13 through:

• The development of novel/refined methodologies for the assessment of the global

emission arising as a result of system inherent asymmetries, and the propagation

of voltage unbalance.

• The examination of the IEC/TR 61000-3-13 principles in relation to the antic-

ipated problem of exceedance of the set planning levels, and proposing appro-

priate revisions.

• The establishment of theoretical bases to broaden the understanding of the

voltage unbalance behaviour exhibited by various sources that exist in inter-

connected network environments, which would provide additional assistance in

the application of the voltage unbalance allocation methodologies.

To develop a comprehensive understanding of the influence of line asymmetries

on the global voltage unbalance levels in MV and HV power systems and the de-

pendency of this influence on various load types/bases, preliminary investigations are


5

carried out in relation to radial power systems. A basis towards the development

of methodologies for evaluating the global emission in MV and HV power systems

which arises as a result of line asymmetries is established through the extension of the

nodal equations [I] = [Y ][V ] to the sequence domain. This basis is integrated with the

outcomes obtained from the preliminary studies for ascertaining the methodologies.

Development of a systematic approach for the assessment of influence coefficients is

also facilitated by an approach similar to above. Verification of the methodologies is

accomplished using unbalanced load flow analysis3 .

Dependency of the propagation of voltage unbalance from MV to LV and HV to

MV levels on specific load types is initially examined through the development of

theoretical bases which describe the behaviour of these load types under unbalanced

supply conditions. Employing these, the impact of a load base which consists of

various load types on the propagation is established in terms of transfer coefficients.

Examination of the IEC/TR 61000-3-13 principles is achieved through two steps

employing a simple three-bus test system. Consideration to cases both with and with-

out the inclusion of the influence of system inherent asymmetries is given. Firstly, the

emission limits to installations are calculated using the prescribed approach together

with some of the methodologies proposed in this thesis. Secondly, the resulting bus-

bar voltage unbalance levels are established using the general summation law when

all installations inject their allocated limits, and examined against the set planning

level. Extension of the suggested CBV allocation technique to voltage unbalance is

accomplished by introducing its principles while addressing the emission which arises

as a result of system inherent asymmetries according to IEC/TR 61000-3-13.


3
This is a program developed in M AT LAB R . This, which is described in Appendix O, is based
on the phase coordinate reference frame and incorporates the component level load flow constraints
and the three-phase modelling of power system components.
6

To develop theoretical bases which provide an insight into the problem of voltage

unbalance in interconnected network environments, deterministic studies supported

by unbalanced load flow analysis are carried out employing a 66kV sub-transmission

system that is known to experience excessive voltage unbalance levels. Using a new

concept termed ‘voltage unbalance emission vector’ which is derived based on IEC/TR

61000-3-13, the behaviour of each of the lines treating as standalone lines and also as

elements in the interconnected system, and of each of the loads operating in the inter-

connected environment is observed. Through an extensive analysis of these results,

approaches for ascertaining the influence of an unbalanced source, in a global sense,

in terms of a single emission vector (which is referred to as ‘global emission vector’)

are established. Employing the linearity of negative sequence variables, these global

emission vectors of individual unbalanced sources are added forming a basis which

provides a comprehensive understanding of the voltage unbalance behaviour of the

entire system.

1.3 Outline of the Thesis

A brief description of the contents of the remaining chapters is given below:

Chapter 2, a literature review, provides an overview on various general aspects

of voltage unbalance, and a critical discussion on IEC/TR 61000-3-13 on which the

thesis is primarily based. A basic introduction, followed by a review on sources, ef-

fects and mitigation techniques of voltage unbalance is given. Various standards and

documents governing measurement and evaluation procedures, indices and limits of

voltage unbalance are reviewed. The key section of this chapter describes concepts,

principles and related aspects prescribed in IEC/TR 61000-3-13 establishing the back-

grounds for Chapters 3 - 6. The last section discusses fundamental deficiencies of,

and suggested revisions to, the IEC allocation policy with regard to harmonics and
7

flicker forming the background for Chapter 6.

Chapter 3 addresses the aspect of the global voltage unbalance emission which

arises as a result of line asymmetries in relation to MV power systems. Investigations

carried out employing a simple radial network on the influence of line asymmetries on

the global emission levels and its dependency on various load types/bases is presented,

emphasising limitations associated with the respective direction given in IEC/TR

61000-3-13. A systematic approach, covering radial and interconnected networks, for

the evaluation of this global emission at nodal level is proposed. Results established

using this method for a three-bus test system are compared with those obtained using

unbalanced load flow analysis.

As a continuation of the work presented in Chapter 3, Chapter 4 addresses the

same, however in relation to HV power systems which require further investigations

on the influence of the downstream MV line asymmetries on the global emission levels.

Dependency of the global voltage unbalance levels on the local HV line asymmetries

as well as on the untransposed downstream MV lines in the presence of passive and

induction motor loads is presented employing a radial network. Additional limita-

tions, applicable to HV networks, associated with the approach given in IEC/TR

61000-3-13 are emphasised. A methodology for evaluating the global emission which

arises as a result of both the local and downstream line asymmetries is proposed.

Results established using this method for a three-bus test system and also for the

IEEE 14-bus test system are compared with those obtained using unbalanced load

flow analysis.

Chapter 5 focuses on the aspect of the propagation of voltage unbalance. Firstly,

the propagation from upstream higher voltage to downstream lower voltage levels

in terms of transfer coefficients is addressed. Following initial studies on the depen-


8

dency of the MV to LV and HV to MV transfer coefficients on specific load types

which include different passive components and three-phase induction motors, gen-

eralised expressions for their estimation are proposed. Ranges of variation of these

transfer coefficients are demonstrated. The accuracy of this new method for esti-

mating the MV to LV transfer coefficient is compared with the respective method

given in IEC/TR 61000-3-13. Secondly, the propagation from one busbar to other

neighbouring busbars of a sub-system in terms of influence coefficients is addressed.

Preliminary studies carried out employing a radial network on the dependency of

these influence coefficients on various load types is presented. A systematic approach

for the evaluation of influence coefficients for interconnected network environments is

proposed. Results established using this method for a three-bus test system and also

for the IEEE 14-bus test system are compared with those obtained using unbalanced

load flow analysis.

Chapter 6 firstly examines the IEC/TR 61000-3-13 voltage unbalance allocation

principles employing a three-bus test system. The calculation procedure of the emis-

sion limits to installations using the prescribed formulae together with some of the

above proposed methodologies is described. The resulting busbar voltage unbalance

levels when all installations inject their allocated limits are derived, and examined.

Secondly, the principles of the suggested CBV allocation policy are introduced to

voltage unbalance ensuring a robust allocation. These new allocation principles are

examined employing the above three-bus test system.

Chapter 7 establishes theoretical bases for studying the problem of voltage un-

balance in interconnected network environments. Deterministic studies carried out

employing a 66kV interconnected sub-transmission system in relation to its line and

load asymmetries are separately described. Outcomes from these studies are presented

in a generalised form such that a systematic approach which allows a comprehensive


9

understanding of the voltage unbalance behaviour of the system is developed. The

proposed approach is applied for assessing the study system on major contributors

to voltage unbalance levels, and line transposition options which effectively correct

the asymmetrical network. These assessments are validated employing the results

obtained using unbalanced load flow analysis.

Finally, Chapter 8 summarises the major outcomes of the work presented in the

thesis, and makes recommendations and suggestions for future work.


Chapter 2

Literature Review

2.1 Introduction

This chapter provides an overview on various general aspects of voltage unbalance,

and a critical discussion on IEC/TR 61000-3-13 on which the thesis is primarily based.

A brief introduction to voltage unbalance, followed by a review on various methods

used in different standards and documents for its quantification is given in Section 2.2.

Sections 2.3 - 2.5 cover sources, effects and mitigation techniques of voltage unbalance

respectively as reported in the literature. The widely used IEC 61000-4-30 and other

standards/documents governing measurement and evaluation procedures and indices

of voltage unbalance are examined in Section 2.6. Various categories of voltage unbal-

ance limits: compatibility levels, voltage characteristics, planning levels and customer

emission limits are discussed, and a review on limiting values is given in Section 2.7.

The key section of this chapter, Section 2.8, describes concepts, principles and related

aspects prescribed in IEC/TR 61000-3-13 establishing the backgrounds for Chapters

3 - 6. Section 2.9 discusses fundamental deficiencies of, and suggested revisions to, the

IEC allocation policy with regard to harmonics and flicker forming the background

for Chapter 6. The chapter is summarised in Section 2.10.

10
11

2.2 Definition of Voltage Unbalance

The Technical Report IEC/TR 61000-3-13 [1] defines voltage unbalance as a condition

in poly-phase electric power systems in which the magnitudes of the fundamental

phase voltages and/or the associated phase angles of separation are not equal. This

is a steady-state condition, and hence short-term unbalance which can occur during

events such as unsymmetrical faults does not fall under the definition [1]. Voltage

unbalance can exist in two forms in three-phase power systems: zero and negative

sequence unbalance. Where there exists a path for the flow of zero sequence currents

such as in grounded-neutral systems, the presence of zero sequence voltage can become

an issue [6, 7] especially when the coupling transformer allows zero sequence currents

to flow from higher voltage to lower voltage systems and vice-versa. Zero sequence

unbalance is generally a parameter of little concern as it does not affect ungrounded-

neutral systems and dual-phase installations, and also as it can be controlled through

system design and maintenance [1, 8]. As the negative sequence voltage propagates

through all power system components similar to the positive sequence voltage, it is

the quantity of significant concern. Thus, it is common in practice to associate voltage

unbalance with the negative sequence.

The modulus of the ratio of the fundamental negative sequence (V− ) to posi-

tive sequence (V+ ) voltage components, which is known as ‘voltage unbalance factor’

(VUF), as given by (2.1), is used in IEC/TR 61000-3-13 to quantify the degree of

negative sequence voltage unbalance. This seems to be in agreement with most other

standards/codes1 and international working groups2 . This is also referred to as ‘true

value’ of negative sequence voltage unbalance [17, 18].


1
e.g. European EN 50160 [9], South African NRS 048-2 [10] and National Electricity Code Aus-
tralia (NECA) [11].
2
e.g. CIGRE/CIRED [12, 13, 14], International Union for Electricity Applications (UIE) [15, 16].
12


V−
V U F = (2.1)
V+

Reproducing from IEC 61000-4-30 [19], IEC/TR 61000-3-13 gives a practical method

for establishing the VUF using the three fundamental line-line rms voltage magni-

tudes as: s √
1 − 3 − 6
V UF = √ (2.2)
1 + 3 − 6

where,
|Vab |4 +|Vbc |4 +|Vca |4
= (|Vab |2 +|Vbc |2 +|Vca |2 )2

Vab , Vbc and Vca - fundamental line-line rms voltages

Alternative methods for the quantification of voltage unbalance are given by the

National Electricity Manufacturer’s Association (NEMA)3 and the Institute of Elec-

trical and Electronics Engineries (IEEE)4 . The NEMA definition which is known

as ‘line voltage unbalance rate’ (LVUR), and the IEEE definition which is known as

‘phase voltage unbalance rate’ (PVUR) that exists in two different forms (P V U R1 and

P V U R2 ) are given by (2.3), (2.4) and (2.5) respectively. However, the recent IEEE

power quality monitoring standard IEEE 1159 [23] lists both the P V U R1 and the VUF.

Maximum voltage deviation from the average line-line voltage


LV U R = (2.3)
Average line-line voltage

Maximum voltage deviation from the average phase voltage


P V U R1 = (2.4)
Average phase voltage

Difference between the maximum and the minimum phase voltages


P V U R2 =
Average phase voltage
(2.5)
3
NEMA MG1 [20].
4
IEEE 112 [21] and IEEE 100 [22].
13

Although angle unbalance is excluded, the LVUR which does not take the presence

of zero sequence voltage into account is similar to the VUF or the true value for

more realistic levels of voltage unbalance [22, 24]. However, the PVUR which is

influenced by the presence of zero sequence voltage deviates significantly away from

the true value in the presence of zero sequence voltage even at lower levels of voltage

unbalance [24]. Among the two IEEE definitions, the P V U R1 is reasonably close to

the true unbalance in the absence of zero sequence voltage [24].

V−
Although the absolute value of the ratio V+
or the VUF is the parameter in general

use, it is worthwhile noting that voltage unbalance is also associated with a phase

angle. One may, in the same way, define this phase angle as the angle between the

fundamental negative and positive sequence voltage components [25]. This concept

of voltage unbalance as a vector is also applied in IEC/TR 61000-3-13 in defining the

emission level5 introduced by an unbalanced installation at a particular point.

2.3 Sources of Voltage Unbalance

Voltage unbalance is caused mainly by the uneven distribution and/or the uneven

connection of single-phase and dual-phase loads6 across the three phases and the op-

eration of unbalanced three-phase loads7 through the injection of unbalanced phase

currents or negative sequence currents into the system. Unequal mutual impedances

which arise as a result of the asymmetrical electromagnetic coupling between the

conductors of untransposed/partially transposed single circuit [29]/multi circuit [30,

31, 32] transmission and distribution [33, 34] overhead lines, which lead to unbal-

anced voltage drops across the three phases, is also a well known source of voltage

unbalance. Although limited, electrostatic unbalance of untransposed/partially trans-


5
See Section 2.8.1.
6
e.g. LV appliances, electric traction motors [26, 27], induction furnaces.
7
e.g. arc furnaces [28, 16].
14

posed overhead transmission lines [30, 35] and asymmetrical transformer banks [36]

in particular open-wye open-delta transformer banks [37] have also been reported as

additional sources of voltage unbalance.

2.4 Effects of Voltage Unbalance

The influence of voltage unbalance on the adverse performance of three-phase induc-

tion motors is well documented [38, 39, 40]. When an induction motor is exposed to

unbalanced voltages, the negative sequence voltage component produces an air gap

flux that rotates against the rotor which is forced by the positive sequence torque,

thus generating an unwanted reverse torque. This results in a reduction of the net

motor torque and speed, in addition to torque and speed pulsations and increased

motor vibration and noise. Further, due to the relatively small negative sequence mo-

tor impedance, unbalance in phase currents drawn by a motor can be 6 to 10 times

the supply voltage unbalance [20] causing increased motor losses and heating. On the

whole, the motor efficiency and lifetime (primarily as a consequence of the prolonged

overheating) will be reduced. To be able to deal with this extra heating, the motor

must be derated, or a motor of a large power rating may be required. According to

the International Union for Electricity Applications (UIE) [16]8 , an induction motor

has to be derated depending on the prevailing degree of voltage unbalance as depicted

by Fig. 2.1.

Power electronic converters having uncontrolled diode rectifier front-ends9 [42,

43] and arc furnaces [42] produce uncharacteristic triplen harmonics in addition to

the characteristic harmonics in the input current in the presence of supply voltage
8
The derating curve given in [16] is preferred, as it uses the VUF in quantifying voltage unbalance,
in comparison to other recommendations such as given in the standards NEMA MG1 [20] (which
uses the LVUR) and AS 1359.31 [41]/IEC Report 892 (which uses the P V U R1 ).
9
e.g. adjustable speed drives.
15
Please see print copy for image

Figure 2.1: Derating of three-phase induction motors (UIE)

unbalance. Significant third harmonic currents can increase harmonics and resonance

problems in power systems, and require large filter ratings. As the degree of voltage

unbalance increases, the input current drawn by a converter becomes significantly

unbalanced and changes from a double pulse waveform to a single pulse waveform as

a result of the asymmetric conduction of the diodes. This results in excessive currents

in one or two of the phases10 , which can lead to the tripping of overload protection

circuits, under voltage and increased ripple on the dc-link, and decreased lifetime of

the diodes and the dc-link capacitor.

Modern ac drive systems comprising synchronous pulse width modulated (PWM)

rectifier front ends generate a second order harmonic component on the dc-link when

they are exposed to supply voltage unbalance [45]. This results in increased ripple

on the dc-link affecting the life and size of the dc-link capacitor. Further, this second

order harmonic component reflects in the input current and also in the inverter output
10
Measurements taken on an adjustable speed drive system has shown 50% over-current for a
supply voltage unbalance where the highest voltage magnitude was 3.6% higher than the lowest
voltage magnitude [44].
16

voltage by generating a third harmonic component and sub-harmonic components11

respectively.

The impact of some fault conditions (other than the traditionally studied three-

phase fault) on the transient stability of synchronous generators has been seen to

be more severe in the presence of voltage unbalance [46, 47]. This indicates the

requirement of advanced algorithms and computer programs for power system stabil-

ity studies.

Power system components such as synchronous generators, transmission and dis-

tribution overhead lines and cables and transformers can also be affected by voltage

unbalance, which is intensified by the fact that a small degree of unbalance in phase

voltages can cause a disproportionately large unbalance in phase currents as discussed

earlier. Synchronous generators exhibit a phenomenon similar to that in induction

motors in the presence of negative sequence current resulting in excess machine losses

and heating and possible hazards to structural components [48]. According to the

Australian standard AS 1359.10112 [49], synchronous machines shall be capable of

operating continuously in unbalanced systems if none of the phase currents exceeds

the rated current and the ratio of the negative sequence current component and the

rated current does not exceed a value between 5% and 10% depending on the type

of construction, the method of cooling and the machine capacity. Flow of negative

sequence currents in overhead lines, cables and transformers increases power losses

lowering their capacity [50, 51]. From a more theoretical point of view, current un-

balance affects the definitions and the measurement techniques of apparent power

and power factor [52, 53] influencing the aspects of the power system economics. In

addition, current unbalance has been seen to result in a degraded power factor [53].
11
These sub-harmonics will be replaced by a dc component when the inverter output frequency is
equal to twice the system frequency.
12
This is based on IEC 34-1: Rotating electrical machines - part 1 - rating and performance.
17

2.5 Mitigation Techniques of Voltage Unbalance

Methods of voltage unbalance mitigation primarily concern the distribution of power

which includes single-phase, dual-phase and unbalanced three-phase loads evenly

across the three phases. This has been facilitated through the development of tech-

niques for the manual or automatic reconfiguration of distribution systems [54, 55], the

phase rearrangement between distribution transformers and primary feeders [55, 56]

and the switching of connected customers between different phases of a particular

feeder [55] such that unbalance in phase currents is minimised.

Theoretically, the complete transposition of overhead lines or the use of tower

arrangements which provide an equilateral triangular spacing between the three phase

conductors (for single-circuit lines) nullify the emission which arises as a result of

line asymmetries [57]. However, as these ideal conditions can rarely be achieved in

practice due to economic constraints and practical difficulties, the implementation of

more appropriate design options in terms of the tower configuration [30, 32], and the

phase positioning/swapping at transposition points of multi-circuit lines [30, 57, 58]

has been recommended.

Furthermore, the increase of the fault level at the point of common coupling (PCC)

of inherently unbalanced large loads (e.g. traction loads and arc furnaces) can make

some contribution towards reducing their impact on voltage unbalance [16].

In cases where excessive voltage unbalance levels are unavoidable, special balanc-

ing equipment can be installed at the utility and/or plant level. Power electronic

based shunt connected static compensators13 where the compensation is achieved by

the injection or the absorbtion of reactive power to or from the system have been pro-

posed with the development of suitable control algorithms for dynamically correcting
13
e.g. passive static var compensators (SVC) [59], active static synchronous compensators (STAT-
COM) [60] and distribution STATCOMs (DSTATCOM) [61].
18

voltage unbalance. Series connected static compensators14 which provide an active

correction through the injection of a compensating voltage signal in series with the

supply have also been reported as a means for mitigating voltage unbalance. Com-

prehensive techniques such as unified power quality conditioners (UPQC) [66, 67]

and hybrid active and passive filters [68] which are capable of compensating various

power quality disturbances simultaneously have been further advanced also to handle

voltage unbalance.

The principle of the representation of an unbalanced three-phase load (three-wire)

using an equivalent balanced section and a two-phase section has been employed in

reducing the influence of large unbalanced loads (e.g. traction loads) by the use of

special transformer connection topologies such as Scott, V and Le-Blance at their

supply sub-stations [16, 26, 69]. Installation of Steinments compensators consisting

of inductive and capacitive elements at supply sub-stations of large unbalanced loads

is also a well known technique of load unbalance reduction [69].

2.6 Measurement and Indices of Voltage Unbalance

Purpose of the measurement of a power quality disturbance which is stochastic in na-

ture is to obtain statistical information on the performance of the supply or connected

equipment. Site indices are used to provide a statistical description of the disturbance

at a particular site. System indices which are derived using site indices of various

sites15 based on a certain statistical criteria are representatives of the disturbance

over a part of the power system. This section discusses measurement procedures and

indices of voltage unbalance described in various standards and documents16 .


14
e.g. static synchronous series compensators (SSSC) [62, 63, 64], dynamic voltage restor-
ers (DVR) [65].
15
Typically of a particular voltage level, or a group of similar voltage levels.
16
IEC/TR 61000-3-13 [1] is excluded here. This will be reviewed in Section 2.8 covering all related
aspects.
19

The widely accepted standard IEC 61000-4-30 [19] for the measurement of power

quality disturbances prescribes also the voltage unbalance measurement and evalu-

ation procedure for instruments with Class A17 performance. Measurement of the

fundamental component of the three line-line rms voltages over 10-cycle and 12-cycle

intervals for 50Hz and 60Hz systems respectively is specified. A minimum mea-

surement period of one week is recommended. Aggregated values are obtained over

standard time intervals of 3-second, 10-minute and 2-hour18 . The method of quan-

tification is as per (2.2). For instruments with Class B19 performance, the above

specifications are to be provided by manufacturers.

Due to the concurrent existence of various power quality disturbances in typi-

cal power systems, the measurement of a particular disturbance can be affected by

the presence of other background disturbances in the input electrical signal to the

measuring instrument. Thus, IEC 61000-4-30 defines limits for the uncertainty of

instruments with Class A performance when each background disturbance is within

a specified range of variation. For the measurement of voltage unbalance, when other

disturbances exist in the input signal fulfil the requirements given in Table 2.1, except

for voltage unbalance levels in the range of 1% to 5% of the declared input voltage

(Udin ), an instrument shall present an uncertainty less than ±0.15%. For instruments

with Class B performance, the uncertainty is specified by manufacturers.

Derivation of site indices using a high percentile (e.g. 95%, 99%) of the aggre-

gated values (3-second, 10-minute, 2-hour) is preferred in general in most standards

and documents [9, 10, 12, 19]. The 95% percentile of the 10-minute aggregated val-

ues over a measurement period of one week is seen to be strongly recommended

for most power quality disturbances including voltage unbalance. This is the only
17
That is, precise measurements such as for the verification of compliance with standards.
18
A 2-hour value is obtained by combining twelve number of 10-minute values.
19
That is, less precise measurements such as for statistical surveys.
20

Table 2.1: Requirements of background disturbances in assessing the uncertainty of


Class A instruments for the measurement of voltage unbalance (IEC 61000-4-30)
Disturbance Requirement
Power frequency fn ± 0.5Hz (fn - nominal frequency)
Voltage magnitude Udin ± 1%
Flicker Pst < 1 (Pst - short-term flicker severity index)
Harmonics 0% to 3% of Udin
Inter-harmonics 0% to 0.5% of Udin

index used in the European standard EN 50160 [9]. IEC 61000-4-30 proposes a num-

ber of voltage unbalance site indices for contractual applications including the 95%

percentile of the 10-minute and 2-hour aggregated values over a week. The issue

of voltage unbalance indices has also been addressed by the CIGRE/CIRED Joint

Working Group C4.07 [12], and the above index together with the 95% percentile of

the 3-second values over a day has been recently recommended. The South African

standard NRS 048-2 [10] uses the highest of the 10-minute values over a week in

addition to the above index as preliminary site indices. Site indices over long mea-

surement periods are typically calculated as the highest of the daily or weekly indices

(e.g. NRS 048-2).

Among various methods of calculating system indices [25], the choice of a high

percentile of site indices is seen to be popular [12, 70, 71]. The IEC electromagnetic

compatibility standards IEC 61000-2-2 [70] and IEC 61000-2-12 [71] use the 95%

percentile of the 95% site indices as the system index. In addition to the above system

index, [25] recommends the highest of the 95% or 99% site indices as a system index

for voltage unbalance. An alternative approach is given by the CIGRE/CIRED Joint

Working Group C4.07 [12] for a system index in assessing a set voltage unbalance

limit as a low percentage of sites (e.g. 1% and 5%) that exceeds the limit.
21

2.7 Limits of Voltage Unbalance

2.7.1 Compatibility Levels

Connection of equipment to a power system requires that it be able to withstand

any disturbance to which it is subjected by itself and other equipment. Alternatively,

the emission of the disturbance must be limited to a level which is tolerable by the

connected equipment. The primary mechanism defined by the IEC20 to achieve a

balance between the emission and the immunity is the compatibility level. Equip-

ment must be designed to ensure the immunity to the disturbance at least up to

the compatibility level, and utilities are required to maintain the disturbance at or

below the compatibility level. Due to the stochastic nature of the power quality phe-

nomenon, an absolute limit or an expectation of 100% compliance at all times and

locations with a set limit is not sensible. Thus, the compatibility levels are generally

set allowing a small exceeding probability (e.g. 5%) as illustrated in Fig. 2.2 where

the probability density function21 of the disturbance level which represent both time

and space variations and the probability density function of the equipment immunity

level are shown.

The IEC compatibility standards IEC 61000-2-2 [70] and IEC 61000-2-12 [71] give

a value for the voltage unbalance compatibility level in LV and MV power systems

respectively of 2% allowing an excursion up to 3% in some areas where predominantly

single-phase loads are connected. These are based on the 95% non-exceeding proba-

bility level of statistical distributions which represent both time and space variations

of the disturbance. The IEC does not define compatibility levels for HV and EHV

systems. One of the CIGRE papers (Working Group 36.05) [72] proposes a com-

patibility level of 1% for HV networks together with the 2% level for LV and MV
20
IEC 61000-2-2 [70] and IEC 61000-2-12 [71].
21
Inherently normal distributions.
22

Please see print copy for image

Figure 2.2: Statistical interpretation of the compatibility level (IEC 61000-2-2,


IEC 61000-2-12)

systems22 . Use of the compatibility levels is seen to be the usual practice in limiting

voltage unbalance at LV, and the value of 2% is commonly applied (e.g. Belgium,

Italy, United Kingdom, Germany) [16].

2.7.2 Voltage Characteristics

Voltage characteristics are limits within which any user of a public power system

can expect the voltage to remain at the point of utilistion under normal operating

conditions [12]. That is, these limits should not be exceeded for 100% of locations for

a high percentile of time [25].

The European standard EN 50160 [9] defines main voltage characteristics for LV

and MV distribution systems. According to this, the 95% of the 10-minute VUF

values during each measurement period of one week shall be limited at 2%. An

excursion up to 3% is allowed in some areas with partly single-phase or dual-phase


22
However, a clearly defined system index is not given other than stating the 95% of the daily
3-second values to retain to compare against the compatibility levels.
23

installations. The CIGRE/CIRED Joint Working Group C4.07 [12] recommends the

EN 50160 limit for MV networks, and further defines limits of 2% and 1.5% for HV

and EHV systems respectively based on the 95% percentile of the weekly 10-minute

values.

The South African standard NRS 048-2 [10] uses the EN 50160 limit for LV net-

works, and extends this limit to MV and HV levels together with the 3% excursion as

long as the 2% limit is not exceeded for more than 80% of time over the assessment

period. This also gives a limit of 1.5% for EHV networks. The compliance with these

limits is assessed based on the highest of the 95% weekly 10-minute values over the

full measurement period.

The National Electricity Code Australia (NECA) [11] specifies various voltage

unbalance limits as given below:

• Except as a consequence of a contingency event, the average voltage unbalance

measured at a connection point should not vary by more than 0.5%, 1.3% and

2% for Vn > 100kV , 10kV < Vn ≤ 100kV and Vn ≤ 10kV (Vn - nominal supply

voltage) respectively when determined over a 30-minute averaging period.

• As a consequence of a credible contingency event, the average voltage unbalance

measured at a connection point should not vary by more than 0.7%, 1.3% and

2% for Vn > 100kV , 10kV < Vn ≤ 100kV and Vn ≤ 10kV respectively when

determined over a 30-minute averaging period.

• Average voltage unbalance measured at a connection point should not vary

by more than 1%, 2% and 2.5% for Vn > 100kV , 10kV < Vn ≤ 100kV and

Vn ≤ 10kV respectively when determined over a 10-minute averaging period.

• Average voltage unbalance measured at a connection point should not vary

more often than once per hour by more than 2%, 2.5% and 3% for Vn > 100kV ,
24

10kV < Vn ≤ 100kV and Vn ≤ 10kV respectively when determined over a

1-minute averaging period.

Beyond the NECA, different states in Australia have adopted their own electricity

distribution codes. As an example, Victorian electricity distribution code [73] specifies

that a distributor must maintain voltage unbalance at a PCC to a customer’s three-

phase electrical installation at or below 1% with an excursion up to 2% only for a

total of 5 minutes in every 30-minute period.

The American National Standards Institute (ANSI) standard C84.1-1995 devel-

oped by the NEMA states that electrical supply systems should be designed and oper-

ated to limit the maximum voltage unbalance23 to 3% when measured at the electric-

utility revenue meter under no-load conditions [43]24 . Concurrently, the NEMA stan-

dard MG1-1993 [20] recommends that three-phase induction motors should be derated

for voltage unbalance levels25 greater than 1%. Some other ANSI/IEEE standards26

indicate that some electronic equipment, such as computers, may experience problems

for voltage unbalance levels more than 2% or 2.5%.

As given in [12] and [48], voltage unbalance limits used in some other coun-

tries/provinces are listed below:

• France - 2% and 1% limits for MV and HV-EHV networks respectively based

on the 10-minute values over a minimum measurement period of one week [12].

• Quebec (Canada) - 2%, 1.5% and 1% limits for MV, HV and EHV networks

respectively based on the 95% percentile of the 2-hour values over a week [12].
23
Note that this is given in terms of the LVUR.
24
This 3% value stated in ANSI C84.1-1995 is based on the minimum combined cost to utilities
and manufacturers for voltage unbalance related issues.
25
In terms of the LVUR.
26
e.g. ANSI/IEEE standard 141-1993 and ANSI/IEEE standard 241-1990.
25

• Russia - 2% and 4% based on the 95% percentile and the maximum value

respectively [48].

2.7.3 Planning Levels

Planning level is a concept adopted by the IEC27 , and is a limit set for a particular

voltage level by the body responsible for the planning and operation of the supply

system. This can be considered as an internal quality objective of the system opera-

tor. Setting of the planning levels is aimed at the coordination of voltage unbalance

between various voltage levels such that the MV/LV compatibility level is not ex-

ceeded. Planning level is usually equal to or lower than the compatibility level, and

may differ from case to case depending on the network structure and circumstances.

In practice, planning levels are reflected in MV, HV and EHV systems. A statistical

interpretation of the planning level is given in Fig. 2.3.

Please see print copy for image

Figure 2.3: Statistical interpretation of the planning level (IEC 61000-2-2, IEC 61000-
2-12)
27
IEC 61000-2-2 [70] and IEC 61000-2-12 [71].
26

As given in [12] and [16], planning levels used in some countries are listed below:

• United Kingdom - 2% for MV, HV and EHV systems.

• Belgium, Italy - 2% and 1% for MV and HV systems respectively.

Considering practices in various countries and measurement results28 , the CIGRE/CI-

RED Joint Working Group C4.07 [12] recommends planning levels of 2%, 1.5% and

1% for MV, HV and EHV systems respectively based on the 95% daily 3-second

values and/or the 99% weekly 10-minute values. Indicative values for planning levels

provided in IEC/TR 61000-3-13 [1] will be given in Section 2.8.

2.7.4 Customer Emission Limits

Beyond the three basic categories of voltage unbalance limits: compatibility levels,

voltage characteristics and planning levels, some countries impose emission limits on

major customers such as high speed railway systems as a means for managing the

global limits. The limits applied in various countries are stated below.

• Denmark - for connecting a single-phase railway supply station to an HV net-

work, the mean VUF value during a 1-minute period is not allowed to exceed

2% at the supply point [16].

• France - an emission limit of 0.7% for 10-minute periods applies to single-phase

loads connected to MV or HV networks. Permissible emission levels for high

speed trains are 1%, and 1.5% for periods equal to or greater than 15 minutes

and less than 15 minutes respectively [16].

• Germany - it is indicated that the specified 2% LV compatibility level can be

normally met if any connected equipment does not cause emission levels higher

than 0.7% for time ranges of minutes and 1% for time ranges of seconds [16].
28
Reported on in [12].
27

• Italy - if the subscribed power Sj of an individual user does not meet the

requirements: Sj ≤ 30kV A and Sj ≤ 1% of Ssc , Sj ≤ 500kV A and Sj ≤

1% of Ssc , and Sj ≤ 1% of Ssc for LV, MV and HV systems respectively (Ssc -

fault level at the PCC), emission limits of 0.7% for time ranges of minutes and

1% for time ranges of seconds are applied on the customer [16].

• Spain - for high speed train systems, a voltage unbalance limit of 1% is applied

at the PCC. An average value of 2% for 10-minute periods is allowed [16].

• Australia - according to the NECA [11], a customer connected at a 30kV or

higher voltage level must ensure that the current in any phase of a three-phase

electrical installation does not deviate away from the average current by more

than 2%. This limit is 5% for a customer connected at a voltage level lower

than 30kV. The Victorian distribution code [73] defines this limit as 5% and

2% for voltage levels up to 1kV and above 1kV respectively29 .

Furthermore, IEC/TR 61000-3-13 [1] prescribes guiding principles for the assess-

ment of emission limits to three-phase unbalanced installations in an equitable manner

rather than simply specifying limiting values. This will be critically reviewed in the

next section.

2.8 Guiding Principles of IEC/TR 61000-3-13 [1] for Voltage

Unbalance Emission Allocation

The recently released IEC Technical Report IEC/TR 61000-3-13, which is based on

the work that has been undertaken by the CIGRE/CIRED Joint Working Group

C4.103 [13, 14], is the most comprehensive document available in international con-

sensus governing voltage unbalance (negative sequence). The Scope of this report
29
Excursions up to 10% and 4% respectively are permitted for periods less than 2 minutes.
28

covers the provision of guiding principles to system operators, which can be used as a

basis for determining the requirements for the connection of unbalanced installations

to MV, HV and EHV public power systems such that adequate service quality to

all connected customers is ensured. The report addresses the coordination of voltage

unbalance between various voltage levels of a power system through the allocation of

the system capacity to absorb voltage unbalance to individual customers.

This report refers to an unbalanced installation as a complete three-phase instal-

lation, i.e. including both balanced and unbalanced parts, causing voltage unbalance.

Connection of single-phase and dual-phase customer equipment is not specifically ad-

dressed, and the distribution of these loads evenly across the three phases is considered

as a responsibility of the system operator.

2.8.1 Basic Concepts Used in IEC/TR 61000-3-13

Development of emission limits to individual customer installations is based on the

effects of these emissions will have on the quality of the voltage. The concepts of

compatibility level30 , planning level and emission level are used to evaluate the voltage

quality.

Based on the 2% compatibility level at LV, existing practices in MV, HV and

EHV systems and measurement results31 , IEC/TR 61000-3-13 gives indicative plan-

ning levels for MV, HV and EHV systems as reproduced in Table 2.2. These levels

consider the need to provide margins between LV, MV and HV-EHV for the purpose

of the overall electromagnetic compatibility coordination. IEC/TR 61000-3-13 also

provides the general guidance for adopting planning levels for specific systems32 . The
30
Compatibility levels given in IEC/TR 61000-3-13 are reproductions of the IEC compatibility
standards IEC 61000-2-2 [70] and IEC 61000-2-12 [71] which were discussed above in Section 2.7.1.
31
Reported on in [74] by the GIGRE/CIRED Joint Working Group C4.103.
32
Refer to Annex A of IEC/TR 61000-3-13.
29

measurement and evaluation procedure for the assessment of planning levels against

actual voltage unbalance levels is as per IEC 61000-4-30 [19] for instruments with

Class A performance33 . The minimum measurement period is a week with normal

business activities. Use of one or more of the following indices is recommended34 .

• 95% of the weekly 10-minute values should not exceed the planning level.

• The highest of the 99% daily 3-second values should not exceed the planning

level times a multiplying factor35 (e.g. 1.25 - 2) which is to be specified by the

system operator depending on the system characteristics and the short-term

capability of equipment along with their protection devices.

Table 2.2: Indicative planning levels given in IEC/TR 61000-3-13


Voltage level MV HV EHV
Planning level (VUF%) 1.8 1.4 0.8

Emission level introduced by an unbalanced installation into the connected power

system is defined as the magnitude of the unbalanced voltage vector which the con-

sidered installation gives rise to at the point of evaluation36 . This, of an installation j,

is the magnitude of the vector Uj illustrated in Fig. 2.437 . When this vector results

in an increased38 voltage unbalance level (i.e. |Upost-connection | > |Upre-connection |), the

emission level as defined above (i.e. |Uj | in terms of the VUF) is required to be at

or below the emission limit. As the recommended voltage unbalance coordination


33
Refer to Section 2.6.
34
Use of the second index is only needed for installations having a significant impact on the system.
35
Customers are allowed to cause higher emission levels for short periods of time such as during
bursts or start-up conditions.
36
This can be a point of connection or a PCC.
37
Upre-connection , Upost-connection are unbalanced voltage vectors under pre-connection and post-
connection conditions respectively.
38
In some cases, the interaction between the installation and the remaining part of the supply
system may result in a decreased voltage unbalance level (i.e. |Upost-connection | < |Upre-connection |).
30

approach relies on individual emission limits which are being derived39 from plan-

ning levels, the compliance against these emission limits should be assessed based

on the same measurement and evaluation procedure and indices applied for assessing

planning levels against actual voltage unbalance levels.

Please see print copy for image

Figure 2.4: Interpretation of the emission level (IEC/TR 61000-3-13)

2.8.2 Emission Limits: Stages 1, 2 and 3

IEC/TR 61000-3-13 specifies three stages governing the approval for the connection

of unbalanced installations into MV, HV and EHV networks.

• Stage 1 - connection of small installations or installations with a small degree

of unbalance, which fulfil the criteria given by (2.6), can be accepted without a

detailed evaluation of their emission characteristics. That is, no emission limit

will be imposed on these installations.

Suj
≤ 0.2% (2.6)
Ssc

39
The procedure of the derivation of individual emission limits will be discussed in Section 2.8.3.
31

where,

Suj - single-phase power equivalent (line-line or line-neutral equivalent) of the

load unbalance of an installation j

Ssc - three-phase short-circuit capacity at the point of evaluation

• Stage 2 - if an installation does not meet Stage 1 requirement, it should com-

ply with an emission limit imposed based on a certain criteria by the system

operator. This approach of setting individual emission limits will be discussed

in Section 2.8.3.

• Stage 3 - connection of an installation which would fail to comply with Stage 2

emission limit can be conditionally accepted under some circumstances.

2.8.3 Development of Stage 2 Emission Limits

This stage of the IEC/TR 61000-3-13 voltage unbalance management procedure ap-

portions the system absorption capacity to individual customers. The philosophy

of this allocation approach is similar to that of the counterpart IEC harmonics [2]

and flicker [3] allocation methods. However, an additional aspect is involved in the

case of voltage unbalance, i.e. the emission which arises as a result of system inher-

ent asymmetries40 . The general principle of the proposed allocation approach is such

that, when a system is utilised to its designed capacity and all connected installations

inject their individual limits, taking also the emission arising due to system inherent

asymmetries into account, the resultant emission level which arises at any point of

the system should be restricted at or below the set planning level.

Determination of the emission limits to individual customers in an equitable man-

ner is achieved through the following sequence steps:


40
Essentially untransposed or partially transposed overhead transmission/distribution lines.
32

• Adoption of a general summation law for combining emissions arising due to

numerous sources of unbalance.

• Sharing of the system capacity to absorb voltage unbalance between various sub-

systems or voltage levels. This share of a sub-system is referred to as ‘global

emission allowance’ (Ug/s for a sub-system S).

• Apportioning of the global emission allowance of the considered sub-system to

its busbars (Ug/s:x for a busbar x of the sub-system S).

• Allocation of the busbar allowance Ug/s:x to connected customers in an equitable

manner taking into account the emission which arises as a result of system

inherent asymmetries.

General Summation Law

Theoretically, a resultant unbalanced voltage at a point of a system which arises as

a result of the interaction of various sources of unbalance41 is the vector summation

of unbalanced voltage components caused by individual sources at the considered

point. These vectors (magnitude and/or phase) of numerous unbalanced installations

are inherently random and independent. Although voltage unbalance arising as a

result of system inherent asymmetries which primarily depends on line construction

practices is not random, it also varies with the load. Thus, representing these vectors

as stochastic quantities, the following general summation law which avoids the need

for phase angle information is adopted on the basis of experience:

qX
result α
Us:x = (Uj/s:x )α (2.7)
41
i.e. numerous unbalanced installations and system inherent asymmetries.
33

where,
result
Us:x - resultant emission level (a probabilistic quantity) at a busbar x of a system S

Uj/s:x - emission level (a probabilistic quantity) caused by any source of unbalance j

to be combined at the busbar x

α - summation law exponent

The exponent α depends upon the chosen value of probability for the actual volt-

age unbalance level not to exceed the calculated value, and the degree to which the

combined individual unbalanced voltages vary randomly in magnitude and phase.

Considering the 95% non-exceeding probability and the fact that the operation of

most unbalanced installations are unlikely to produce simultaneous or in-phase emis-

sions in practice, IEC/TR 61000-3-13 gives an indicative value for α in the absence of

specific information42 as 1.4. This indicative value would be more applicable in cases

of about eight or more number of dominating unbalanced installations.

Global Emission Allowance

Determination of the global emission allowance of a sub-system is illustrated using the

simple radial system shown in Fig. 2.5. Each of the busbars (US, S and DS) represents

a sub-system43 , where S is the sub-system under evaluation and US and DS are the

upstream and the downstream sub-systems respectively. Adopted planning levels of

the sub-systems US and S are Lus and Ls respectively.

According to IEC/TR 61000-3-13, the global emission in a particular sub-system is

the emission arising due to unbalanced sources that exist in the considered sub-system

and its downstream. Then, two sources which contribute to voltage unbalance in the
42
e.g. when it is known that unbalances are likely to be in-phase and coincident in time, a
summation exponent closer to unity should be used.
43
e.g. US - HV, S - MV and DS - LV.
34

system S can be identified: the global emission (Ug/s ), and the voltage unbalance

(UUus /s ) which propagates from US. By combining these two emissions using the

general summation law, the resultant voltage unbalance level Usresult at S can be

established as:

(Usresult )α = (Ug/s )α + (UUus /s )α (2.8)

Voltage unbalance UUus /s is expressed as a product of a transfer factor and the voltage

unbalance Uus prevailing at US, i.e. UUus /s = Tus−s Uus . This factor Tus−s referred

to as ‘US to S voltage unbalance transfer coefficient’ will be further discussed in

Section 2.8.4. Rearranging (2.8) while restricting the voltage unbalance levels Uus

and Us to the respective planning levels Lus and Ls , the global emission allowance of

the sub-system S can be established as:

p
α
Ug/s = (Ls )α − (Tus−s Lus )α (2.9)

This is the level of voltage unbalance that unbalanced installations supplied by the

sub-system S44 including customers supplied at its downstream, and asymmetrical

lines which exist in the sub-system and its downstream are allowed to cause at any

busbar of the sub-system.

Apportioning of the Global Emission Allowance to Busbars

Avoiding this intermediate step, IEC/TR 61000-3-13 directly gives the formula for

allocating the global emission allowance to individual customers. However, the inclu-

sion of this sequence of steps would assist in providing a comprehensive description

of the allocation process.

44
This can be a single-bus or a multi-bus system.
35
Please see print copy for image

Figure 2.5: Illustration of the global emission allowance (IEC/TR 61000-3-13)

The underlying technique employed for apportioning the global emission allowance

of a sub-system45 to its busbars uses two power definitions, which are described below:

• Total apparent power of installations which are to be supplied by the busbar

under evaluation in foreseeable future. This can be estimated as the sum of

all power flows leaving the busbar while ignoring all power flows between the

considered busbar and other busbars. Considering the sub-system S (Figs. 2.6 -

2.7), this power Ss:x for the busbar x includes customers to be supplied directly

at the busbar (Ss:x−local ) and also at the downstream system (Ss:x−ds ), i.e. Ss:x =

Ss:x−local + Ss:x−ds .

• Total apparent power, as seen at the busbar under evaluation, of installations

which are to be supplied by the entire sub-system. A preliminary approximation

for this power is given as the sum of power flows leaving the busbar while in-

cluding all power flows between the considered busbar and other busbars. This

approximation assumes that the emissions of installations supplied by neighbor-


45
Consider the interconnected sub-system S shown in Fig. 2.6, where Fig. 2.7 shows the system
representation of any busbar x of the sub-system.
36

ing busbars make a direct impact on the considered busbar, which is intended

to be conservative particularly for highly meshed HV and EHV networks. Thus,

the following method given by (2.10) which takes the influences of neighbouring

busbars using a transfer factor into account is recommended for assessing this

power.

Ss:x−total = k1−x Ss:1 + k2−x Ss:2 + ... + Ss:x + ... + ki−x Ss:i + ... + kn−x Ss:n (2.10)

where,

Ss:x−total - total available power of the entire sub-system as seen at the busbar x

Ss:i - total power46 to be supplied by a busbar i, where i = 1, 2, 3..., n, i 6= x

ki−x - voltage unbalance influence coefficient between the busbar i and the

busbar x, that is defined as the voltage unbalance which arises at the busbar x

when 1pu of negative sequence voltage source is applied at the busbar i

The global emission allowance Ug/s is then apportioned to the busbar x in proportion

to the ratio between Ss:x and Ss:x−total as:

s
α
Ss:x
Ug/s:x = Ug/s (2.11)
Ss:x−total

This is the level of voltage unbalance that unbalanced installations supplied by the

busbar x including customers supplied at its downstream, and asymmetrical lines

existing in the sub-system and the downstream system supplied by the busbar x are

allowed to cause at the busbar.


46
The first power definition.
37

Busbar 2

Upstream
system Busbar 1 Busbar x Busbar n
… …

Busbar 3

Sub-system S

Figure 2.6: Interconnected sub-system S

Busbar x

Ss:x-local

Downstream system
supplied by the busbar x

Ss:x-ds

Figure 2.7: System representation of any busbar x of the system S shown in Fig. 2.6
38

Individual Emission Limits

As in the case of the IEC harmonics and flicker allocation, IEC/TR 61000-3-13 also

considers that the allocation of the busbar allowance Ug/s:x to a customer to be con-

nected at the busbar x based on the ratio between the agreed apparent power and

the total power to be supplied by the busbar as an equitable criteria. However, not-

ing that the global voltage unbalance at the busbar is generally caused not only by
loads lines
unbalanced installations (Ug/s:x ) but also by system inherent asymmetries (Ug/s:x ) as

expressed by (2.12), the allocation of the total Ug/s:x to installations may result in

exceedance of the set planning levels.

(Ug/s:x )α = (Ug/s:x
loads α lines α
) + (Ug/s:x ) (2.12)

IEC/TR 61000-3-13 addresses this issue by introducing a new factor Kues:x . This

factor which is defined by (2.13) represents the fraction of Ug/s:x that can be allocated

to installations. Conversely, the factor K 0 ues:x (= 1−Kues:x ) which is given by (2.14)

represents the fraction of Ug/s:x that accounts for the emission arising as a result of

system inherent asymmetries47 .

loads

Ug/s:x
Kues:x = (2.13)
Ug/s:x

lines

Ug/s:x
K 0 ues:x = (2.14)
Ug/s:x

Then, the busbar allowance Ug/s:x is allocated to any customer installation j to be

connected at the busbar x as:


s  
α Ss:x−j
Es:x−j = Ug/s:x Kues:x (2.15)
Ss:x
47
Section 2.8.5 gives a further discussion on this factor.
39

where,

Es:x−j - emission limit of the customer installation j to be connected at the busbar x

Ss:x−j - agreed apparent power of the installation j to be connected at the busbar x

2.8.4 Voltage Unbalance Transfer Coefficients

As indicated in Section 2.8.3 by (2.9), the prescribed voltage unbalance allocation

procedure requires quantitative measures of the propagation of voltage unbalance

from upstream higher voltage (Uus ) to downstream lower voltage (UUus /s ) systems

in terms of transfer coefficients. In general terms, the transfer coefficient Tus−s is

defined as:
UUus /s
Tus−s = (2.16)
Uus

Precise estimation of transfer coefficients facilitates an equitable allocation as the

underestimation of these results in emissions above the set planning levels, whereas

the overestimation causes unnecessary limitations on individual customers.

In addition to the indicative values of 0.9 and 0.95 given for the MV to LV and

HV to MV transfer coefficients respectively, IEC/TR 61000-3-13 prescribes a formula

for estimating the MV to LV transfer coefficient (Tmv−lv ) using the system and load

characteristics and the downstream load composition as:

1
Tmv−lv =   (2.17)
ks −1
1 + km ksc−lv +1

where,

km - ratio between the rated motor load (in MVA) and the total load (in MVA) sup-

plied by the downstream LV system


40

ks - ratio between the positive and negative (which is inductive) sequence impedances

of the aggregated motor load supplied by the LV system48

ksc−lv - ratio between the short-circuit capacity (in MVA) at the LV busbar and the

total load (in MVA) supplied by the LV system

Equation (2.17) indicates a value less than unity for Tmv−lv in the presence of indus-

trial load bases containing large proportions of mains connected three-phase induction

motors, and a unity transfer coefficient in relation to passive loads49 in general. That

is, motor loads help attenuating voltage unbalance as it propagates from higher volt-

age to lower voltage systems, whereas there is no attenuation or amplification in the

presence of passive loads. Fig. 2.8 illustrates the variation of Tmv−lv with km de-

rived using (2.17) for various combinations of ks and ksc−lv values demonstrating that

Tmv−lv can be as small as 0.6. This attenuation has also been seen from measure-

ments taken by the CIGRE/CIRED Joint Working Group C4.103 [74] at a remote

mine with a large proportion of induction motor loads.

1.0

0.9

0.8
Tmv-lv

0.7 ks = 5, ksc-lv = 20
ks = 5, ksc-lv = 10
0.6 ks = 5, ksc-lv = 5
0.5 ks = 7, ksc-lv = 20
ks = 7, ksc-lv = 10
0.4
0.0 0.2 0.4 0.6 0.8 1.0

km

Figure 2.8: Variation of Tmv−lv with km established using (2.17) for various combina-
tions of ks and ksc−lv values

48
Typically, ks can be in the range of 5 to 7.
49
e.g. constant current, constant impedance, constant power loads.
41

2.8.5 Factor K 0 ue

As defined in Section 2.8.3 by (2.14), the factor K 0 ue represents the fraction of a

busbar emission allowance that accounts for the emission arising as a result of sys-

tem inherent asymmetries. IEC/TR 61000-3-13 recommends system operators to

determine this factor based on the line construction practices and the system char-

acteristics in their specific networks. In any case, system operators are responsible

to maintain their networks such that K 0 ue allows an equitable share of the busbar

allowance between unbalanced installations and system inherent asymmetries.

The technical report gives a rudimentary direction towards evaluating this factor

together with a set of indicative values which are reproduced in Table 2.8.5. Consid-

ering a simple radial network having an asymmetrical line50 , the direction mentioned

above indicates that the global emission caused by the line at its receiving end busbar

can be derived using51 :


t |Z−+:t I+:t |
Ug/s:rec ≈ (2.18)
Vnp

where,

t - represents the asymmetrical line


t
Ug/s:rec - global voltage unbalance (in terms of the VUF) caused by the line t at its

receiving end busbar

Z−+:t - negative-positive sequence coupling impedance of the line t

I+:t - positive sequence current in the line t

Vnp - nominal phase voltage of the system


50
An overhead, single-circuit line.
51
The nominal system voltage is taken as the prevailing positive sequence voltage.
42

Table 2.3: Indicative values for the factor K 0 ue given in IEC/TR 61000-3-13
System characteristics K 0 ue
- Highly meshed systems with generation locally connected near load
centers
- Transmission lines are fully transposed, otherwise lines are very
short (few km) 0.1-0.2
- Distribution systems supplying high density load areas with short
lines or cables
- Meshed systems with some radial lines which are either fully or
partly transposed
- Mix of local and remote generation with some long lines 0.2-0.4
- Distribution systems supplying a mix of high density and suburban
areas with relatively short lines (< 10km)
- Long transmission lines which are generally transposed, generation
mostly remote
- Radial sub-transmission lines which are partly transposed or
untransposed
- Distribution systems supplying a mix of medium and low density 0.5-0.4
load areas with relatively long lines (> 20km)
- three-phase motors account for only a small part of the peak load
(e.g. 10%)
43

2.9 A Revised Harmonics/Flicker Allocation Technique Based

on the IEC Guidelines - A Preamble to Voltage Unbal-

ance Allocation

As stated earlier, the IEC approach for managing continuous power quality distur-

bances (IEC/TR 61000-3-13 [1] for voltage unbalance, IEC 61000-3-6 [2] for harmon-

ics, IEC 61000-3-7 [3] for flicker) in power systems through the allocation of emission

limits to individual customer installations is based on a common philosophy. This

section reviews the work that has been undertaken on fundamental deficiencies of, and

suggested revisions to, the counterpart harmonics and flicker allocation approaches.

Being based on a common philosophy, these deficiencies and revisions may appli-

cable to the recently introduced voltage unbalance allocation approach of IEC/TR

61000-3-13, which will be addressed in Chapter 6.

Four key objectives are assumed in the IEC allocation policy:

• Firstly, the level of disturbance at any point of any part of the power system

should not lead to the LV compatibility level is being exceeded. Planning levels

in higher voltage parts should be set accordingly.

• Secondly, the allocation should not distinguish between different types of cus-

tomer installations, i.e. installations of equal MVA demand connected at a

common busbar should receive equal emission limits.

• Thirdly, the allocation should be equitable in some sense, i.e. larger installations

should be entitled to larger emission levels.

• Finally, the emission levels should be as large as possible utilising as much of

the network absorption capacity as possible.


44

Australian/New Zealand standards AS/NZS 61000-3-6 [75] and AS/NZS 61000-3-

7 [76] are essentially based on the respective IEC technical reports on harmonics and

flicker allocation. As a result of difficulties found by Australian utilities in applying

these standards, rigorous studies addressing associated deficiencies have been carried

out by the Integral Energy Power Quality and Reliability Centre at the University of

Wollongong [4, 5]. Consequently, Standards Australia commissioned the writing of

the handbook HB-264-2003 [4] which gives more prescriptive procedures for the use of

these standards. Arising as a result of these studies, it has been revealed that the ap-

plication of the standards to even a simple radial network, let alone relatively complex

meshed systems, is a highly non-trivial exercise. The prescribed allocation procedure

has been seen to lead to situations where the set planning levels are being exceeded

even when no customer exceeds the allocated emission limit. This behaviour has been

identified when using a uniform planning level across the entire network (at a partic-

ular voltage level) as per the IEC approach. Reference [5] demonstrates, employing

a simple example, that it is not possible to derive a practical set of emission limits

such that all busbars in a network reach an uniform planning level when all emission

limits are met. Evidently, a requirement exists for either non-uniform planning levels

for various busbars, or a more suitable allocation policy.

A revised allocation technique for both harmonics and flicker, which closely aligns

with the IEC policy, whereby emission levels at network busbars are explicitly forced

to be at or below the set planning level when all loads inject their limits derived

under the new approach, has been introduced [4, 5]. The principles of this technique

which is referred to as ‘constraint bus voltage (CBV) method’ will be reviewed in the

remaining part of this section.

Consider the requirement that two customer installations (say, m and n) of agreed

apparent power Ss:x−m and Ss:x−n supplied at a busbar x shall receive the same
45

combined emission limit Es:x−mn as a load of Ss:x−mn = Ss:x−m + Ss:x−n connected at

the same busbar. According to the general summation law, the combined emission

limit Es:x−mn can be written in terms of the individual emission limits Es:x−m and

Es:x−m as:

(Es:x−mn )α = (Es:x−m )α + (Es:x−n )α (2.19)

Now, suppose that the emission limit of an installation p is some function f of its

MVA demand Ss:x−p :

Es:x−p = f (Ss:x−p ) (2.20)

Then, (2.19) can hold true if and only if:

(Es:x−m )α + (Es:x−n )α = f α (Ss:x−m ) + f α (Ss:x−n ) = f α (Ss:x−m + Ss:x−n ) (2.21)

That is, f α is associative. The simplest way to satisfy (2.21) is to make the function
p
f proportional to α Ss:x−p :

p
α
f (Ss:x−p ) = Es:x−p = ka Ss:x−p (2.22)

where ka is a proportionality coefficient which is referred to as ‘allocation constant’.

This constant is yet to be determined.

The allocation policy as suggested by the IEC gives the emission limit Es:x−p as52 :

Ug/s p
Es:x−p = √
α
α
Ss:x−p (2.23)
Ss:x−total

By comparing (2.22) and (2.23), it can be observed that the allocation constant under

the IEC allocation method can be given by:


52
For the cases of harmonics (IEC 61000-3-6) and flicker (IEC 61000-3-7).
46

Ug/s
ka = √
α
(2.24)
Ss:x−total

Note that, under the IEC allocation policy, ka is a busbar dependant parameter as

the power Ss:x−total 53 is busbar dependant.

The simplest way to ensure that busbar emission levels do not exceed the set

planning level is the relaxation of the constraint imposed by (2.24). Instead, the

allocation constant ka , making it a global constant, can be chosen simply to be the

largest value such that:

result
Ug/s:x ≤ Ug/s for every busbar x (2.25)

result
The resulting global emission level Ug/s:x at any busbar x can be derived using the

general summation law in terms of busbar emission limits Es:i 54 for i = 1, 2, ..., x, ..., n

and influence coefficients ki−x between busbars i and the busbar x as:

result
p
α
Ug/s:x = (k1−x Es:1 )α + (k2−x Es:2 )α + ... + (Es:x )α + ... + (kn−x Es:n )α (2.26)

Employing (2.22), (2.26) can be written as:

result
p
α α α α
Ug/s:x = ka k1−x Ss:1 + k2−x Ss:2 + ... + Ss:x + ... + kn−x Ss:n (2.27)
53
See Section 2.8.3.
54
i.e. the combined emission limit of a load of which the agreed apparent power is equal to the
total apparent power Ss:i supplied by the busbar.
47

Then, ka can be established in order to satisfy (2.25) as:

Ug/s
ka =  v  (2.28)
u n
u X
α

max Ss:x + t
α ki−x Ss:i 
i=1,i6=x

In summary, the suggested allocation policy is given by (2.22) with ka determined

using (2.28). This new policy meets the four key allocation objectives stated above.

Firstly, (2.25) ensures that the set planning levels are not exceeded when all con-

sumers inject their allocated emission limits. Secondly, based on (2.22), customer

installations of equal MVA demand, whether connected at the same busbar or dif-

ferent busbars, receive identical emission limits. Thirdly, (2.22) ensures that larger

customer installations (in MVA demand terms) to receive larger emission levels than

smaller installations. Finally, the absorption capacity of the network is fully utilised

in the sense that at least one busbar will reach the network planning level.

2.10 Chapter Summary

This chapter has provided general information in relation to voltage unbalance, which

include methods of quantification, sources, effects, mitigation techniques, measure-

ment and evaluation procedures, indices and limits.

The key section of the chapter has given a critical discussion on the IEC/TR

61000-3-13 guidelines for voltage unbalance allocation on which the thesis is primarily

based. Step by step procedure of the development of Stage 2 emission limits together

with the related aspects, i.e. the propagation from higher voltage to lower voltage

levels, the propagation from one busbar to other neighboring busbars of a particular

sub-system, and the emission arising as a result of system inherent asymmetries, has

been clearly described establishing the backgrounds for the remaining chapters.
48

The last section of the chapter has covered the counterpart IEC approaches to

harmonics and flicker allocation in relation to their fundamental deficiencies and

suggested revisions. The principles of this revised allocation technique for harmon-

ics/flicker have been explained forming the background for Chapter 6.


Chapter 3

Global Voltage Unbalance in MV

Power Systems due to System

Inherent Asymmetries

3.1 Introduction

As indicated in Chapter 2 by (2.15), the recommended voltage unbalance allocation

approach requires the determination of the factor K 0 ue (= 1 − Kue) which is system

dependant. For completeness, the definition of K 0 ue is replicated here by (3.1):

lines

Ug/s:x
K 0 uex = (3.1)
Ug/s:x

where,

K 0 uex - factor K 0 ue at any busbar x of any sub-system S (see Figs. 2.6 and 2.7)

Ug/s:x - emission allowance of the busbar x


lines
Ug/s:x - voltage unbalance at the busbar x arising as a result of line asymmetries that

exist in the sub-system S and the downstream system supplied by the busbar x

49
50

The emission which arises as a result of the asymmetries associated with LV feeders

is not of significant concern owing to their shorter lengths and smaller loading levels.

Thus, when aiming at MV power systems (i.e. where sub-system S is an MV system)

on which this chapter is focused, it is only the local MV lines that are responsible for
lines
the emission Ug/mv:x .

Considering a simple radial MV network (refer to Fig. 3.1) with an untransposed


t
overhead single-circuit line (labelled ‘t’), voltage unbalance Ug/mv:rec caused by the

line at its receiving end busbar (labelled ‘rec’) can be expressed as:

V t
t −:g/mv:rec
Ug/mv:rec = (3.2)
V+:rec

where,

t t
|V−:g/mv:rec | = |V−:g/mv:send − (Z−+:t I+:t + Z−−:t I−:t/t + Z−0:t I0:t )| (3.3)

t
V−:g/mv:rec - negative sequence voltage caused by the line at its receiving end busbar

V+:rec - positive sequence voltage at the receiving end busbar

Z−+:t , Z−0:t - negative-positive, negative-zero sequence coupling impedances respec-

tively of the line

Z−−:t - negative sequence impedance of the line, which is inherently equal to its pos-

itive sequence impedance Z++:t 1

I+:t , I0:t - positive and zero sequence currents respectively in the line

I−:t/t - negative sequence current in the line arising as a result of the asymmetry

associated with the line itself


t
V−:g/mv:send - negative sequence voltage caused by the line at its sending end busbar

(labelled ‘send’), which arises as a result of the flow of negative sequence current
1
Impedance Z−−:t will be replaced with Z++:t hereafter.
51

(that is caused by the line) through the transformer coupling the upstream system

and the considered system

Upstream
HV system

send

MV line
(t)

MV system under
consideration
rec

Figure 3.1: Simple MV network

Assuming that zero sequence unbalance in the network is controlled through sys-

tem design, maintenance and operation or an ungrounded-neutral system, the term

Z−0:t I0:t can be set to zero. Then, by comparing (3.3) with the IEC approach stated

in Chapter 2 by (2.18), the IEC approach is seen to assume that the negative se-

quence current I−:t/t is not significant enough to introduce considerable influences by


t t
the terms Z++:t I−:t/t and V−:g/mv:send on the negative sequence voltage |V−:g/mv:rec |.

This seems equitable when the line supplies primarily passive loads (e.g. constant

impedance loads) due to the high negative sequence impedance2 associated with such

loads. However, the validity of the above simplification is questionable when the line

supplies an industrial load base containing a large proportion of mains connected

three-phase induction motors, considering the fact that the current unbalance can
2
Which is equal to the positive sequence impedance.
52

be 6 to 10 times the supply voltage unbalance [20] for an induction motor due to

its relatively low negative sequence impedance. Referring to Table 2.8.5 (last entry),

IEC/TR 61000-3-13 also indicates that K 0 ue is related to the proportion of motor

loads. These emphasise that although by definition K 0 ue is a parameter which ac-

counts for system inherent asymmetries it has some degree of load dependency, an

aspect which requires detailed examination.

Objectives of the work presented in this chapter are:

• To investigate the influence of line asymmetries on the global voltage unbalance

in MV power systems, and its dependency on various load types/bases including

three-phase induction motors. The work carried out in this regard in relation

to a radial line is presented in Section 3.2.

• To develop a generalised methodology, covering radial and interconnected net-

works, for the evaluation of the global voltage unbalance in MV systems that

is caused by line asymmetries at nodal level3 . This facilitates the assessment of

the nodal K 0 ue factors. This is given in Section 3.3.

The proposed methodology is verified in relation to a three-bus test system using un-

balanced load flow analysis in Section 3.4. Section 3.5 summarises the work presented

in the chapter emphasising major conclusions.

3.2 Influence of Line Asymmetries on the Global Emission

and its Dependency on Load Types/Bases

Consider the radial MV-LV system shown in Fig. 3.24 where the MV line is untrans-
t
posed. The purpose is to assess the voltage unbalance Ug/mv:rec caused by the MV line
3 lines
i.e. emissions Ug/mv:x referring to the system shown in Fig. 3.5.
4
Loads, MV-LV coupling transformers and downstream LV busbars supplied by the MV line are
represented as aggregated elements.
53

at its receiving end busbar while the voltage at the sending end busbar is balanced.

For this, the loads Smv:rec−local and Smv:rec−ds are considered as balanced.

t t
The emission Ug/mv:rec can be expressed by (3.2) where |V−:g/mv:rec | given by (3.3)
t
can be simplified for the considered scenario (i.e. V−:g/mv:send = 0) as:

t
|V−:g/mv:rec | = |Z−+:t I+:t + Z++:t I−:t/t | (3.4)

The behaviour of the negative sequence current I−:t/t or the influence of the term
t
Z++:t I−:t/t on |V−:g/mv:rec | may depend on the load type/base supplied by the line.

Thus, the aim of this section is to develop theoretical bases which describe the be-

haviour exhibited by four basic load types5 and also various load bases6 in this regard.

send

MV line
(t)

rec

Smv:rec-local (MVA)

Downstream
LV system

Smv:rec-ds (MVA)

Figure 3.2: Radial MV-LV system

5
i.e. constant impedance, constant current, constant power and three-phase induction motor
loads.
6
i.e. mixes of various load types.
54

3.2.1 Constant Impedance (Z) Loads

When the loads Smv:rec−local and Smv:rec−ds (Fig. 3.2) represent constant impedance

loads (balanced: decoupled and equal positive, negative and zero sequence impedances)7 ,

(3.4) can be re-expressed8 using the decoupled nature of sequence impedances as:

 
t
Z −−:rec
|V−:g/mv:rec | ≈ Z−+:t I+:t (3.5)
Z−−:send

where, Z−−:rec , Z−−:send - downstream9 negative sequence impedances seen at the

receiving end and the sending end busbars respectively of the MV line. Employing

the fact that the negative sequence impedances associated with all system components

are equal to their positive sequence impedances, (3.5) can be re-expressed as:

t
|V−:g/mv:rec | ≈ |Z−+:t I+:t | (1 − V Rt ) (3.6)

where, V Rt - voltage regulation of the line, which is defined as the ratio between

the positive sequence voltage drop across the line (i.e. Z++:t I+:t ) and the sending

end positive sequence voltage. That is, for constant impedance loads, the negative

sequence current I−:t/t behaves in such a manner that the term Z++:t I−:t/t causes
t
the negative sequence voltage |V−:g/mv:rec | to be smaller than the term |Z−+:t I+:t |

considered in the IEC approach: (2.18) by the factor (1 − V Rt ).


7
i.e. the three phase supply is connected to equal impedances where each of the impedances is
modelled as per O.5 - O.6 (Appendix O, Section O.4.2) with λp = λq = 2.
8
Derivation of (3.5) is given in Appendix A.
9
This term, which will be using frequently in the thesis, is employed to indicate an impedance
that is seen at a particular point into the direction where the power flows.
55

3.2.2 Constant Current (I) Loads

The negative sequence current I−:t/t can be considered to be negligible when the line

supplies constant current loads (balanced)10 , as such loads draw equal magnitudes of

three phase currents regardless of the prevailing voltage condition. Hence, (3.4) can

be simplified for constant current loads as:

t
|V−:g/mv:rec | ≈ |Z−+:t I+:t | (3.7)

That is, in contrary to the case of constant impedance loads, the IEC approach: (2.18)

requires no modifications for constant current loads.

3.2.3 Constant Power (P Q) Loads

As the linearisation of system equations and simplifying assumptions (as in the case

of constant current loads) are not supported by constant power loads11 , through

careful examination of the results obtained from unbalanced load flow analysis, (3.8)

is established considering operating scenarios most likely to occur in practice as a


t
close approximation to the negative sequence voltage |V−:g/mv:rec |:

t
|V−:g/mv:rec | ≈ |Z−+:t I+:t | (1 − V Rt )β (3.8)

where, β ≈ −1 and −2 for low (∼ 0.9) and high (∼ 1) lagging power factor (pf)

conditions respectively. That is, for constant power loads, the term Z++:t I−:t/t causes
10
i.e. the three phase supply is connected to equal loads where each of the loads is modelled as
per O.5 - O.6 (Appendix O, Section O.4.2) with λp = λq = 1. In this case, the three-phase load
bank draws equal magnitudes of three phase currents regardless of the voltage condition (including
unbalance) which prevails at the terminals of the load bank.
11
i.e. when the three phase supply is connected to equal loads where each of the loads is modelled
as per O.5 - O.6 (Appendix O, Section O.4.2) with λp = λq = 0. In this case, the power drawn by
each of the three phases of the load bank is equal and does not depend on the voltage condition
(including unbalance) which prevails at the terminals of the load bank.
56
t
the negative sequence voltage |V−:g/mv:rec | to be greater than the term |Z−+:t I+:t |

considered in the IEC approach: (2.18) by the factor (1 − V Rt )β .

3.2.4 Induction Motor (IM ) Loads

Three-phase induction motors12 can be represented using decoupled, unequal and

constant13 sequence impedances. Hence, (3.4) can be re-expressed14 in the form given

by (3.5). This can be expressed15 in terms of the system, line and load characteristics

and the system operating conditions as:

t 1
|V−:g/mv:rec | ≈ |Z−+:t I+:t |    (3.9)
V Rt 1
1+ 1−V Rt 1
+k 1
ks sc−lvagg

where,

ks - ratio between the positive and negative sequence impedances of the aggregated

motor load supplied by the aggregated LV busbar16


Ssc−lvagg
ksc−lvagg = Smv:rec−ds

Smv:rec−ds - total load (in MVA) supplied by the aggregated LV busbar (see Fig. 3.2)
(Vn−lv )2
Ssc−lvagg = |Z++:sys−lv |
, the short-circuit capacity (in MVA) at the aggregated LV

busbar, that is derived using the positive sequence system impedance Z++:sys−lv which

exists between the busbar under evaluation and the downstream LV busbar17

Vn−lv - nominal line-line voltage of the LV system


12
Induction motors are considered to be supplied at the LV level, which is generally the case.
Hence, for this case of motor loads, Smv:rec−local = 0.
13
For a given motor speed.
14
Employing the decoupled nature of sequence impedances.
15
This specific case with klv = 1 and km = 1 can be deduced using the generalised expression (3.14)
of which the derivation is given in Appendix C.
16
Typically, ks can be in the range of 5 to 7.
17
This definition for Ssc−lvagg , which is given in general, will be applied in Chapters 4 and 5 as
well. For the particular case of the network shown in Fig. 3.2, Z++:sys−lv is equal to the impedance
Z++:ml−lv of the MV-LV coupling transformer, and the MV busbar rec is the busbar under evalua-
tion.
57

Equation (3.9) implies that the negative sequence current I−:t/t in the presence of

induction motor loads behaves in such a manner that the term Z++:t I−:t/t causes
t
the negative sequence voltage |V−:g/mv:rec | to be smaller than the term |Z−+:t I+:t |

considered in the IEC approach: (2.18) by the factor 01 1.


“ ”
V Rt 1
1+ 1−V Rt
@
1 + 1
A
ks ksc−lv
agg

Z loads: LF
Z loads: Eq. (3.6) IEC approach:
|Z-+:t I+:t|
80 I loads: LF
I loads: Eq. (3.7)
PQ loads: LF
PQ loads: Eq. (3.8)
IM loads: LF
|V -:g/mv:rec| (V)

IM loads: Eq. (3.9)

40
t

0
|I+:t| (A) 50 150 250 350 450 550 650

VRt (%) 1 3 5 6 8 10 12

t
Figure 3.3: Variation of |V−:g/mv:rec | with |I+:t | (V Rt values corresponding to various
|I+:t | are also indicated) for the four basic load types

3.2.5 Discussion

Taking a generalised view, the additional factors associated with the above proposed

expressions with respect to the IEC approach can be expressed in a form (1 − V Rt )γ

for the three passive load types (where γ = 1 for constant impedance loads, γ = 0

for constant current loads, and −2 ≤ γ ≤ −1 for constant power loads). Consider-

ing most practical circumstances where V Rt < 10%, the factor (1 − V Rt )γ can be

approximated to unity (in other words I−:t/t ≈ 0) supporting the IEC approach for

passive loads in general.


58

The additional factor associated with the proposed expression for induction motor

loads is considerably smaller18 than that for passive loads, implying that the IEC

approach is conservative for induction motor loads as expected.

t
Fig. 3.3 illustrates the variation of the negative sequence voltage |V−:g/mv:rec | with

the current |I+:t | for the four load types19 obtained employing the test system20 de-

scribed in Appendix B. This shows the variations established using the proposed

expressions (3.6) - (3.9) in comparison to the results obtained using unbalanced load

flow (LF) analysis21 , justifying the new formulation and also the above discussion.

3.2.6 Mixes of Passive and Induction Motor Loads

When the MV line supplies a mix of passive loads (at MV and/or LV) and induction

motors (at LV), the negative sequence current I−:t/t can be decomposed as:

I−:t/t = I−:t/ps + I−:t/im (3.10)

where, I−:t/ps , I−:t/im - negative sequence currents (referred to MV side) in passive

(PS) and induction motor (IM) branches respectively arising as a result of the asym-

metry of the MV line. Employing (3.10), (3.4) can be expanded as:

t
|V−:g/mv:rec | = |Z−+:t I+:t + Z++:t I−:t/ps + Z++:t I−:t/im | (3.11)

As discussed in Section 3.2.5, the current component I−:t/ps associated with passive

loads is not significant enough to introduce considerable influence on the negative


18
e.g. 0.6 and 0.9 for induction motor loads and constant impedance loads respectively with
ksc−lvagg = 20, ks = 6.7 and V Rt = 10%.
19
Supplied at the LV level.
20
Where, |Z−+:t | = 0.112Ω, ksc−lvagg ≈ 19 and ks = 6.7.
21
Refer to Appendix O. Out of the two models proposed for three-phase induction motors (Section
O.4.6), the impedance type model is used in the results presented in this thesis.
59
t
sequence voltage |V−:g/mv:rec |. Thus, (3.11) can be simplified as:

t
|V−:g/mv:rec | ≈ |Z−+:t I+:t + Z++:t I−:t/im | (3.12)

Equation (3.12) can be rearranged in a form similar to that of (3.5) as:

 
t
Z −−:rec−im
|V−:g/mv:rec | ≈ Z−+:t I+:t (3.13)
Z−−:send−im

where, Z−−:rec−im , Z−−:send−im - downstream negative sequence impedances seen at

the receiving end and the sending end busbars respectively of the MV line taking

into account only induction motors supplied by the downstream LV busbar. Re-

expressing22 (3.13):

t 1
|V−:g/mv:rec | ≈ |Z−+:t I+:t |    (3.14)
V Rt klv
1+ 1−V Rt 1
+k 1
ks km sc−lvagg

where,

km - ratio between the rated motor load (in MVA) and the total load (in MVA) sup-

plied by the aggregated LV busbar


Smv:rec−ds
klv = Smv:rec−local +Smv:rec−ds
, the fraction of the LV loads supplied by the MV busbar

rec under evaluation

t
The emission Ug/mv:rec can be established in a generalised form by substituting (3.14)

in (3.2) as:
t |Z−+:t I+:t | 1
Ug/mv:rec ≈  × (3.15)

V Rt klv |V+:rec |
1+ 1−V Rt 1
+k 1
ks km sc−lvagg

22
Derivation of (3.14) is given in Appendix C.
60
t
Fig. 3.4 illustrates the variation of Ug/mv:rec with km established using (3.15) in

comparison to the results obtained using unbalanced load flow analysis for the test

system23 in relation to three cases where:

• klv = 1 (i.e. Smv:rec−local = 0)

• klv = 0.5

• klv = 0 (i.e. Smv:rec−ds = 0, which implies that no motor loads are supplied by

the MV line)

The results presented in Fig. 3.4 correspond to a selected operating scenario where

the MV line supplies a total of 10MVA load at 0.9 lagging pf resulting in |I+:t | ≈ 470A,

V Rt ≈ 8.5% and |V+:rec | ≈ 7.2kV (1pu). These results confirm the above basis given

by (3.15) which describes the behaviour of various load bases in relation to the global

emission arising as a result of line asymmetries, also demonstrating the dependency

of the global emission on the motor proportion.

3.3 Methodology for Evaluating the Global Emission Arising

Due to Line Asymmetries

Considering the interconnected network shown in Fig. 3.524 as an MV system with

untransposed lines (t12 , t13 , t23 , ..., tij , ...), the system representation of any busbar x

of the system can be taken as per Fig. 3.625 where the downstream system represents

an aggregated LV system. For the purpose of assessing the global emission arising

as a result of the line asymmetries, the voltage at the upstream system and all loads

supplied by the network are considered to be balanced.


23
Appendix B: |Z−+:t | = 0.112Ω, ksc−lvagg ≈ 19, ks = 6.7.
24
Reproduction of Fig. 2.6.
25
Reproduction of Fig. 2.7.
61

klv = 0 (Smv:rec-ds = 0, i.e. no motor loads are supplied)


0.8

0.7
klv = 0.5
U g/mv:rec (%)

0.6
t

klv = 1 (Smv:rec-local = 0)
0.5
Load flow results
Eq. (3.15)
0.4
0.0 0.2 0.4 0.6 0.8 1.0
km

t
Figure 3.4: Variation of Ug/mv:rec with km for the cases where klv = 1, klv = 0.5 and
klv = 0

Busbar 2

Upstream
HV system Busbar 1 Busbar x Busbar n
… …

Busbar 3

MV sub-system

Figure 3.5: Interconnected MV sub-system


62

Busbar x

Smv:x-local

Downstream
system supplied
by the busbar x

Smv:x-ds

Figure 3.6: System representation of any busbar x of the MV system shown in Fig. 3.5

Employing the linearity of negative sequence variables [1], the resultant negative
lines
sequence voltage V−:g/mv:x arising as a result of the interaction of the untransposed

lines t12 , t13 , t23 , ..., tij , ... at any busbar x can be written as:

lines t12 t13 t23 ij t


V−:g/mv:x = V−:g/mv:x + V−:g/mv:x + V−:g/mv:x + ... + V−:g/mv:x + ... (3.16)

ij t
where, V−:g/mv:x - negative sequence voltage caused by any line tij on its own at the
lines
busbar x. Then, the emission Ug/mv:x can be expressed as:


V lines
lines −:g/mv−x
Ug/mv:x = (3.17)
V+:x

where, V+:x - positive sequence voltage at the busbar x.

Extending the nodal equations [I] = [Y ][V ]26 to the sequence domain, xth element

of [I], (x, y)th element of [Y ] and xth element of [V ] can be written respectively for

the considered MV network as:


26
Where, [I], [Y ] and [V ] are the nodal current, admittance and voltage matrices respectively.
63

 
 I0:x 
 
Ix = 
 I+:x

 (3.18)
 
I−:x
 
 Y00:xy Y0+:xy Y0−:xy 
 
Yxy =
 Y+0:xy Y++:xy Y+−:xy

 (3.19)
 
Y−0:xy Y−+:xy Y−−:xy
 
 V0:x 
 
Vx = 
 V+:x

 (3.20)
 
V−:x

where,

Ix - xth element of [I] or the nodal current vector at the busbar x, which is considered

to be negative when leaving the node

Iλ:x - λ (= 0, +, −) sequence component of Ix

Yxy - (x, y)th element of [Y ], where:

• for x = y, Yxy is equal to the summation of all network admittances connected

to the busbar x (= y)

• for x 6= y, Yxy is equal to the negative value of the admittance of the network

element which connects the busbar x and any other busbar y

Yλ∆:xy - λ−∆ sequence coupling admittance element of Yxy for λ 6= ∆, and λ sequence

impedance element of Yxy for λ = ∆, where λ = 0, +, − and ∆ = 0, +, −

Vx - xth element of [V ] or the voltage at the busbar x

Vλ:x - λ (= 0, +, −) sequence component of Vx


64

The nodal equations [I] = [Y ][V ] of which the elements are given by (3.18) - (3.20)

can be expanded while setting zero sequence voltages and currents to zero to establish

negative sequence nodal currents as:

    
−I−:1 Y Y−+:12 . . . Y−+:1n V
   −+:11   +:1 
    
 −I−:2   Y−+:21 Y−+:22 . . . Y−+:2n   V+:2 
 =   .  +
    
 . .. .. .. ..
 .  . 
 . . . . .  . 
 
 
    
−I−:n Y−+:n1 Y−+:n2 . . . Y−+:nn V+:n
  
Y Y−−:12 . . . Y−−:1n V
 −−:11   −:1 
  
 Y−−:21 Y−−:22 . . . Y−−:2n   V−:2 
+ (3.21)
  
.. .. .. ..  . 
 . 
. . . .  . 


  
Y−−:n1 Y−−:n2 . . . Y−−:nn V−:n

Equation (3.21) can be rewritten in a concise form also employing the inherent rela-

tionship: Y−−:xy = Y++:xy (for x = y and x 6= y) as:

−[I− ] = [Y−+ ][V+ ] + [Y++ ][V− ] (3.22)

Emphasising that these negative sequence nodal currents and voltages arise as a re-

sult of the untransposed lines of the considered network, the matrices [I− ] and [V− ]
lines
are relabelled as [I−:lines ] and [V−:g/mv ] respectively for consistency. As shown in Sec-

tion 3.2 above, the influence of the negative sequence currents [I−:lines ] on the negative
lines
sequence voltages [V−:g/mv ] should be taken into account in the presence of consider-

able proportions of induction motor loads. Thus, the negative sequence nodal current

I−:lines/x at any busbar x can be generally written when the busbar supplies a mix of

passive (at MV and/or LV) and motor (at LV) loads as:
65

lines
I−:lines/x ≈ Y−−:x−im V−:g/mv:x (3.23)

where, Y−−:x−im - downstream negative sequence admittance seen at the busbar x

taking into account only induction motors supplied by the downstream LV busbar.

This admittance Y−−:x−im , which is inherently inductive owing to the inductive nature

of the associated induction motor negative sequence impedances and MV-LV trans-

former impedances, can be expressed27 in terms of the system and load characteristics,

system operating conditions and downstream load composition as:


! √ !
klv:x 3 |I+:x |
Y−−:x−im ≈ −j 1 (3.24)
ksc−lvagg :x
+ ks:x1km:x Vn−mv

where,

ks:x , ksc−lv:x , km:x , klv:x - as defined for (3.9) and (3.14)28

Vn−mv - nominal line-line voltage of the MV system

Substitution of (3.23) in (3.22) and rearrangement gives:

0
lines
[V−:g/mv ]n×1 ≈ −[Y++ ]−1
n×n [Y−+ ]n×n [V+ ]n×1 (3.25)

where,
0
Y++:xy ≈ Y++:xy + Y−−:x−im for x = y
0
Y++:xy = Y++:xy for x 6= y

27
Derivation of (3.24) is given in Appendix D.
28
Additional subscript ‘x’ indicates the quantities corresponding to the busbar x.
66

That is, taking the nodal positive sequence voltages as known quantities as they

can be easily obtained from conventional balanced load flow analysis, the negative
lines
sequence voltage V−:g/mv:x at any busbar x can be established using (3.25).

3.4 Verification of the Methodology

The proposed methodology is applied to the three-bus MV system (60Hz, 12.47kV,

three-wire) shown in Fig. 3.7. Considered operating scenario and resulting positive

sequence system conditions29 are also indicated in Fig. 3.7. Lengths of the lines

which are taken as identical in construction (including the phase positioning) and

untransposed are shown alongside the lines. Relevant admittance data of the lines

are30 :

• positive sequence admittance = (1.0098 − j2.0630)Skm

• negative-positive sequence coupling admittance = (0.0258 + j0.1821)Skm

Busbars 1 and 3 supply MV loads with equal compositions of constant impedance

and constant power elements. That is, km:x = 0 implying that Y−−:x−im ≈ 0 for

x = 1 and 3. Busbar 2 supplies loads at the LV level which account for 40% of the

total load supplied by the system. Two cases based on the type of LV loads are

considered:

• Case 1 - LV loads represent passive elements31 . That is, km:2 = 0 implying that

Y−−:2−im ≈ 0.

• Case 2 - LV loads represent three-phase induction motors. That is, km:2 = 1.

Aggregation of 50hp motors of which the details are given in Appendix B is

considered. This results in an admittance Y−−:2−im ≈ −j0.1316S.


29
Nodal voltages and line currents, which are obtained using load flow analysis.
30
See Appendix B for further details.
31
With equal compositions of constant impedance and constant power loads.
67
lines
Fig. 3.8 illustrates the emissions Ug/mv:x at the MV busbars for the two cases

listed above. This shows the results established using the proposed methodology32 in

comparison to the results obtained using unbalanced load flow analysis validating the

proposed technique. Further, these results reveal that the presence of considerable
lines
proportions of induction motor loads increases the emission Ug/mv:x when x represents

the busbar that is directly connected to the upstream system (e.g. busbar 1 of the

system shown in Fig. 3.7), compared to the case where only passive loads are supplied
lines
by the network. In addition, induction motor loads tend to reduce the emission Ug/mv:x

at all other busbars (e.g. busbars 2 and 3 of the system shown in Fig. 3.7), compared

to the case where only passive loads exist.

204A
MV busbar 2
(0.98pu, -6.740) LV (460V)
4MVA
0.9 lagging pf
10km
Upstream HV 135A
(66kV) system MV busbar 1
(1.05pu, 00) (1.05pu, -4.070)
10km
59A

5km
152A

4MVA
0.9 lagging pf
2MVA
0.9 lagging pf
MV busbar 3
(1.01pu, -5.570)

HV-MV coupling transformer – 12MVA, winding resistance = 1%, leakage reactance = 10%,
secondary tap setting = 1.05pu

MV-LV coupling transformer – aggregated representation of fully loaded 1MVA transformers with
winding resistance = 1%, leakage reactance = 5% and secondary tap setting = 1.05pu

Figure 3.7: Three-bus MV test system considered for applying the proposed method-
ology
32
Application of the methodology to the test system is described in Appendix E.
68

0.8 Load flow: Case 1


Methodology: Case 1
Load flow: Case 2
0.6
Methodology: Case 2
/ mv − x (%)

0.4
U glines

0.2

0.0
1 2 3
Busbar

lines
Figure 3.8: Emissions Ug/mv:x for the three-bus MV test system for the two cases
where km:2 = 0 and km:2 = 1

3.5 Chapter Summary

This chapter has addressed the global voltage unbalance in MV power systems which

arises as a result of line asymmetries. This is a key aspect in assessing emission limits

to individual installations connected to MV power systems essentially based on the

IEC/TR 61000-3-13 recommendations.

The dependency of the global emission on various load types/bases including three-

phase induction motors has been examined in relation to an untransposed radial MV

line. The following major conclusions can be drawn from the study:

• The approach given in IEC/TR 61000-3-13 to assess the influence of an asym-

metrical radial line on the global emission can be applied to an MV network

only when the system supplies primarily passive loads. In this case, the impact

of the negative sequence currents arising as a result of line asymmetries on the

global emission is insignificant.


69

• The IEC approach has been seen to be conservative when the network supplies

a large proportion of induction motor loads. In this case, the negative sequence

currents arising as a result of line asymmetries make a significant influence

on the global emission. The degree of this influence is dependant primarily

on the proportion of motor loads, and secondarily on the system and motor

characteristics.

A systematic approach, covering radial and interconnected networks, for evalu-

ating the global emission caused by line asymmetries at nodal level taking the line,

system and load characteristics, system operating conditions and downstream load

composition into account has been proposed. The results established using the pro-

posed methodology in relation to a three-bus test system has been seen to be in close

agreement with the results obtained using unbalanced load flow analysis. Further,

these results clearly demonstrated the following:

• The presence of considerable proportions of induction motor loads at down-

stream LV systems increases the global emission at the MV busbar which is

directly connected to the upstream system, compared to the case where only

passive loads exist.

• Induction motor loads tend to reduce the global emission levels at all other

busbars of the network, compared to the case where only passive loads exist.
Chapter 4

Global Voltage Unbalance in HV

Power Systems due to System

Inherent Asymmetries

4.1 Introduction

The global voltage unbalance in MV power systems which arises as a result of line

asymmetries and the dependency of this emission on the proportion of induction

motor loads has been investigated, emphasising the limitations associated with the

IEC approach1 in Chapter 3. As a continuation of the work presented in Chapter 3,

this chapter focuses on HV systems with regard to the same subject.

When aiming at an HV network (i.e. where sub-system S is an HV system),


lines
the emission Ug/hv:x at any busbar x can arise as a result of the local HV lines and

also the MV lines that is present in the downstream system supplied by the busbar.

Considering a simple radial HV network (refer to Fig. 4.1) with an asymmetrical

HV line (labelled ‘t’) which supplies an untransposed MV line (labelled ‘td ’) at the
1
Stated in Chapter 2 by (2.18).
70
71
t+td
downstream, the emission Ug/hv:rec caused by the lines2 at the receiving end busbar

(labelled ‘rec’) of the HV line can be expressed as:

t+t
V d
t+td −:g/hv:rec
Ug/hv:rec = (4.1)
V+:rec

where,
t+td t td
|V−:g/hv:rec | = |V−:g/hv:rec + V−:g/hv:rec | (4.2)

t+td
V−:g/hv:rec - negative sequence voltage caused both by the HV and MV lines at the

receiving end busbar of the HV line


t
V−:g/hv:rec - negative sequence voltage caused only by the HV line3 at its receiving end

busbar
td
V−:g/hv:rec - negative sequence voltage caused only by the MV line4 at the receiving

end busbar of the HV line

V+:rec - positive sequence voltage at the receiving end busbar of the HV line

As in the case of the radial MV network considered in Chapter 35 , the negative


t
sequence voltage V−:g/hv:rec that is caused by the local HV line can be expressed,

ignoring zero sequence unbalance, as:

t t
V−:g/hv:rec = V−:g/hv:send − (Z−+:t I+:t + Z−−:t I−:t/t ) (4.3)

td
The symbols in (4.3) are as defined in Chapter 3 for (3.3)6 . The second term V−:g/hv:rec

in (4.2), which is caused by the downstream MV line can be expressed as7 :


2
i.e. both HV and MV lines.
3
i.e. taking the MV line as balanced.
4
i.e. taking the HV line as balanced.
5
See Section 3.1.
6
Note that t represents the local HV line here.
7
The HV line is treated as balanced in this case, i.e. Z−+:t = 0.
72

td td
V−:g/hv:rec = V−:g/hv:send − Z−−:t I−:td /t (4.4)

where,

I−:td /t - negative sequence current in the HV line, which arises as a result of the

asymmetry associated with the MV line


td
V−:g/hv:send - negative sequence voltage caused by the MV line at the sending end

busbar (labelled ‘send’) of the HV line, which arises as a result of the flow of the

negative sequence current (that is caused by the MV line) through the transformer

coupling the upstream system and the considered HV system

Upstream
system

send

HV line HV system under


(t) consideration

rec

MV line
(td)

Figure 4.1: Simple HV network


73

By comparing (4.2) of which the two terms are given by (4.3) and (4.4) respectively,

with the IEC approach, it can be identified that the IEC approach infers the following

when applied to HV networks:

• Negligible influence from the negative sequence currents that are caused by the

local HV lines, which is similar to the associated simplification in assessing MV

systems as discussed in Chapter 3. As evident from Chapter 3, this simplifica-

tion can lead to a significant degree of error when the load supplied by an HV

network accounts for a large proportion of induction motors.

• Negligible influence from the negative sequence currents that are caused by the

downstream MV lines. In other words, the IEC approach leaves out the presence

of the downstream line asymmetries in evaluating the global emission in HV

networks. However, the asymmetry associated with MV lines, which was seen

in Chapter 3 to introduce negative sequence currents that are significant enough

to influence the local emission in the presence of large motor proportions, can

also make a considerable impact on the global emission in HV networks. This

demands additional investigations on the subject applicable to HV systems.

Objectives of the work presented in this chapter are:

• To investigate the influence of line asymmetries, which include the local HV lines

and downstream MV lines, on the global emission in HV systems in the presence

of induction motor loads. The work carried out in this regard in relation to a

simple radial network is presented in Section 4.2.

• To develop a generalised methodology, covering interconnected network envi-

ronments, for the evaluation of the global voltage unbalance in HV systems

that is caused by the local and downstream line asymmetries at nodal level.

This is covered in Section 4.3.


74

The proposed methodology is applied to a three-bus test system and the results are

compared with those obtained using unbalanced load flow analysis in Section 4.4.

Section 4.5 further verifies the proposed methodology employing the IEEE 14-bus

test system which supplies passive loads locally8 . Section 4.6 summarises the work

presented in the chapter emphasising major conclusions.

4.2 Influence of Line Asymmetries on the Global Emission

in the Presence of Induction Motor Loads

Consider the radial HV-MV-LV network shown in Fig. 4.29 where the HV and MV
t+td
lines of interest are untransposed. The purpose is to assess the emission Ug/hv:rec that

is caused by the line asymmetries at the receiving end busbar of the HV line while

the voltage at the sending end busbar is balanced. For this, all loads supplied by the

system are considered to be balanced.

t+td
The emission Ug/hv:rec can be expressed by (4.1) where the two components of
t+td
the resultant negative sequence voltage |V−:g/hv:rec | which are given by (4.3) and (4.4)
t
respectively can be simplified for the considered scenario (i.e. V−:g/hv:send = 0 and
td
V−:g/hv:send = 0) as10 :

t
V−:g/hv:rec = −(Z−+:t I+:t + Z++:t I−:t/t ) (4.5)

td
V−:g/hv:rec = −Z++:t I−:td /t (4.6)
8
i.e. at the HV level itself.
9
Fig. 4.2 gives a generalised representation of the downstream system supplied by the HV line.
LV busbars, where motor loads are connected, which can be supplied through MV lines (labelled
‘LVr ’ where subscript ‘r’ indicates the receiving ends of the MV lines) and also directly by HV-MV
coupling transformers (i.e. not through MV lines, labelled as ‘LVs ’ where subscript ‘s’ indicates
the sending ends of the MV lines) are separately considered. Loads, HV-MV and MV-LV coupling
transformers, and downstream MV lines and MV and LV busbars supplied by the HV line are
represented as aggregated elements.
10
The impedance Z−−:t is replaced with Z++:t .
75

send

HV line
(t)

rec

MV line
(td)

LVs

Downstream
LVr system supplied
by the HV line

Figure 4.2: Radial HV-MV-LV system

t+td
The resultant negative sequence voltage |V−:g/hv:rec |, which is the absolute value of

the sum of (4.5) and (4.6), can be generally expressed11 when the network supplies a

mix of passive loads (at HV, MV and/or LV) and induction motors (at LV) as:

t+td
|V−:g/hv:rec | ≈ |Z−+:t I+:t (1 − µ − ζ)| (4.7)

where,

Z++:t
µ = Z−−:send−im
11
Derivation of (4.7) is given in Appendix F.
76

Z−−:send−im - downstream negative sequence impedance seen at the sending end bus-

bar of the HV line taking into account only induction motors supplied by the LV

busbars (i.e. LVs and LVr )

ζ = µktd kn σ
Z−+:td −hv
σ= Z−+:t
(a complex quantity)

Z−+:td −hv - negative-positive sequence coupling impedance, referred to HV side, of

the MV line

ktd - ratio between the total load supplied by the MV line, and the total load supplied

by the HV busbar rec

kn - ratio between the downstream negative sequence impedance12 seen at the sending

end busbar of the MV line taking into account the total motor load13 supplied by the

HV line, and the downstream negative sequence impedance seen at the sending end

busbar of the MV line taking into account only the motor load14 supplied the MV line

t+td
The emission Ug/hv:rec can be established in a generalised form by substituting (4.7) in

(4.1) as:

t+td
Z−+:t I+:t (1 − µ − ζ)
Ug/hv:rec ≈
(4.8)
V+:rec

The factors µ and ζ (= µktd kn σ) represent the impact of the term Z++:t I−:t/t in

(4.5) and of the term given by (4.6) or of the downstream MV line respectively on the
t+td
emission Ug/hv:rec in the presence of induction motor loads. On the whole, the term

µ + ζ = µ (1 + ktd kn σ) represents the overall influence introduced by motor loads,

of which the magnitude can be greater or smaller than µ and ζ noting the phase

angle involved with σ. For the purpose of demonstrating the level of this influence,
12
Ignoring the presence of any passive loads.
13
i.e. the motor load supplied both by the busbars LVs and LVr .
14
i.e. the motor load supplied only by the busbar LVr .
77

the factor µ is expressed15 in terms of the system and load characteristics, system

operating conditions and downstream load composition for a simplified case where

motor loads are supplied only by the MV line at the busbar LVr 16 as:

1
µr ≈ „ « (4.9)
1 1
(1−V Rt )(1−V Rtd )2 ksr kmr
+k
sc−lvragg
1+ V Rt klvr

where,

µr - factor µ corresponding to the simplified case where motor loads are supplied only

by the busbar LVr

V Rt , V Rtd - voltage regulations of the HV and MV lines respectively17

ksr - ratio between the positive and negative sequence impedances of the aggregated

motor load supplied by the busbar LVr 18

kmr - ratio between the rated motor load (in MVA) and the total load (in MVA)

supplied by the busbar LVr

klvr - ratio between the total load (in MVA) supplied by the busbar LVr , and the

total load (in MVA) supplied by the HV busbar rec under evaluation

ksc−lvragg - ratio between the short-circuit capacity19 (in MVA) at the aggregated

busbar LVr , and the the total load (in MVA) supplied by the busbar LVr

15
Derivation of (4.9) is given in Appendix G.
16
i.e. the busbar LVs , which is directly supplied by the HV-MV coupling transformer, supplies
primarily passive loads resulting in a kn = 1.
17
Voltage regulation is defined as the ratio between the positive sequence voltage drop across the
line (e.g. Z++:t I+:t for the HV line), and the sending end positive sequence voltage.
18
Typically, ksr can be in the range of 5 to 7.
19
Which is derived using the positive sequence system impedance Z++:sys−lv = Z++:hm−lv +
Z++:td −lv +Z++:mlr −lv that exists between the HV busbar rec under evaluation and the downstream
busbar LVr , where Z++:hm−lv , Z++:td −lv , Z++:mlr −lv are the positive sequence impedances, referred
to LV, of the HV-MV coupling transformer, MV line and MV-LV coupling transformer supplying
the busbar LVr respectively.
78
t+td
Fig. 4.3 illustrates the variation of Ug/hv:rec with klvr established using (4.8) in

comparison to the results obtained using unbalanced load flow analysis for the sim-

plified test case20 described in Appendix H in relation to two cases where:

• kmr = 0 (i.e. busbar LVr supplies only passive loads)

• kmr = 1 (i.e. busbar LVr supplies only induction motors)

The operating scenario considered corresponds to |I+:t | ≈ 490A, V Rt ≈ 7%, V Rtd ≈

9% and |V+:rec | ≈ 39.6kV (1.04pu). Values of ksc−lvragg and σ corresponding to various

klvr are given in Appendix H. Further, the levels of the influence of the factors µ21

(i.e. of the term Z++:t I−:t/t ) and ζ 22 (i.e. of the downstream MV line) on the emission

in each case are also indicated in Fig. 4.323 . Arising from these results, the following

can be concluded:

t+td
• For passive loads, the influence of the factors µ and ζ on the emission Ug/hv:rec

is insignificant implying that the IEC approach can be accepted only when the

line supplies primarily passive loads also in assessing HV systems. That is, in

the case of passive loads, the local HV lines are totally responsible for the global

emission and the contribution made by the downstream MV lines is negligible.

t+td
• The presence of induction motors affects the level of emission Ug/hv:rec given by

the IEC approach noticeably24 . This dependency of the global emission on the
20
Where, |Z−+:t | = 0.5226Ω, ksr = 6.7, kn = 1, ktd = klvr .
21
For the case of kmr = 0, µ is the difference between the results established using (4.8) for
kmr = 0, and those obtained using unbalanced load flow analysis for kmr = 0 without introducing
the effects of the downstream MV line asymmetry. For the case of kmr = 1, µ is the difference
between the results established using (4.8) for kmr = 0, and those obtained using unbalanced load
flow analysis for kmr = 1 without introducing the effects of the downstream MV line asymmetry.
22
For the both cases of kmr = 0 and kmr = 1, ζ is the difference between the results established
using unbalanced load flow analysis with and without introducing the effects of the downstream MV
line asymmetry.
23
The overall influence introduced by motor loads on the emission for the test case is given by the
summation µ + |ζ| noting that the phase angle (see Table H.1) involved with σ is zero (i.e. phase
angles of the impedances Z−+:t and Z−+:td of the HV and MV lines respectively are equal).
24
e.g. motor loads weighted only 20% cause 10% reduction in the emission compared to that when
only passive loads exist.
79

motor proportion can be seen in two forms which are explained by the factors µ

and ζ respectively. Prominently, in the case of motor loads, not only the local

HV line but also the downstream MV line are equally responsible in determining

the global emission.

Load flow results


μ | Z − +:t I +:t |
d
Eq. (4.8) for kmr = 0
0.7 kmr = 0 caused only by t
ζ | Z − +:t I +:t |
d
for kmr = 0

caused by both t and t d μ | Z − +:t I +:t |


U gt +/thvd :rec (%)

0.6 d
for kmr = 1

kmr = 1
caused only by t
ζ | Z − +:t I +:t |
0.5
kmr = 1
d
caused by both t and t d for kmr = 1

0.4
0.0 0.2 0.4 0.6 0.8 1.0

k lvr

t+td
Figure 4.3: Variation of Ug/hv:rec with klvr for the two cases where kmr = 0 and
kmr = 1

4.3 Methodology for Evaluating the Global Emission Arising

Due to Line Asymmetries

Considering the network shown in Fig. 4.425 as an HV system with asymmetrical lines

(t12 , t13 , t23 , ...), the system representation of any busbar x of the system can be taken

as per Fig. 4.5. This system supplied by the busbar x consists of a radial26 MV-LV

network with an untransposed MV line td:x . For assessing the global emission arising
25
Reproduction of Fig. 2.6.
26
Which is the usual practice.
80

as a result of the line asymmetries, the voltage at the upstream system and all loads

supplied by the network are considered to be balanced.

Busbar 2

Upstream
EHV system Busbar 1 Busbar x Busbar n
… …

Busbar 3

HV sub-system

Figure 4.4: Interconnected HV sub-system

As seen in Section 4.2 above, the global voltage unbalance in HV networks is

determined not only by the local line asymmetries but also by the downstream MV

lines asymmetries when the network supplies considerable proportions of motor loads.
lines
Thus, the resultant negative sequence voltage V−:g/hv:x which arises as a result of the

line asymmetries at the busbar x can be generally written as [1]:

lines t12 t13 t23 ij t


V−:g/hv:x = (V−:g/hv:x + V−:g/hv:x + V−:g/hv:x + ... + V−:g/hv:x + ...)
td:1 td:2 td:x td:i
+ (V−:g/hv:x + V−:g/hv:x + ... + V−:g/hv:x + ... + V−:g/hv:x + ...)

(4.10)

where,
tij
V−:g/hv:x - negative sequence voltage caused by any local HV line tij on its own at

the busbar x
81
td:i
V−:g/hv:x - negative sequence voltage caused by any downstream MV line td:i on its

own at the busbar x

lines
Then, the emission Ug/hv:x and the factor K 0 uex can be expressed in the forms which

are similar to (3.17) and (3.1) respectively given in Chapter 3.

Busbar x

MV line
(td:x)

LVs:x

Downstream
system supplied by
the HV busbar x
LVr:x

Figure 4.5: System representation of any busbar x of the HV system shown in Fig. 4.4

As described in Chapter 3 (Section 3.3), the nodal negative sequence currents

([I−:lines ]) which arise as a result of the line asymmetries can be written in terms

of the nodal negative-positive sequence coupling admittances ([Y−+ ]), nodal positive
 
lines
sequence admittances ([Y++ ]) and nodal positive ([V+ ]) and negative [V−:g/hv ] se-

quence voltages as:


82

lines
−[I−:lines ] = [Y−+ ][V+ ] + [Y++ ][V−:g/hv ] (4.11)

The current I−:lines/x which is the xth element of [I−:lines ] or the negative sequence

nodal current at the busbar x can be generally written when the busbar supplies a

mix of passive (at HV, MV and/or LV) and motor (at LV) loads as:

lines
I−:lines/x ≈ Y−−:x−im V−:g/hv:x + Y−+:x V+:x (4.12)

The admittance Y−−:x−im is the downstream negative sequence admittance seen at

the busbar x taking into account only induction motors that are normally supplied

at the LV level, which is inductive in nature. As an example, this can be expressed27

for the simplified case where motor loads are supplied only by the MV line td:x at the

busbar LVr:x as:

√ !
klvr:x 3 |I+:x |
Y−−:x−imr ≈ −j   (4.13)
(1 − V Rtd:x )2 1
+ 1 Vn−hv
ksr:x kmr:x ksc−lvragg :x

where,

Y−−:x−imr - admittance Y−−:x−im corresponding to the simplified case stated above

V Rtd:x , ksr:x , kmr:x , klvragg :x , ksc−lvr:x - as defined for (4.9)28

Vn−hv - nominal line-line voltage of the HV system

The admittance Y−+:x is the downstream negative-positive sequence coupling ad-

mittance seen at the busbar x, which arises as a result of the asymmetry of the MV

line td:x . This admittance can be generally expressed29 by (4.14) and (4.15):
27
Derivation of (4.13) can be described using (D.1), (D.3) and (G.6).
28
Additional subscript ‘x’ indicates the quantities corresponding to the busbar x.
29
Refer to Appendix I for the derivation.
83

√ !
3 |I+:x |
|Y−+:x | ≈ ktd:x kn:x |Y−−:x−im Z−+:td:x −hv | (4.14)
Vn−hv
θY−+:x ≈ 900 + θZ−+:td:x + θpf :x (4.15)

where,

ktd:x , kn:x - as defined for (4.7)

Z−+:td:x −hv - negative-positive sequence coupling admittance, referred to HV, of the

MV line td:x

θY−+:x , θZ−+:td:x , - phase angles of the admittance Y−+:x and the negative-positive se-

quence coupling impedance Z−+:td:x of the MV line td:x respectively

θpf :x - pf angle30 at the busbar x

Substitution of (4.12) in (4.11) and rearrangement gives:

0
lines
[V−:g/hv ]n×1 ≈ −[Y++ ]−1
n×n [Y−+ ]n×n [V+ ]n×1 (4.16)

where,
0
Y++:xy ≈ Y++:xy + Y−−:x−im for x = y
0
Y−+:xy ≈ Y−+:xy + Y−+:x for x = y
0
Y++:xy = Y++:xy for x 6= y
0
Y−+:xy = Y−+:xy for x 6= y

That is, taking the nodal positive sequence voltages as known quantities, the negative
lines
sequence voltage V−:g/hv:x at any busbar x can be established using (4.16).

30
− and + for lagging and leading conditions respectively.
84

The presence of positive sequence voltage controlled components such as PV gen-

erators and synchronous condensers in a system force the negative sequence voltage

at the connected busbars to be zero, disregarding the existence of sources of unbal-

ance. Equation (4.16) which gives the nodal negative sequence voltages arising as a

result of line asymmetries does not consider the presence of such components, and

thus requires suitable adjustments such that the influence of zero voltage unbalance

(or of voltage controlled components) at given busbars on the emission levels at other

busbars is accommodated.

Consider that a voltage controlled component is connected at any busbar i of

the considered HV network (see Fig. 4.4). Hence, the negative sequence voltage V−:i

at the busbar i is zero, and there are only (n − 1) number of busbars at which the

negative sequence voltage to be determined. Then, by expanding the basic equation

(3.21) given in Chapter 3, the negative sequence current (labelled I−ci :x 31 ) at any

other busbar x can be written as:

I−ci :x = (Y−+:x1 V+:1 + ... + Y−+:xi V+:i + ... + Y−+:xn V+:n ) + (Y++:x1 V−:1 + ...

... + Y++:xh V−:h + Y++:xj V−:j + ... + Y++:xn V−:n ) (4.17)

Noting that the term associating the admittance Y++:xi is absent in (4.17), the ma-

trix equation (3.22) can be modified accordingly to incorporate the influence of the

constraint V−:i = 0 on the emission arising as a result of line asymmetries at other

busbars by:

0
• Reducing the dimension of the matrix [Y++ ] down to (n − 1) × (n − 1) by

removing both the ith row and column.


31
Subscript ci indicates the additional constraint of the controlled voltage unbalance at the bus-
bar i.
85
0
• Reducing the dimension of the matrix [Y−+ ] down to (n − 1) × n by removing

the ith row.

4.4 Verification of the Methodology Using a Three-bus Test

System

The proposed methodology is applied to the three-bus HV network (66kV, 60Hz,

three-wire) shown in Fig. 4.6. Considered operating scenario and resulting positive

sequence system conditions32 are also indicated in Fig. 3.7. Lengths of the HV

lines which are taken as identical in construction (including the phase positioning33 :

a b c) and untransposed are shown alongside the lines. Relevant admittance

data of the lines are34 :

• positive sequence admittance = (0.6265 − j2.3517)Skm

• negative-positive sequence coupling admittance = (0.1040 + j0.1779)Skm

(0.2061∠600 Skm)

Busbars 1 and 3 supply MV loads with equal compositions of constant impedance

and constant power elements directly at the HV-MV coupling transformers. This

implies that Y−−:x−im ≈ 0 and Y−+:x ≈ 0 for x = 1 and 3. Busbar 2 supplies loads at

the LV level through 3.2187km of untransposed MV lines35 , which account for 40%

of the total load supplied by the system. Two cases based on the type of LV loads

are considered:

32
Nodal voltages and line currents, which are obtained using load flow analysis.
33
This shows the considered arrangement of the three phase conductors (a, b and c) of the hori-
zontal tower.
34
Refer to Appendix H for further details.
35
Refer to Appendix B for the tower construction and conductor data.
86

208A
HV busbar 2
(1.01pu, -10.70) LV (460V)
10MVA
0.9 lagging pf
tap = 1.1pu
10MVA
0.9 lagging pf
50km MV lines –
Upstream EHV 130A 12.47kV
(230kV) system HV busbar 1 3.2187km
(1.1pu, 00) (1.07pu, -8.30) 150A
VR = 9%
40km
78A
20km
167A
tap = 1pu

tap = 1.03pu
10MVA
0.9 lagging pf
HV busbar 3
(1.04pu, -9.60)

10MVA 10MVA
0.9 lagging pf 0.9 lagging pf

EHV-HV coupling transformer – 60MVA, winding resistance = 1%, leakage reactance = 20%,
secondary tap setting = 1.1pu

HV-MV coupling transformers – 12MVA, winding resistance = 1%, leakage reactance = 10%

MV-LV coupling transformers – aggregated representation of fully loaded 1MVA transformers with
winding resistance = 1% , leakage reactance = 5%, and secondary tap setting = 1.08pu

Figure 4.6: Three-bus HV test system considered for applying the proposed method-
ology
87

• Case 1 - LV loads represent passive elements36 . That is, km:2 = 0, implying that

Y−−:2−im ≈ 0 and Y−+:2 ≈ 0.

• Case 2 - LV loads represent three-phase induction motors. That is, km:2 = 1.

Aggregation of 50hp motors of which the details are given in Appendix B is

considered. This results in an admittance Y−−:2−im ≈ −j0.0154S.

lines
Fig. 4.7 illustrates the emissions Ug/hv:x at the HV busbars established using the

proposed methodology37 in comparison to the results obtained using unbalanced load

flow analysis for the two cases listed above, clearly indicating the influence of the

downstream line asymmetries38 in each case. The considered phase positioning of

the MV lines39 is similar to that of the HV lines resulting in a Y−+:2 = (−0.0077 +

j0.1093) × 10−3 S.

lines
Fig. 4.8 illustrates the emission Ug/hv:x at each of the HV busbars established

using the proposed methodology and unbalanced load flow analysis for the case where

km:2 = 1 (i.e. when busbar 2 supplies motor loads) in relation to two different phase

arrangements of the MV lines, indicating their influence40 in each case on the resultant

emission levels.

• Phase positioning I - a b c (case considered in Fig. 4.7), of which

the negative-positive sequence coupling impedance and admittance values are

0.0349∠300 Ω/km and 0.1839∠820 Skm respectively as given in Appendix B.

Note that this phase positioning is as same as that of the local HV lines.

36
With equal compositions of constant impedance and constant power loads.
37
Application of the methodology to the test system is described in Appendix J.
38
Which is the difference between the results established using unbalanced load flow analysis with
and without introducing the effects of the downstream MV line asymmetries.
39
See Appendix B.
40
Which is the difference between the results established using unbalanced load flow analysis with
and without introducing the effects of the downstream MV line asymmetries.
88

0.7 Caused both by the HV and


MV lines for Case 1: load flow
0.6
Case 1: methodology

0.5
/ hv:x (%)

Caused both by the HV and


0.4 MV lines for Case 1: load flow
U glines

Caused only by the HV lines


0.3 for Case 2: load flow

0.2 Caused only by the HV lines


for Case 2: methodology
0.1
Caused both by the HV and
MV lines for Case 2: load flow
0.0
1 2 3 Caused both by the HV and
MV lines for Case 2:
Busbar methodology

lines
Figure 4.7: Emissions Ug/hv:x for the three-bus HV test system for the cases where
km:2 = 0 and km:2 = 1

• Phase positioning II - a c b, of which the negative-positive se-

quence coupling impedance and admittance values are 0.0349∠1500 Ω/km and

0.1839∠−1580 Skm respectively. This gives rise to an admittance Y−+:2 ≈

(−0.0908 − j0.0613) × 10−3 S.

Figs. 4.7 and 4.8 demonstrate that there is a good agreement between the results

obtained using the proposed technique and unbalanced load flow analysis. Further,

based on the results presented in Figs. 4.7 and 4.8, the following can be revealed:

• As in the case of MV networks discussed in Chapter 3, the emission/s arising as

a result of the local HV lines in the presence of considerable motor proportions

is higher at the busbar which is directly connected to the upstream system

(e.g. busbar 1 of the system shown in Fig. 4.7) and lower at all other busbars

(e.g. busbars 2 and 3 of the system shown in Fig. 4.7), compared to the case

where only passive loads exist.


89

0.6 Caused only by the HV lines:


load flow

0.5
Caused only by the HV lines:
methodology
/ hv: x (%)

0.4

Caused both by the HV and MV


0.3 lines for Phase positioning I:
U glines

load flow

0.2 Caused both by the HV and MV


lines for Phase positioning I:
methodology
0.1 Caused both by the HV and MV
lines for Phase positioning II:
load flow
0.0
1 2 3 Caused both by the HV and MV
lines for Phase positioning II:
Busbar methodology

lines
Figure 4.8: Emissions Ug/hv:x for the three-bus HV test system for the case where
km:2 = 1 in relation to the Phase arrangements I and II of the MV lines

• In addition, the global emission in HV networks is dependant on the downstream

line asymmetries in the presence of considerable motor proportions as seen

in Section 4.2 above. This influence of the downstream MV lines can either

decrease (e.g. for Phase positioning I in Fig. 4.8) or increase (e.g. for Phase
lines
positioning II in Fig. 4.8) the resultant emission Ug/hv:x compared to the local

emission levels depending on the impedance/admittance characteristics of the

downstream lines relative to the local lines.

4.5 Verification of the Methodology Using the IEEE 14-bus

Test System

The proposed methodology is further applied to the IEEE 14-bus test system shown

in Fig. 4.9 which consists of positive sequence voltage controlled busbars (busbars

1, 2, 3, 6 and 8), taking it as a 66kV, 60Hz and three-wire network supplying constant
90

power loads at the HV level. The system data are as per [77] with appropriate

and minor modifications, which are given in Appendix K together with the nodal

positive sequence voltages41 . Lines are taken as identical in construction (including

phase positioning) and untransposed, of which the relevant admittances per km as

pu quantities (on a 100M V A base) are42 :

• Positive sequence admittance = (0.2729 + j1.0244) × 102 pu

• Negative-positive sequence coupling admittance = (0.4530 + j0.7749) × 10pu

Lengths43 of the lines are established such that their positive sequence impedance

magnitudes (in pu) are approximately as per [77].

lines
Fig. 4.10 illustrates the emission Ug/hv:x at each of the HV busbars established

using the methodology in comparison to the results obtained using unbalanced load

flow analysis, further validating the proposed technique.

4.6 Chapter Summary

As a continuation of the work presented in Chapter 3, this chapter has addressed

the global voltage unbalance in HV power systems which arises as a result of line

asymmetries. This is a key aspect in assessing emission limits to individual installa-

tions connected to HV power systems essentially based on the IEC/TR 61000-3-13

recommendations.

The dependency of the global emission on the local HV lines as well as on the

downstream MV lines in the presence of passive and induction motor loads has been

examined in relation to a simple radial network. The following major conclusions can

be drawn from the study:


41
Which are obtained using load flow analysis.
42
See Appendix H for further details.
43
Given in Appendix K.
91

Please see print copy for image

Figure 4.9: IEEE 14-bus test system

0.5
Load flow results
Methodology
0.4
/ hv:x (%)

0.3
U glines

0.2

0.1

0.0
1 2 3 4 5 6 7 8 9 10 11 12 13 14

Busbar

lines
Figure 4.10: Emissions Ug/hv:x for the IEEE 14-bus test system
92

• The direction given in IEC/TR 61000-3-13 to assess the influence of an asym-

metrical radial line on the global emission can be applied also to an HV system

however only when the network supplies primarily passive loads. In this case,

the local HV lines are totally responsible for the global emission, and the con-

tribution made by the downstream MV lines is negligible.

• The presence of induction motors has been seen to affect the level of emission

given by the IEC approach noticeably. This dependency of the global emission

on the motor proportion can be seen in two forms:

– As in the case of MV networks, the local emission or the emission arising

as a result of HV lines is influenced by motor loads.

– In addition, the presence of motor loads makes the downstream emission

or the emission arising as a result of MV lines accountable for the global

emission in HV networks.

A systematic approach, covering interconnected environments, for evaluating the

global emission caused both by the local and downstream line asymmetries at nodal

level has been proposed. The results established using the proposed methodology

in relation to a three-bus test system has been seen to be in close agreement with

the results obtained using unbalanced load flow analysis. Furthermore, these results

clearly demonstrated the following:

• As in the case of MV networks, the emission/s arising as a result of the local

HV lines in the presence of considerable motor proportions is higher at the HV

busbar which is directly connected to the upstream system and lower at all other

busbars of the network, compared to the case where only passive loads exist.

• The influence of the downstream MV lines can either decrease or increase the

resultant emission levels with respect to the local emission levels depending on
93

the impedance/admittance characteristics of the downstream lines relative to

the local lines.

The proposed methodology is further validated in relation to the IEEE 14-bus test

system.
Chapter 5

Propagation of Voltage Unbalance

5.1 Introduction

As discussed in Chapter 2 (Section 2.8.3), the propagation of voltage unbalance is a

key aspect considered in the IEC/TR 61000-3-13 allocation procedure. Quantitative

measures of this propagation exist in two forms:

• Transfer coefficients1 which give a measure of the propagation from upstream

higher voltage to downstream lower voltage systems through coupling trans-

formers. This is employed in the allocation procedure in determining the global

emission allowance Ug/s of any sub-system S to quantify the level of voltage

unbalance which propagates from the upstream sub-system.

• Influence coefficients which give a measure of the propagation from one busbar

to another busbar of a sub-system at a particular voltage level through lines.

The influence coefficient ki−x between busbars i and x is defined as the voltage

unbalance which arises at the busbar x when 1pu of negative sequence voltage

source is applied at the busbar i. This is employed in the allocation procedure in


1
See Section 2.8.4.

94
95

determining the total available apparent power Stotal−x of the entire sub-system

as seen at the busbar x under evaluation to take into account the contributions

from neighbouring busbars.

1.15

1.10

1.05
Tmv-lv

1.00

0.95

0.90
10 13 16 19 22 25
ksc-lv

Figure 5.1: Variation of Tmv−lv with ksc−lv obtained for constant power loads using
unbalanced load flow analysis

The method given in IEC/TR 61000-3-13 for the evaluation of the MV to LV

transfer coefficient Tmv−lv , which is reproduced in Section 2.8.4 by (2.17), assumes

a unity transfer coefficient in relation to passive loads in general. A transfer coeffi-

cient = 1 is mathematically trivial for constant impedance loads. However, it may not

be valid for other load types such as constant power and constant current loads, owing

to the different behaviours exhibited by these load types under unbalanced supply

conditions as noted in Chapter 3 (Section 3.2). As an example, Fig. 5.1 illustrates

the variation of the transfer coefficient Tmv−lv with ksc−lv 2 ) established when an LV

system supplies a load base primarily having constant power loads with 0.9 lagging pf

using unbalanced load flow analysis, compared against unity. Noting that the transfer

coefficient Tmv−lv is considerably greater than unity at relatively lower ksc−lv values
2
As defined in Chapter 2 for (2.17), ksc−lv is the ratio between the short-circuit capacity (in
MVA) at an LV busbar and the total load (in MVA) supplied by the LV busbar.
96

(i.e. heavily loaded LV systems), it is evident that the assumption of a unity transfer

coefficient for passive loads used in the IEC method cannot be generally applied with

a high degree of accuracy. That is, a requirement exists for cautious examination

of the dependency of transfer coefficients on various load types/bases. Moreover,

systematic approaches for assessing other transfer coefficients (e.g. HV to MV) and

influence coefficients are not covered in IEC/TR 61000-3-13.

Objectives of the work presented in this chapter are:

• To develop theoretical bases which describe the behaviour exhibited by various

load types with regard to the propagation of voltage unbalance from higher

voltage to lower voltage systems, and to propose improved/novel approaches for

evaluating the MV to LV and HV to MV transfer coefficients. This is covered

in Section 5.2.

• To carry out preliminary studies in order to investigate the dependency of in-

fluence coefficients on various load types/bases, and to develop a generalised

methodology for their evaluation covering interconnected network environments.

Section 5.3 presents this, together with a comparison of the results obtained us-

ing the proposed technique and unbalanced load flow analysis in relation to an

MV three-bus test system and the IEEE 14-bus test system.


97

5.2 Voltage Unbalance Transfer Coefficients

Considering the system shown in Fig. 5.23 where the US (upstream higher voltage

system) to S (lower voltage system under evaluation) transfer coefficient Tus−s which

is defined in Section 2.8.4 by (2.16) can be written in an expanded form as:



V−:Uus /s−us
V−:Uus /s
× 1 = V−:us

Tus−s = (5.1)
V+:s V−:us V+:s−us
V+:us V+:us

where,

V+:s , V+:us - positive sequence voltages at the busbars S and US respectively

V−:us - negative sequence voltage at the busbar US

V−:Uus /s - negative sequence voltage at the busbar S which is transferred from US

V+:s−us , V−:Uus /s−us - V+:s and V−:Uus /s respectively referred to US

US

Uus/s

Ss-local

DS

Ss-ds

Figure 5.2: Radial system considered for the illustration of transfer coefficients

3
Reproduction of Fig. 2.5.
98

The positive sequence voltage ratio VV+:s−us can be expressed in a general form,

+:us

disregarding the load type, as:


V+:s−us 1
V+:us ≈ (5.2)

Z −s
1 + j ++:tf ∠θ

Z++:s pf :s

where,

Z++:tf −s - positive sequence impedance (assumed as inductive), referred to S, of the

US-S coupling transformer

Z++:s - downstream positive sequence impedance (or equivalent) seen at the busbar S

θpf :s - pf angle4 at the busbar S


Z++:tf −s
The impedance ratio Z++:s can be written as:


Z++:tf −s 1
Z++:s = ksc−s (5.3)

where,
Ssc−s
ksc−s = Ss
(Vn−s )2
Ssc−s = |Z++:tf −s |
, the short-circuit capacity (in MVA) at the busbar S
(Vn−s )2
Ss = |Z++:s |
, the total load (in MVA) supplied by the system S

Vn−s - nominal line-line voltage of the system S


Then, the positive sequence voltage ratio VV+:s−us can be rewritten as:

+:us


V+:s−us 1
V+:us ≈ (5.4)

1
1 + j ksc−s ∠θpf :s

4
− and + for lagging and leading conditions respectively.
99

The negative sequence voltage |V−:Uus /s−us | can be generally written, ignoring zero

sequence unbalance, as:

|V−:Uus /s−us | = |V−:us − (Z−−:tf −us I−:Uus /tf )| (5.5)

where,

Z−−:tf −us - negative sequence impedance5 , referred to US, of the US-S coupling trans-

former

I−:Uus /tf - negative sequence current (referred to US) in the US-S coupling trans-

former, which arises as a result of the voltage unbalance at the busbar US

Based on the evidence from Chapter 3 (Section 3.2), the behaviour of the negative

sequence current I−:Uus /tf seem to be influenced by the load type/base supplied by

the system S making the transfer coefficient Tus−s dependant on the load type/base.

The IEC approach6 for assessing the MV to LV transfer coefficient accounts for this

influence, however it distinguishes only motor loads from passive loads, or in other

words it does not take the differences7 that exist between various passive load types

into account. This section addresses four basic load types8 and various load bases

developing theoretical bases which describe their behaviours in this regard.

Constant Impedance (Z) loads

When the system S supplies constant impedance loads (balanced), (5.5) can be re-

V
arranged such that the negative sequence voltage ratio −:UVus /s−us
−:us
is equal to the
5
Which is inherently equal to the positive sequence impedance Z++:tf −us of the transformer.
6
Given in Chapter 2 by (2.17).
7
As seen from Fig. 5.1.
8
i.e. constant impedance, constant current, constant power and three-phase induction motor
loads.
100

positive sequence voltage ratio given by (5.4), as the positive and negative sequence

impedances of the loads and the US-S coupling transformer are equal:


V−:Uus /s−us 1
V−:us = (5.6)

1
1 + j ksc−s ∠θpf :s

Substitution of (5.4) and (5.6) in (5.1) gives:

Tus−s = 1 (5.7)

That is, for constant impedance loads, the term Z++:tf −us I−:Uus /tf causes the down-

stream negative sequence voltage |V−:Uus /s−us | to be smaller than the upstream neg-
1
ative sequence voltage |V−:us | by the factor ˛
1
˛, leading to a unity Tus−s .
∠θpf :s ˛
˛ ˛
˛1+j k
sc−s

Constant Current (I) loads

As in Section 3.2.2, the negative sequence current I−:Uus /tf can be considered to be

negligible when the system supplies constant current loads (balanced), as such loads

draw equal magnitudes of three phase currents regardless of the prevailing voltage

condition. Hence, (5.5) can be simplified for constant current loads as:


V−:Uus /s−us
V−:us ≈ 1 (5.8)

Substitution of (5.4) and (5.8) in (5.1) gives:


1
Tus−s ≈ 1 + j ∠θpf :s (5.9)
ksc−s

That is, in contrary to the case of constant impedance loads, the downstream negative

sequence voltage |V−:Uus /s−us | is equal to the upstream negative sequence voltage
101

|V−:us | in the presence of constant current loads, resulting in an increase in Tus−s by



1
the factor 1 + j ksc−s ∠θpf :s relative to unity.

Constant Power (P Q) Loads

As in Section 3.2.3, through careful examination of the results obtained using unbal-

anced load flow analysis, (5.10) is established as a close approximation to the negative

V
sequence voltage ratio −:UVus /s−us
−:us
:


V−:Uus /s−us 1
V−:us ≈ (5.10)
β
1
1 + j ksc−s ∠θpf :s

where, β ≈ −1 and −2 for low (∼ 0.9) and high (∼ 1) lagging pf conditions respec-

tively. Substitution of (5.4) and (5.10) in (5.1) gives:

1−β
1
Tus−s ≈ 1 + j
∠θpf :s (5.11)
ksc−s

That is, for constant power loads, the term Z++:tf −us I−:Uus /tf causes the downstream

negative sequence voltage |V−:Uus /s−us | to be greater than the upstream negative se-
1
quence voltage |V−:us | by the factor ˛˛ ˛β , resulting in an increase in Tus−s
˛1+j k 1 ∠θpf :s ˛
˛
sc−s
1−β
1
by the factor 1 + j ksc−s ∠θpf :s relative to unity.

Discussion

Taking a generalised view, the scaling factors applied to the downstream negative

sequence voltage |V−:Uus /s−us | relative to the upstream negative sequence voltage
γ
1
|V−:us | can be expressed in a form 1 + j ksc−s ∠θpf :s for the three passive load types,

where γ = 1 for constant impedance loads, γ = 0 for constant current loads, and

−2 ≤ γ ≤ −1 for constant power loads. Note that this exponent γ, which distin-
102

guishes various load behaviours, is similar to that established in Section 3.2 in relation

to the emission arising as a result of an asymmetrical radial line. Alternatively, the


τ
1
transfer coefficient Tus−s can be expressed in a form 1 + j ksc−s ∠θpf :s , where τ = 0

for constant impedance loads, τ = 1 for constant current loads, and 2 ≤ τ ≤ 3 for

constant power loads. Noting that the factor ksc−s can be in the range of 59 to 2510

for various systems and the exponent τ can vary in the range of 0 to 3 for different

load types, a uniform behaviour or the behaviour of constant impedance loads as

assumed in the IEC approach cannot be used to represent all load types with a high
τ
1
degree of accuracy. The accuracy of the formulation Tus−s ≈ 1 + j ksc−s ∠θpf :s will

be demonstrated in Sections 5.2.1 and 5.2.2 as applicable to MV to LV and HV to

MV transfer coefficients respectively.

Induction Motor (IM ) Loads

When the system S supplies three-phase induction motors11 which can be repre-

sented using decoupled, unequal and constant12 sequence impedances, (5.5) can be

written as:

V−:Uus /s−us 1
V−:us = (5.12)

Z++:tf −s
1 + Z−−:s

where, Z−−:s - downstream negative sequence impedance seen at the busbar S. When

US and S represent MV and LV (i.e. s = lv) systems respectively (i.e. the case of

the MV to LV propagation), the impedance Z−−:lv is equal to the negative sequence

impedance Z−−:im of the aggregated motor load supplied by the aggregated LV sys-

tem13 . In the case of the HV to MV (i.e. s = mv) propagation, the impedance Z−−:mv
9
e.g. fully loaded 60MVA HV-MV transformer with 20% impedance.
10
e.g. fully loaded 400kVA MV-LV transformer with 4% impedance.
11
Which are usually supplied at the LV level.
12
For a given motor speed.
13
Note that the busbar DS of Fig. 2.5, which represents the downstream system of the system S,
does not exist in this case.
103

is not simply equal to Z−−:im due to the additional system impedance Z++:sys−lv 14

that exists between the MV busbar under evaluation and the downstream LV systems

at which motor loads are supplied. That is, transfer coefficients, in the presence of

motor loads, seem to depend on the systems in which the propagation is being con-

sidered. This will be further discussed separately in relation to the MV to LV and

HV to MV propagation in Sections 5.2.1 and 5.2.2 respectively.

5.2.1 MV to LV Transfer Coefficient, Tmv−lv

This section considers that the busbars US and S (Fig. 5.2) represent MV and LV

systems respectively (i.e. the subscripts ‘us’ and ‘s’ used in the above formulae (5.1)

- (5.12) are to be replaced with ‘mv’ and ‘lv’ respectively).

Passive Loads
τ
1
As derived above, Tmv−lv ≈ 1 + j ksc−lv ∠θpf :lv for passive loads. Representing most

practical circumstances, the factor ksc−lv for LV systems can take a value in the

range of 1015 to 2516 . Figs. 5.3: I − II and 5.4: I − II illustrate the variation of

Tmv−lv with ksc−lv established using this formulation in comparison to the results

obtained using unbalanced load flow analysis for constant current and constant power

loads respectively, where sub-figures I and II correspond to 0.99 and 0.9 lagging pf

conditions respectively. These illustrate the accuracy of the new formulation, while

demonstrating the deviation of the actual transfer coefficient from the unity value

assumed in the IEC approach.


14
e.g. MV-LV coupling transformer and MV line impedances. The subscript ‘lv’ indicates the
impedance referred to the LV level.
15
e.g. fully loaded 10MVA transformer with 10% impedance.
16
e.g. fully loaded 400kVA transformer with 4% impedance.
104

1.10 I - 0.99 lagging pf 1.10 II - 0.9 lagging pf

1.05 1.05
Tmv-lv

Tmv-lv
1.00 1.00

0.95 Load flow results 0.95 Load flow results


Proposed formulation Proposed formulation
0.90 0.90
10 13 16 19 22 25 10 13 16 19 22 25
ksc-lv ksc-lv

Figure 5.3: Variation of Tmv−lv with ksc−lv for constant current loads: I - 0.99 lagging
pf, II - 0.9 lagging pf

1.15 I - 0.99 lagging pf 1.15 II - 0.9 lagging pf

1.10 1.10

1.05 1.05
Tmv-lv

Tmv-lv

1.00 1.00
Load flow results Load flow results
0.95 0.95
Proposed formulation Proposed formulation

0.90 0.90
10 13 16 19 22 25 10 13 16 19 22 25
ksc-lv ksc-lv

Figure 5.4: Variation of Tmv−lv with ksc−lv for constant power loads: I - 0.99 lagging
pf, II - 0.9 lagging pf
105

Induction Motor (IM ) Loads

Equation (5.12) is reproduced here for the propagation of the negative sequence volt-

age from MV to LV as:


V−:Umv /lv−mv 1
= (5.13)
V−:mv Z++:ml−lv
1 + Z−−:im

where, the subscript ‘ml’ in Z++:ml−lv is a replacement of the subscript ‘tf ’ used

in Z++:tf −s to specifically indicate the MV-LV coupling transformer. Noting the

inductive nature of the impedances Z++:ml−lv and Z−−:im , (5.13) can be rewritten as:


V−:Umv /lv−mv 1
=   (5.14)
V−:mv ks
1 + ksc−lv

where, ks - ratio between the positive and negative (which is inductive) sequence

impedances of the aggregated motor load supplied by the LV system17 . Substitution

of (5.4) and (5.14) in (5.1) gives:



1
1 + j ksc−lv ∠θpf :lv

Tmv−lv ≈   (5.15)
ks
1 + ksc−lv

That is, for induction motor loads, the term Z++:ml−mv I−:Umv /ml causes the down-

stream negative sequence voltage |V−:Umv /lv−mv | to be smaller than the upstream neg-
1
ative sequence voltage |V−:mv | by the factor „
ks
«. Noting that 5 < ks < 7, this
1+ k
sc−lv
results in a Tmv−lv < 1. This reduction in Tmv−lv relative to unity is significant com-

pared to the increment in Tmv−lv caused by passive loads18 . Fig. 5.5 illustrates the

variation of Tmv−lv with ksc−lv , for motor loads with ks = 6.719 and pf = 0.9 lagging,
17
Typically, ks can be in the range of 5 to 7.
18
e.g. Tmv−lv = 0.81 for motor loads with ks = 6.7, ksc−lv = 25 and pf = 0.9 lagging, whereas
Tmv−lv = 1.04 for constant power loads with ksc = 25 and pf = 0.9 lagging.
19
e.g. 50hp motors described in Appendix B.
106

established using the IEC method20 and (5.15) in comparison to the results obtained

using unbalanced load flow analysis. This demonstrates that both the IEC method

and the proposed new formulation are satisfactory for estimating Tmv−lv for mo-

tor loads.

1.10

1.00

0.90
Tmv-lv

0.80

0.70 Load flow results

0.60 Prposed formulation


IEC method
0.50
10 13 16 19 22 25
ksc-lv

Figure 5.5: Variation of Tmv−lv with ksc−lv for induction motor loads with ks = 6.7
and pf = 0.9 lagging

Generalisation for Mixes of Various Load Types

Consider a mix of constant impedance, constant current, constant power and induc-

tion motor loads is supplied by the LV system. Then, the negative sequence current

I−:Umv /ml can be decomposed as:

I−:Umv /ml = I−:z + I−:i + I−:pq + I−:im (5.16)

where, I−:z , I−:i , I−:pq , I−:im - negative sequence currents (referred to MV side) in Z,

I, P Q and IM loads respectively arising as a result of the MV unbalance. Employing


20
Given by (2.17).
107

(5.16), (5.5) can be written in an expanded form as:



X
|V−:Umv /lv−mv | = V−:mv − (Z++:ml−mv I−:L ) (5.17)


L=z,i,pq,im

The impact of the components Z++:ml−mv I−:L for the four load elements can be com-

bined, forming the resultant influence on the propagation, as21 :


V−:Umv /lv−mv 1
≈ β  (5.18)
V−:mv

kpq
kz k m ks
1 + j ksc−lv ∠θpf :z 1 + j ksc−lv ∠θpf :pq 1 +

ksc−lv

where,

km - ratio between the rated motor load (in MVA) and the total load (in MVA) sup-

plied by the LV system

kz , kpq - ratios of the constant impedance and constant power loads (in MVA) respec-

tively to the total load (in MVA) supplied by the LV system

θpf :z , θpf :pq - power factor angles of the constant impedance and constant power loads

respectively supplied by the LV system

β  
kz kpq km k s
The terms 1 + j ksc−lv ∠θpf :z , 1 + j ksc−lv ∠θpf :pq and 1 + in the numera-

ksc−lv

tor of (5.18) account for the influence of the constant impedance, constant power and

induction motor elements respectively in the mix on the propagation of the negative

sequence voltage. Note that these terms are modified versions of the respective terms

for the individual load types22 , where the modifications are introduced by multiply-
 
1
ing ksc−lv of each of the terms for the individual load types by the respective load

proportion kL (L = z, pq, im). Substituting (5.4) and (5.18) in (5.1), the transfer
21
See Appendix
L for details.
γ
22 1
i.e. 1 + j ksc−lv ∠θpf :lv with different γ values for the various passive loads types, and

 
ks
1 + ksc−lv for motor loads.
108

coefficient Tmv−lv can be expressed in a generalised form as:



1
1 + j ksc−lv ∠θpf :lv

Tmv−lv ≈ β   (5.19)
kz kpq km k s
1 + j ksc−lv ∠θpf :z 1 + j ksc−lv ∠θpf :pq 1 +

ksc−lv

Figs. 5.6: I − II illustrate the variation of Tmv−lv with ksc−lv for two load bases

which are dominated by motor loads (Z − 10%, I − 5%, P Q − 15% and IM − 70%)

and passive loads (Z − 25%, I − 5%, P Q − 60%, IM − 10%) respectively established

using the IEC method, the proposed formulation (5.19) and unbalanced load flow

analysis. A lagging pf of 0.9 for all load components and a ks = 6.7 for motor loads

are assumed. These results demonstrate that although the IEC method provides an

accurate estimation to Tmv−lv for load bases containing large proportions of induc-

tion motors, it associates a considerable degree of error for load bases dominated by

passive elements. Furthermore, the proposed new formulation gives a more accurate

estimation to Tmv−lv for both of the loads bases.

0.90 1.05 II - Dominated by passive elements


I - Dominated by induction motors

0.85
1.00
Tmv-lv

Tmv-lv

0.80
Load flow results 0.95 Load flow results
0.75 Proposed formulation Proposed formulation
IEC method IEC method
0.70 0.90
10 13 16 19 22 25 10 13 16 19 22 25
ksc-lv
ksc-lv

Figure 5.6: Variation of Tmv−lv with ksc−lv : I - for a load base dominated by induction
motors, II - for a load base dominated by passive elements
109

Figs. 5.7: I−II illustrate the variation23 of Tmv−lv with km , considering load mixes

of constant impedance and motor loads (i.e. km = 1 − kz ) and constant power and

motor loads (i.e. km = 1 − kpq ) respectively, for two extreme cases where ksc−lv ≈ 10

and ksc−lv ≈ 25. A lagging pf of 0.9 for all load components and ks = 6.7 for motor

loads are assumed. These demonstrate that the transfer coefficient Tmv−lv can be in

the range of 0.6 to 1.1.

I - Load mixes of Z and IM loads II - Load mixes of PQ and IM loads


1.1 1.1

1.0 1.0 ksc-lv ≈ 25


ksc-lv ≈ 25

0.9 0.9
Tmv-lv

Tmv-lv
0.8 0.8

0.7 ksc-lv ≈ 10 0.7 ksc-lv ≈ 10

0.6 0.6
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
km = 1-kz km = 1-kpq

Figure 5.7: Variation of Tmv−lv with km for ksc−lv ≈ 25 and ksc−lv ≈ 10: I - for load
mixes of Z and IM loads, II - for load mixes of P Q and IM loads

The generalised expression (5.19) can be simplified, easing the computation, for

industrial load bases which contain large proportions of motor loads by neglecting

the negative sequence current components I−:z , I−:i , and I−:pq compared to I−:im in

(5.16) as:
1
1 + j ksc−lv ∠θpf :lv

Tmv−lv ≈   (5.20)
1 + kksc−lv
m ks

Fig. 5.8 provides a comparison of the variations of Tmv−lv with km established using

the IEC method, the generalised expression (5.19), the simplified expression (5.20)

for industrial load bases, and unbalanced load flow analysis. Load mixes of constant
23
Derived using the proposed formulation.
110

power and motor loads, a lagging pf of 0.9, ks = 6.7 and ksc−lv ≈ 10 are considered.

This shows that the estimation from the various methods are in close agreement for

km values above 0.5. For lower km values, (5.19) gives the closest estimation to the

actual24 Tmv−lv , and (5.20) is seen to provide a better estimation compared to the

IEC method.

1.2 Load flow results


Eq. (5.19)
1.1 Eq. (5.20)
IEC method
1.0
Tmv-lv

0.9

0.8

0.7

0.6
0.0 0.2 0.4 0.6 0.8 1.0
km

Figure 5.8: Variation of Tmv−lv with km established using the IEC method, (5.19),
(5.20) and unbalanced load flow analysis

5.2.2 HV to MV Transfer Coefficient, Thv−mv

This section considers that the busbars US, S and DS of Fig. 5.2 represent HV, MV

and LV25 systems respectively (i.e. the subscripts ‘us’ and ‘s’ in the above formulae

(5.1) - (5.12) are to be replaced with ‘hv’ and ‘mv’ respectively).


24
i.e. the values obtained using unbalanced load flow analysis.
25
This is an aggregated representation of all LV systems supplied by the upstream MV system.
111

Passive Loads
τ
1
As derived above, Thv−mv ≈ 1 + j ksc−mv ∠θpf :mv for passive loads. Representing

most practical circumstances, the factor ksc−mv for MV systems can take a value in

the range of 526 to 1527 . Noting that the values of ksc−mv for MV systems are usually

smaller than the ksc−lv values for LV systems, the amplification which takes place in

the presence of passive loads in the HV to MV propagation is greater than that in

the case of the MV to LV propagation.

Induction Motor (IM ) Loads

Considering that the MV system under evaluation supplies only induction motors at

the aggregated LV busbar, (5.12) is reproduced here for the HV to MV propagation as:


V−:Uhv /mv−hv 1
= (5.21)
V−:hv
1 +
Z++:hm−mv
Z−−:mv

where, the subscript ‘hm’ in Z++:hm−mv is a replacement of the subscript ‘tf ’ used

in Z++:tf −s to specifically indicate the HV-MV coupling transformer. Noting that

the impedance Z−−:mv is equal to the sum of the impedances Z−−:im and Z++:sys−lv

(referred to MV), the rearrangement of (5.21) in terms of the load and system char-

acteristics gives :


V−:Uhv /mv−hv 1
=    (5.22)
V−:hv 
1

ks
1 + ksc−mv 1+ ks
ksc−lv
agg

26
e.g. fully loaded 60MVA transformer with 20% impedance.
27
e.g. fully loaded 10MVA transformer with 6.5% impedance.
112

where, ksc−lvagg - ratio between the short-circuit capacity28 (in MVA) at the aggregated

LV busbar, and the total load (in MVA) supplied by the LV system. Substitution of

(5.4) and (5.22) in (5.1) gives:



1
1 + j ksc−mv ∠θpf :mv

Thv−mv ≈    (5.23)
1 ks
1 + ksc−mv 1+ ks
ksc−lv
agg

That is, for induction motor loads, the term Z++:hm−hv I−:Uhv /hm causes the down-

stream negative sequence voltage |V−:Uhv /mv−hv | to be smaller than the upstream neg-

ative sequence voltage |V−:hv | by the factor 2 01 13 . Noting that


“ ”
41+ 1 @ ks A5
ksc−mv 1+ ks
ksc−lv
  agg

ks
1+ k ks > 1, a value less than unity can be expected for the transfer coefficient
sc−lvagg

Thv−mv . Similar to the case of the MV to LV propagation, this reduction in Thv−mv

relative to unity is significant compared to the increment in Thv−mv introduced by

passive loads29 . Note that the degree of this reduction in the HV to MV transfer

coefficient Thv−mv is lower than that in the HV to MV transfer coefficient Tmv−lv for

similar system and load characteristics30 . However, usually ksc−mv < ksc−lv , and thus

a higher degree of this reduction can be expected in the HV to MV propagation than

that in the MV to LV propagation31 .


28
Which is derived using the positive sequence system impedance Z++:sys−lv that exists between
the MV busbar under evaluation and the downstream LV busbar.
29
e.g. Thv−mv = 0.79 for motor loads with ks = 6.7, ksc−mv = ksc−lvagg = 15 and pf = 0.9 lagging,
whereas Tmv−lv = 1.06 for constant power loads with ksc−mv = 15 and pf = 0.9 lagging.
30
e.g. Tmv−lv = 0.71 for ks = 6.7, ksc−lv = 15 and pf = 0.9 lagging, and Thv−mv = 0.79 for
ks = 6.7, ksc−mv = ksc−lv = 15 (i.e. the HV-MV transformer supplies a similar MV-LV transformer.)
and pf = 0.9 lagging.
31
e.g. Thv−mv = 0.76 and Tmv−lv = 0.81 for ks = 6.7, ksc−mv = 15, ksc−lv = ksc−lvagg = 25 and
pf = 0.9 lagging.
113

Generalisation for Mixes of Various Load Types

Consider that the MV system under evaluation supplies a mix of constant impedance,

constant current, constant power loads (at MV and/or LV) and induction motors (at

LV). Based on the same approach32 used to establish (5.18) in the case of the MV to

LV propagation, the various load behaviours in the above load mix can be combined33 ,

forming the resultant impact of the load mix on the HV to MV propagation, as:


V−:Uhv /mv−hv

V−:hv
1 2 0 13
˛ ˛˛ ˛β “ ”
kzmv kpqmv kmmv ks
∠θpf :zmv ˛˛1+j k ∠θpf :pqmv ˛ 1+ k
˛ ˛˛ ˛ 4 @ A5
˛1+j k
sc−mv sc−mv sc−mv 1+ km ks
ksc−lv
agg

(5.24)

where,

kzmv , kpqmv , kmmv - ratios of the constant impedance, constant power and motor loads

(in MVA) respectively to the total load (in MVA) supplied by the MV system under

evaluation

θpf :zmv , θpf :pqmv - power factor angles of the constant impedance and constant power

loads respectively supplied by the MV system

Substituting (5.4) and (5.24) in (5.1), and replacing kmmv = km klv (where klv is

the fraction of LV loads supplied by the MV system under evaluation), the transfer

coefficient Thv−mv can be expressed in a generalised form as:


32
Described in Appendix L.
33 1 1 kLmv
Replacing ksc−mv and/or ksc−lv of the respective terms for the individual load types by ksc−mv
agg
kL
and ksc−lvagg respectively for L = z, pq, im.
114


1
1 + j ∠θ

ksc−mv pf :mv
Thv−mv ≈ β    
kzmv kpqmv km klv ks
1 + j ksc−mv ∠θpf :zmv 1 + j ksc−mv ∠θpf :pqmv 1 + ksc−mv

km ks
1+ ksc−lv
agg
(5.25)

Figs. 5.9: I - II and 5.10: I - II illustrate the variation of Thv−mv with klv for two

extreme cases where ksc−mv = 12 and ksc−mv = 4 respectively considering systems

where the loads are supplied directly at the MV busbar34 . Sub-figures I and II cor-

respond to load mixes of constant impedance and motor loads, and constant power

and motor loads respectively. Each figure shows the variation established using (5.25)

in comparison to the result obtained using unbalanced load flow analysis for three

sub-cases where:

• km = 1 (i.e. LV system supplies only motor loads)

• km = 0.5 (i.e. LV system supplies equal proportions of passive and motor loads)

• km = 0 (i.e. LV system supplies only passive loads)

A case where ksc−lvagg = 2035 , pf = 0.9 lagging for all load components and ks = 6.7

for induction motors is considered.

Figs. 5.11: I - II illustrate the variation of Thv−mv with klv for the two cases where

ksc−mv = 12 and ksc−mv = 4 respectively considering systems where the LV loads are

supplied through MV lines36 . Each figure shows the variation for the three sub-cases

km = 1, km = 0.5 and km = 0 established using (5.25) in comparison to the result

obtained using unbalanced load flow analysis. Passive loads are represented as a mix

of constant impedance and constant power elements with equal compositions. A case
34
LV loads are supplied through MV-LV transformers, i.e. Z++:sys−lv accounts for the impedances
of the MV-LV coupling transformers.
35
e.g. aggregation of fully loaded 1MVA transformers with 5% impedance.
36
i.e. Z++:sys−lv accounts for the impedances of the MV lines and the MV-LV coupling transform-
ers.
115

where ksc−lvagg = 637 , pf = 0.9 lagging for all load components and ks = 6.7 for

induction motors is considered.

According to Figs. 5.9 - 5.11, the proposed formulation (5.25) provides an estima-

tion to the transfer coefficient Thv−mv , which is seen to be in close agreement with the

values given by unbalanced load flow analysis except for the minor discrepancies38

arise when a system supplies primarily constant power loads under heavy loading

or lower ksc−mv conditions. Further, these demonstrate that the transfer coefficient

Thv−mv can be a value in the range of 0.5 to 1.4, which is a wider range than that of

Tmv−lv (0.6 - 1.1), depending on the prevailing system and load characteristics and

downstream load composition.

1.2 I - Load mixes of Z and IM loads 1.2 II - Load mixes of PQ and IM loads
1.1 1.1
km = 0
km = 0.5
1.0 1.0 km = 0
km = 0.5
Thv-mv

Thv-mv

0.9 0.9
km = 1
0.8 0.8
Load flow results Load flow results
0.7 Proposed formulation 0.7
km = 1 Proposed formulation
0.6 0.6
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
klv klv

Figure 5.9: Variation of Thv−mv with klv for ksc−mv = 12 (loads are supplied directly
at the MV busbar): I - for load mixes of Z and IM loads, II - for load mixes of PQ
and IM loads

37
e.g. the MV lines described in Appendix B supplying 10MVA loads per line with 10% voltage
regulation through fully loaded 1MVA transformers with 5% impedance and 1.1pu secondary tap
setting.
38
10% error in the case considered in Fig. 5.10: II.
116

1.7 I - Load mixes of Z and IM loads II - Load mixes of PQ and IM loads


1.7
1.5
Load flow results 1.5
1.3 1.3
Proposed formulation km = 0
1.1

Thv-mv
Thv-lv

1.1 km = 0.5 km = 0
0.9 km = 0.5 0.9
0.7 0.7 km = 1

0.5 0.5 Load flow results


km = 1 Proposed formulation
0.3 0.3
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
klv klv

Figure 5.10: Variation of Thv−mv with klv for ksc−mv = 4 (loads are supplied directly
at the MV busbar): I - for load mixes of Z and IM loads, II - for load mixes of PQ
and IM loads

1.2 I - ksc-mv = 12 1.4 II - ksc-mv = 4 km = 0


km = 0
1.3
1.1
1.2
1.0 1.1
km = 0.5
Thv-mv
Thv-mv

1.0 km = 0.5
0.9
0.9
0.8 0.8 km = 1
km = 1 0.7
Load flow results Load flow results
0.7 0.6
Proposed formulation Proposed formulation
0.6 0.5
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
klv klv

Figure 5.11: Variation of Thv−mv with klv (LV loads are supplied through MV lines):
I - for ksc−mv = 12, II - for ksc−mv = 4
117

5.3 Voltage Unbalance Influence Coefficients

5.3.1 Preliminary Investigations - Dependency of Influence Coefficients

on Load Types/Bases

Consider the radial MV-LV system shown in Fig. 5.1239 where the voltage at the

sending end busbar (labelled ‘send’) of the line is taken as unbalanced. The purpose

is to assess the voltage unbalance that propagates from the sending end busbar to

the receiving end busbar (labelled ‘rec’) of the MV sub-system, or the influence

coefficient ksend−rec between the busbars send and rec. For this, the loads Smv:rec−local

and Smv:rec−ds , and the MV line (t) are considered as balanced.

send

MV line
(t)

rec

Smv:rec-local (MVA)

Downstream
LV system

Smv:rec-ds (MVA)

Figure 5.12: Radial MV-LV system (reproduction of Fig. 3.2)

39
Reproduction of Fig. 3.2.
118

Based on the definition40 , the influence coefficient ksend−rec can be expressed as:


V−:Usend /rec
ksend−rec = (5.26)
V−:send

where,

V−:Usend /rec - negative sequence voltage at the receiving end busbar that propagates

from the sending end busbar, which can be expressed as41 :


V−:Usend /rec = V−:send − Z++:t I−:Usend /t (5.27)

V−:send - negative sequence voltage that exists at the sending end busbar

I−:Usend /t - negative sequence current in the line, which arises as a result of the voltage

unbalance existing at the sending end busbar

Equation (5.27) is a similar version of (5.5) considered above in relation to trans-

fer coefficients, where the impedance Z−−:tf −us (= Z++:tf −us ) of the US-S coupling

transformer is replaced here with Z++:t of the line. Thus, as in the case of transfer co-

V send /rec
efficients, the negative sequence voltage ratio −:U
V−:send
, i.e. the influence coefficient

ksend−rec , can be written for passive loads as:


V−:Usend /rec 1
V−:send = ksend−rec ≈ (5.28)
γ
Z++:t
1 + Z++:rec

where,

Z++:rec - downstream positive sequence impedance (or equivalent) seen at the MV

busbar rec
40
See Section 5.1.
41
Ignoring zero sequence unbalance, and replacing the negative sequence impedance Z−−:t of the
line with the positive sequence impedance Z++:t .
119



 1 for constant impedance loads



 0 for constant current loads

γ=
β for constant power loads, where β = −1 and − 2 for low (∼ 0.9) and







 high (∼ 1) lagging pf conditions respectively

Equation (5.28) can be rearranged as:

ksend−rec ≈ (1 − V Rt )γ (5.29)

where, V Rt - voltage regulation of the line42 . That is, the influence coefficient ksend−rec

is equal to unity, smaller than unity by the factor (1−V Rt ), and greater than unity by

the factor (1 − V Rt )β for constant current, constant impedance, and constant power

loads respectively. In other words, the term Z++:t I−:Usend /t causes the receiving end

negative sequence voltage V−:Usend /rec to deviate away from the sending end negative

sequence voltage |V−:send | by the factor (1 − V Rt )γ for passive loads. Note that an

identical behaviour was observed in Chapter 3 (Section 3.2) for the negative sequence

current or the term Z++:t I−:t/t which arises as a result of an asymmetrical line, where
t
the term Z++:t I−:t/t was seen to cause the negative sequence voltage |V−:g/mv:rec | to

deviate away from the principal term |Z−+:t I+:t | by the factor (1 − V Rt )γ for passive

loads. Employing this similarity, for three-phase induction motor loads supplied at

the LV level43 , the influence of the term Z++:t I−:send/t on ksend−rec can be written44 as:

1
ksend−rec ≈    (5.30)
V Rt 1
1+ 1−V Rt 1
+k 1
ks sc−lvagg

42
Voltage regulation is defined as the ratio between the positive sequence voltage drop across the
line (i.e. Z++:t I+:t ), and the sending end positive sequence voltage.
43
i.e. Smv:rec−local = 0.
44
Refer to (3.9).
120

As in Chapter 3 (Section 3.2), the factor (1 − V Rt )γ can be approximated to

unity (in other words I−:Usend /t ≈ 0), resulting in a unity ksend−rec for passive loads

in general. However, the influence coefficient ksend−rec for induction motor loads is

considerably smaller than unity45 , implying that the impact of the negative sequence

current I−:Usend /t in the presence of motor loads on ksend−rec cannot be ignored as in

the case of passive loads. Based on this, ksend−rec for a mix of passive (at MV and/or

LV) and motor (at LV) loads can be expressed46 in a generalised form as:

1
ksend−rec ≈    (5.31)
V Rt klv
1+ 1−V Rt 1
+k 1
ks km sc−lvagg

Fig. 5.13 illustrates the variation of ksend−rec with km established using (5.31) in

comparison to the results obtained using unbalanced load flow analysis for the test

system47 described in Appendix B in relation to three cases where:

• klv = 1 (i.e. Smv:rec−local = 0)

• klv = 0.5

• klv = 0 (i.e. Smv:rec−ds = 0, which implies that no motor loads are supplied by

the MV line)

The results presented in Fig. 5.13 correspond to a selected operating scenario where

the MV line supplies a total of 10MVA load at 0.9 lagging pf resulting in a V Rt ≈

8.5%. Fig. 5.13 confirms the above basis given by (5.31) which describes the be-

haviour of different load bases in relation to the propagation of voltage unbalance in

a sub-system, also demonstrating the dependency of this propagation on the motor

proportion.
45
e.g. 0.6 for ksc−lvagg = 20, ks = 6.7 and V Rt = 10%.
46
Refer to (3.14).
47
Where ksc−lvagg ≈ 19, ks = 6.7, passive load composition - equal proportions of constant
impedance and constant power elements, and the MV line is taken as ideally transposed.
121

1.1 klv = 0 (Smv:rec-local = 0, i.e. no motor loads are supplied)

1.0
klv=0.5
0.9
ksend-rec

0.8

0.7 klv = 1 (Smv:rec:ds = 0)

0.6 Load flow results


Eq. (5.31)
0.5
0.0 0.2 0.4 0.6 0.8 1.0
km

Figure 5.13: Variation of ksend−rec with km for the cases where klv = 1, klv = 0.5 and
klv = 0

5.3.2 Methodology for Evaluating Influence Coefficients

Consider the network shown in Fig. 5.1448 . For the purpose of assessing the voltage

unbalance that propagates from any busbar i to other busbars 1, 2, ..., n, i.e. influence

coefficients ki−1 , ki−2 , ..., ki−n , the voltage at the upstream system, all loads supplied

by the system, and all lines in the network49 are considered to be balanced.

Based on a similar approach used to establish (3.21) in Chapter 3 (Section 3.3), the

nodal negative sequence currents50 I−:Ui /x at the busbars x (x = 1, 2, ..., n and x 6= i)

can be written in terms of the nodal negative sequence admittances Y−−:xy 51 and

nodal negative sequence voltages52 V−:Ui /x at the busbars x, and the nodal negative

sequence voltage V−:i which exists at the busbars i as53 :


48
Reproduction of Fig. 3.5.
49
Including lines that exist at lower voltage levels, e.g. MV lines when assessing influence coeffi-
cients in an HV network.
50
Which arise as a result of the voltage unbalance that exists at the busbar i.
51
Y−−:xy = Y++:xy , where x, y = 1, 2, ..., n and x 6= i.
52
Which are caused by the voltage unbalance that exists at the busbar i.
53
Note that Y−+:xy = 0 for ideally transposed lines.
122

 
V−:Ui /1
   


−I−:Ui /1 Y Y++:12 . . . Y++:1i . . . Y++:1n  V
−:Ui /2

   ++:11  
 ..
   
 −I−:U /2   Y++:21 Y++:22 . . . Y++:2i . . . Y++:2n 
.

i

 = 
    
 .
.. .. .. .. .. .. ..  
. . . . . .   V−:i
    
   
..
    
−I−:Ui /n Y++:n1 Y++:n2 . . . Y++:ni . . . Y++:nn 
 . 

 
V−:Ui /n
(5.32)

Busbar 2

Upstream
system Busbar 1 Busbar x Busbar n
… …

Busbar 3

Sub-system S

Figure 5.14: Interconnected sub-system S (reproduction of Fig. 2.6)


123

Equation (5.32) can be decomposed as:

      
−I−:Ui /1 Y Y++:12 . . . Y++:1n V Y
   ++:11   −:Ui /1   ++:1i 
      
 −I−:U /2   Y++:21 Y++:22 . . . Y++:2n   V−:U /2   Y++:2i 
i i
= + V
      
−:i 
 ..   .. .. .. ..  ..  .. 
. . . . . . .
      
      
      
−I−:Ui /n Y++:n1 Y++:n2 . . . Y++:nn V−:Ui /n Y++:ni
(5.33)

Equation (5.33) can be written in a concise form as:

−[I−:Ui/x ] = [Yxz:++ ][V−:Ui/x ] + V−:i [Yxi:++ ] (5.34)

where, x, z = 1, 2, ..., n and x, z 6= i. As shown in Section 5.3.1 above, the influence of

the negative sequence currents [I−:Ui ] on the negative sequence voltages [V−:Ui ] should

be taken into account in the presence of considerable proportions of induction motor

loads. Thus, the current I−:Ui /x at any busbar x can be generally written when the

busbar supplies a mix of passive (at HV, MV and/or LV) and motor (at LV) loads as:

I−:Ui /x ≈ Y−−:x−im V−:Ui /x (5.35)

where, Y−−:x−im - downstream negative sequence admittance seen at the busbar x

taking only induction motors into account. As discussed in Chapters 3 (Section 3.3)

and 4 (Section 4.3) with regard to MV and HV systems respectively, this admittance

Y−−:x−im is inductive in nature, and dependant on the system and load characteris-

tics, system operating conditions and downstream load composition. As an example,

taking the network shown in Fig. 5.14 as an MV network with Fig. 5.15 representing

the system54 supplied by any busbar x, the admittance Y−−:x−im can be expressed as
54
Where the downstream system represents an aggregated LV system.
124

given by (3.24) which is reproduced here by (5.36) for completeness55 :


! √ !
klv:x 3 |I+:x |
Y−−:x−im ≈ −j 1 (5.36)
ksc−lvagg :x
+ ks:x1km:x Vn−mv

Substitution of (5.35) in (5.34) and rearrangement gives:


]−1
0
[ki−x ](n−1)×1 ≈ [Y++:xz [Y ] (5.37)

(n−1)×(n−1) ++:xi (n−1)×1

where,
0
Y++:xz ≈ Y++:xz + Y−−:x−im for x = z
0
Y++:xz = Y++:xz for x 6= z

Busbar x

Smv:x-local

Downstream
system supplied
by the busbar x

Smv:x-ds

Figure 5.15: System representation of any busbar x of the MV system shown in


Fig. 5.14 (reproduction of Fig. 3.6)

As discussed in Chapter 4 (Section 4.3) in relation to the emission arising as

a result of system inherent asymmetries, the presence of positive sequence voltage

controlled components such as PV generators and synchronous condensers in a system

force the negative sequence voltage at the connected busbars to be zero disregarding

the existence of sources of unbalance. Equation (5.37) which gives the influence
55
Refer to Chapter 3 for the definitions of the symbols.
125

coefficients ki−x between busbar i and other neighbouring busbars x does not consider

the presence of such components, and thus requires suitable adjustments such that

the impact of zero voltage unbalance (or of voltage controlled components) at given

busbars on these influence coefficients is accommodated. Similar to the approach

taken in Chapter 4, the matrix equation (5.37) can be modified to incorporate the

influence of a constraint V−:j = 0 at any busbar j on the influence coefficients [ki−x ]

by56 :

0
• Reducing the dimension of the matrix [Y++:xz ] down to (n − 2) × (n − 2) by

removing both the j th row and column.

• Reducing the dimension of the matrix [Y++:ix ] down to (n − 2) × n by removing

the j th row.

5.3.3 Verification of the Methodology Using a Three-bus MV Test

System

The proposed methodology is applied to the three-bus MV network (60Hz, 12.47kV,

three-wire) shown in Fig. 5.1657 for evaluating the voltage unbalance which propa-

gates from busbar 1 to busbars 2 and 3, i.e. influence coefficients k1−x for x = 2, 3.

Lengths of the lines which are taken as identical in construction58 and ideally trans-

posed are shown alongside the lines. The positive sequence admittance per km of the

lines is (1.0098−j2.0630)Skm. Busbar 3 supplies MV loads (2MVA at 0.9 lagging pf)

with equal compositions of constant impedance and constant power elements. That

is, km:x = 0 implying that Y−−:x−im ≈ 0 for x = 3. Busbar 2 supplies a mix (4MVA
56
Note that there is only (n − 2) number of influence coefficients are to be determined as the
influence coefficient ki−j is known to be zero.
57
Reproduction of Fig. 3.7.
58
See Appendix B for further details.
126

at 0.9 lagging pf) of passive loads59 and induction motors (ks = 6.760 ) at the LV

level, which accounts for 40% of the total load supplied by the system. Note that

Y−−:x−im 6= 0 when km:x > 0 for x = 2.

Fig. 5.17 illustrates the variations of k1−2 and k1−3 with km:2 established using the

proposed methodology61 in comparison to the results obtained using unbalanced load

flow analysis, demonstrating the accuracy of the proposed technique. Further, these

results reveal that motor loads help reducing the voltage unbalance that propagates

between neighboring busbars compared to the case where only passive loads exist.

5.3.4 Verification of the Methodology Using the IEEE 14-bus Test

System

The proposed methodology is further applied to the IEEE 14-bus test system shown

in Fig. 5.18 which consists of positive sequence voltage controlled busbars (busbars

1, 2, 3, 6 and 8), taking it as a 66kV, 60Hz and three-wire network supplying constant

power loads at the HV level. System and line62 data are given in Appendix K.

Fig. 5.19 illustrates the influence coefficients ki−x where i = 4, x = 1−14 and x 6=

463 established using the proposed methodology in comparison to the results obtained

using unbalanced load flow analysis, further validating the proposed technique.
59
Which are represented using a mix of constant impedance and constant power elements with
equal compositions.
60
See Appendix B for further details.
61
Application of the methodology to the test system is described in Appendix M.
62
Which are taken as identical in construction as described in Appendix H, and ideally transposed.
The positive sequence admittance per km of the lines = (0.2729 + j1.0244) × 102 pu (based on a
100M V A base).
63
i.e. the propagation of voltage unbalance from busbar 4 to other busbars.
127

204A
MV busbar 2
(0.98pu, -6.740) LV (460V)
4MVA
0.9 lagging pf
10km
Upstream HV 135A
(66kV) system MV busbar 1
(1.05pu, 00) (1.05pu, -4.070)
10km
59A

5km
152A

4MVA
0.9 lagging pf
2MVA
0.9 lagging pf
MV busbar 3
(1.01pu, -5.570)

HV-MV coupling transformer – 12MVA, winding resistance = 1%, leakage reactance = 10%,
secondary tap setting = 1.05pu

MV-LV coupling transformer – aggregated representation of fully loaded 1MVA transformers with
winding resistance = 1%, leakage reactance = 5% and secondary tap setting = 1.05pu

Figure 5.16: Three-bus MV test system considered for applying the proposed method-
ology (reproduction of Fig. 3.7)

1.1
k1-3
1.0

0.9
k1-2, k1-3

0.8

0.7 k1-2
Load flow results
0.6
Methodology
0.5
0.0 0.3 0.5 0.8 1.0
km:2

Figure 5.17: Variations of k1−2 and k1−3 with km:2 for the three-bus MV test system
128

Please see print copy for image

Figure 5.18: IEEE 14-bus test system (reproduction of Fig. 4.9)

0.7
Load flow results
0.6
Methodology
0.5

0.4
k4-x

0.3

0.2

0.1

0.0
1 2 3 5 6 7 8 9 10 11 12 13 14
Busbar x

Figure 5.19: Influence coefficients k4−x (x = 1 − 14, x 6= 4) for the IEEE 14-bus test
system
129

5.4 Chapter Summary

This chapter has addressed the propagation of voltage unbalance. This is a key aspect

in assessing emission limits to individual installations connected to EHV, HV, MV

and LV public power systems essentially based on the IEC/TR 61000-3-13 recommen-

dations. This propagation from higher voltage to lower voltage systems in terms of

transfer coefficients (Section 5.2), and from one busbar to other neighbouring busbars

of a sub-system in terms of influence coefficients (Section 5.2) has been addressed.

Theoretical bases which describe the behaviour exhibited by four basic load types

with regard to transfer coefficients and influence coefficients have been developed.

The following can be drawn as a summary of this work:

• Transfer coefficients, for passive loads, have to be scaled up by the factor


τ
1
1 + j ksc−s ∠θpf :s relative to the value of unity assumed in the IEC method,

where τ = 0, 1, 2 ∼ 3 for constant impedance, constant current and constant

power loads respectively. Noting that the factor ksc−s can be a value in the

range of 5 to 25 for various systems and the exponent τ varies in the range of 0

to 3 for different load types, a uniform behaviour or the behaviour of constant

impedance loads as assumed in the IEC approach cannot be used to represent

all load types with a high degree of accuracy.

• Transfer coefficients, for induction motor loads, have to be scaled down by the
1 01
factors „ « and 2 13 relative to unity in the cases of
1+ k ks
“ ”
41+ 1 ks
sc−lv ksc−mv
@
ks
A5
1+
ksc−lv
agg
the MV to LV and HV to MV propagation respectively. Noting that 5 < ks < 7,

the degree of this reduction is significant compared to the increment introduced

by passive loads.
130

• Considering a simple two-bus radial sub-system, the influence coefficients can

be approximated to unity for passive loads in general. However, these influence

coefficients can be considerably smaller than unity when the network supplies

a large proportion of induction motor loads. The conclusion of this observation

is that the negative sequence currents which arise as a result of the voltage

unbalance that exists at a particular busbar introduce a considerable impact

on influence coefficients in the presence of large proportions of motor loads,

although it is insignificant for passive loads.

Systematic methods for evaluating the MV to LV and HV to MV transfer coeffi-

cients, and influence coefficients for interconnected network environments have been

developed. These have been verified using unbalanced load flow analysis. In summary:

• It has been demonstrated that the proposed new formulation for assessing the

MV to LV transfer coefficient gives a more accurate estimation, particularly

for load bases which are dominated by passive elements, compared to the IEC

method.

• The MV to LV and HV to MV transfer coefficients can vary in the ranges

of 0.6 to 1.1 and 0.5 to 1.4 respectively depending on the system and load

characteristics and downstream load composition.

• Verification of the proposed method for estimating influence coefficients has

been undertaken employing a three-bus MV test system and also the IEEE

14-bus test system.


Chapter 6

A Revised Voltage Unbalance

Allocation Technique Based on the

IEC/TR 61000-3-13 Guidelines

6.1 Introduction

The IEC approach of managing continuous power quality disturbances (e.g. har-

monics, flicker and voltage unbalance) through the allocation of emission limits to

installations is based on a common philosophy. Thus, the problem of violating the

set planning levels, which has been identified with regard to harmonics and flicker1 ,

is an anticipated problem with the new IEC/TR 61000-3-13 voltage unbalance al-

location aproach as well. This chapter examines the IEC/TR 61000-3-13 procedure

employing a simple three-bus HV test system in Section 6.2. The principles of the

constraint bus voltage (CBV) method, that was discussed in Section 2.9 of Chap-

ter 2 as an alternative approach for harmonics and flicker allocation, are introduced

to voltage unbalance so that a robust allocation technique which closely aligns with
1
Refer to Section 2.9 of Chapter 2.

131
132

IEC/TR 61000-3-13 is developed in Section 6.3 of this chapter. Section 6.4 examines

this revised allocation technique using the above mentioned test system. A summary

of the chapter is given in Section 6.5.

6.2 Examination of the IEC/TR 61000-3-13 Approach

Consider the three-bus HV system (60Hz, 66kV, three-wire) shown in Fig. 6.1. The

system supplies constant power loads at the HV level. Considered operating scenario

and resulting positive sequence system conditions2 are also indicated in Fig. 6.1.

Lengths of the lines which are taken as identical in construction (including the phase

positioning) are shown alongside the lines. The relevant admittance data (per km)

of the lines are3 :

• positive sequence admittance = (0.6265 − j2.3517)Skm

• negative-positive sequence coupling admittance when untransposed =

(0.1040 + j0.1779)Skm

The examination procedure involves the following two steps:

• Calculation of the emission limits to individual installations using the IEC/TR

61000-3-13 prescribed formulae which have been reproduced in Section 2.8.3

of Chapter 2, together with the methodologies that have been proposed in

Chapters 3 - 5.

• Derivation of the busbar emission levels, which result in when all individual

installations are injecting at their allocated limits, using the general summa-

tion law.
2
Nodal voltages and line currents, which are obtained using load flow analysis.
3
Refer to Appendix H for further details.
133

HV busbar 2
(0.992pu, -10.840)

20MVA
0.95 lagging pf
50km
Upstream EHV 112A
(230kV) system HV busbar 1
(1.100pu, 00) (1.026pu, -7.980)
40km
64A

20km
151A

20MVA
0.95 lagging pf
10MVA
0.95 lagging pf
HV busbar 3
(1.007pu, -9.510)

EHV-HV coupling transformer – 60MVA, winding resistance = 1%,


leakage reactance = 20%, secondary tap setting = 1pu

Figure 6.1: Three-bus HV test system considered for examining the IEC/TR 61000-
3-13 approach
134

6.2.1 Calculation of Individual Emission Limits

The procedure of the calculation of the emission limits to the three aggregated loads

supplied by the test system is described in the following steps:

Global Emission Allowance

The voltage at the upstream system of the HV network under evaluation is taken as

balanced. That is, the upstream contribution to voltage unbalance in the considered

HV system is zero, resulting in a global emission allowance Ug/hv which is equal to the

network planning level. As indicated in IEC/TR 61000-3-13, a uniform HV planning

level of 1.4% is assumed, i.e. Ug/hv = 1.4%.

Apportioning of the Global Emission Allowance to Busbars

Referring to (2.10) and (2.11), the derivation of the busbar emission allowances Ug/hv:x

requires the initial evaluation of influence coefficients. A method for estimating influ-

ence coefficients was proposed in Section 5.3 of Chapter 5 as given by (5.37). Table 6.1

gives these influence coefficients for the considered test system, which are derived us-

ing the proposed method4 . Fig. 6.2 gives a comparison of these values with those

obtained using unbalanced load flow analysis. Arising from these values, an inter-

esting fact to note is that influence coefficients do not essentially hold the reciprocal

relationship ki−x = kx−i implying that the propagation from busbar i to busbar x is

not necessarily equal to that from busbar x to busbar i.

The total apparent power Shv:x supplied by any busbar x, the total available

apparent power Shv:x−total of the entire sub-system as seen at the busbar x which is
4
As the test system supplies constant power loads at all three busbars (i.e. km:x = 0 implying
0
that Y−−:x−im ≈ 0 for x = 1, 2, 3), the admittances Y++:xz ≈ Y++:xz not only for x 6= z but also for
x = z.
135

Table 6.1: Influence coefficients for the test system shown in Fig. 6.1
k1−2 k1−3 k2−1 k2−3 k3−1 k3−2
1 1 0.57 0.71 0.69 0.86

1.2
Load flow results
1.0 Methodology

0.8
Value

0.6

0.4

0.2

0.0
k1-2 k1-3 k2-1 k2-3 k3-1 k3-2

Influence coefficient

Figure 6.2: A comparison of the influence coefficients for the test system derived using
the proposed method: (5.37), and unbalanced load flow analysis

Table 6.2: Shv:x , Shv:x−total and Ug/hv:x for the test system shown in Fig. 6.1
Busbar (x) Shv:x (MVA) Shv:x−total (MVA) Ug/hv:x (%)
1 20 38.2 0.88
2 20 48.6 0.74
3 10 44.2 0.48
136

derived using (2.10), and the busbar emission allowance Ug/hv:x at the busbar x that is

calculated using (2.11) with a summation law exponent of 1.45 are given in Table 6.2

for each of the three busbars. Note that, as the contributions from the neighboring

busbars 1 and 3 to voltage unbalance at busbar 2 (k1−2 = 1 and k3−2 = 0.86) are seen

to be greater than that at busbar 1 (k2−1 = 0.57 and k3−1 = 0.69), the allocation

approach allows a lower level of emission for busbar 2 than that for busbar 1 although

both busbars 1 and 2 supply loads of equal MVA demand.

Individual Emission Limits

Two cases are considered for examining the IEC/TR 61000-3-13 approach:

• Case 1 - all HV lines are ideally transposed. That is, the contribution from

system inherent asymmetries to the global emission levels is zero, resulting in a

K 0 uex = 06 or Kuex = 17 for all busbars. This leads to voltage unbalance allo-

cation formulae which are identical to that of the harmonics/flicker allocation

[2, 3], where the total busbar allowance Ug/hv:x can be allocated to installations.

• Case 2 - the line between busbars 1 and 3 is ideally transposed, and the other

two lines are untransposed. That is, system inherent asymmetries make some

contribution to the global emission levels resulting in a K 0 uex > 0 or Kuex < 1

for some or all busbars, implying that only a fraction of the busbar emission

allowance Ug/hv:x can be allocated to installations. A method for evaluating the


lines
global emissions Ug/hv:x which arise as a result of line asymmetries was proposed

in Section 4.3 of Chapter 4 as given by (4.16). These emissions for the test
5
Which is the indicative value given in IEC/TR 61000-3-13.
6
K 0 uex is the factor which accounts for the emission arising as a result of system inherent asym-
metries.
7
Kuex is the factor which represents the fraction of the busbar emission allowance that can be
allocated to installations.
137

system which are derived using the proposed method8 , the K 0 uex factors which

are calculated using (2.14) with the above busbar allowances Ug/hv:x (Table 6.2),

and the Kuex factors (= 1 − K 0 uex ) are given in Table 6.3. Fig. 6.3 gives a

comparison of these K 0 uex values with those obtained using unbalanced load

flow analysis.

The busbar emission limits Ehv:x 9 or the emission limits to the three aggregated

loads, which are derived using (2.15) with the above busbar allowances Ug/hv:x (Table

6.2) and Kuex factors (unity for Case 1, and as given in Table 6.3 for Case 2), are

given in Table 6.4 for Cases 1 and 2. Note that, in Case 2, only 70% of the busbar 2

allowance can be allocated to the connected load as the emission arising as a result

of line asymmetries at the busbar is considerably high.

lines
Table 6.3: Ug/hv:x , K 0 uex and Kuex for Case 2 of the test system shown in Fig. 6.1
lines
Busbar (x) Ug/hv:x (%) K 0 uex Kuex
1 0 0 1
2 0.39 0.41 0.59
3 0.05 0.04 0.96

Table 6.4: Ehv:x according to IEC/TR 61000-3-13 for the test system shown in Fig. 6.1
Busbar (x) Ehv:x for Case 1 (%) Ehv:x for Case 2 (%)
1 0.88 0.88
2 0.74 0.51
3 0.48 0.47

8
As the test system supplies constant power loads at all three busbars (i.e. km:x = 0 implying
0 0
that Y−−:x−im ≈ 0 for x = 1, 2, 3), the admittances Y++:xy ≈ Y++:xy and Y−+:xy ≈ Y−+:xy not only
for x 6= y but also for x = y.
9
i.e. the combined emission limit of a load of which the agreed apparent power is equal to the
total apparent power Shv:x supplied by the busbar.
138
0.5
Load flow results
Methodology
0.4

0.3

K'uex
0.2

0.1

0.0
1 2 3
Busbar (x)

Figure 6.3: A comparison of the K 0 uex factors for the test system derived using the
proposed method: (4.16), and unbalanced load flow analysis

6.2.2 Resulting Busbar Emission Levels and Examination Remarks

When an installation supplied by a sub-system is injecting at its allocated limit, the

installation introduces an emission level at the connected busbar which is equal to

the imposed limit while making some contributions to voltage unbalance at other

neighboring busbars as well. According to IEC/TR 61000-3-13, this emission level

introduced at a neighboring busbar can be quantified10 as a product of the emission

limit and the influence coefficient between the two busbars. Thus, when all instal-

lations supplied by the system are injecting at their individual limits, the resulting
reult
global emission level Ug/s:x at any busbar x of the sub-system can be expressed using

the general summation law, also taking the global emission which arises as a result

of system inherent asymmetries into account, as:

reult
(k1−x Es:1 )α + (k2−x Es:2 )α + ... + (Es:x )α + ... + (kn−x Es:n )α

Ug/s:x =
lines α 1/α

+ (Ug/s:x ) (6.1)
10
Refer to the definition of the influence coefficient given in Section 2.8.3 of Chapter 2.
139

The busbar emission levels for the test system, which result in when the three

loads are injecting at the above allocated limits (Table 6.4), are derived using (6.1)

for the two cases and given in Table 6.5.

reult
Table 6.5: Ug/hv:x arising as a result of the IEC/TR 61000-3-13 allocation procedure
for the test system shown in Fig. 6.1
reult reult
Busbar (x) Ug/hv:x for Case 1 (%) Ug/hv:x for Case 2 (%)
1 1.24 1.15
2 1.52 1.51
3 1.40 1.30

Note that although the resulting emission levels at busbars 1 and 3 are below the set

planning level of 1.4%, this at busbar 2 exceeds the planning level by approximately

8% in each of the cases. This indicates, as anticipated, that the IEC allocation policy

is unlikely to be robust enough to comply with one of the key allocation objectives not

only in the cases of harmonics and flicker but also with regard to voltage unbalance.

6.3 A Revised Voltage Unbalance Allocation Technique Based

on the CBV Allocation Principles

The CBV allocation method which has been suggested for harmonics and flicker11

cannot be applied in its present form to voltage unbalance, as the case of voltage

unbalance involves an additional aspect which is the emission arising as a result of

system inherent asymmetries. Due to the presence of this additional emission, the

total busbar emission allowance Ug/s:x cannot be allocated to installations. That is,

in the case of voltage unbalance, the busbar emission limit Es:x 12 is usually smaller
11
Refer to Section 2.9.
12
Es:x of any busbar x is the emission limit of the load of power Ss:x .
140

than the busbar emission allowance Ug/s:x 13 , whereas Es:x = Ug/s:x for the cases of

harmonics and flicker. Thus, appropriate revisions addressing this issue are required

in extending this CBV method to voltage unbalance allocation.

Closely following the IEC/TR 61000-3-13 approach, in which the busbar emis-

sion allowance is derived14 by apportioning the global emission allowance Ug/s in



proportion to the term α Ss:x , the principle given by (2.22) for determining the bus-

bar emission limit under the CBV harmonics/flicker allocation policy is extended to

voltage unbalance for deriving the busbar allowance Ug/s:x as:

p
α
Ug/s:x = ka Ss:x (6.2)

The allocation constant ka is yet to be determined. As addressed in IEC/TR 61000-

3-1315 , the emission which arises as a result of system inherent asymmetries can be

accounted for by using the factor Kuex (= 1 − K 0 uex )16 in determining the busbar

emission limit Es:x as:

p
α
p
α
Es:x = Kuex Ug/s:x = ka Kuex Ss:x (6.3)

The factor Kuex can be expressed17 for the new allocation policy as a function of the

allocation constant as:

lines
!α lines α
Ug/s:x (Ug/s:x )
Kuex = 1 − =1− (6.4)
Ug/s:x kaα Ss:x
13
Ug/s:x for any busbar x accounts for the emissions which arise at the busbar as a result of both
the load of power Ss:x and line asymmetries.
14
See (2.11).
15
See (2.15).
16
The factors Kuex and K 0 uex√are defined by (2.13) and (2.14) respectively.
17
i.e. substituting Ug/s:x = ka α Ss:x .
141

Substitution of (6.4) in (6.3) gives:

q
lines α
Es:x == α
kaα Ss:x − (Ug/s:x ) (6.5)

As suggested in relation to harmonics and flicker, ensuring voltage unbalance levels

at network busbars do not exceed the set planning level, the allocation constant ka

can be chosen to be the largest value such that:

result
Ug/s:x ≤ Ug/s for every busbar x (6.6)

result
where, the resulting emission level Ug/s:x at any busbar x when all installations

are injecting at their individual limits can be written as given by (6.1) which can

rearranged for the new allocation policy given by (6.5) as:


result
kaα k1−x
α α α

Ug/s:x = S1 + k2−x S2 + ... + Sx + ... + kn−x Sn −
1/α
lines α lines α lines α
(Ug/s:1 ) + (Ug/s:2 ) + ... + (Ug/s:n ) (6.7)

Then, ka can be established in order to satisfy (6.6) as:

vu n

lines α
X 
α
 u (Ug/s ) + Ug/s:x
u

u 
α i=1,i6=x
ka = min  u (6.8)
u 
n 
u X
α
 
 t Ss:x + ki−x Ss:i 
i=1,i6=x

In summary, under the suggested alternative voltage unbalance allocation policy,

the emission limit to any customer installation j of which the MVA demand is Ss:x−j

to be connected at any busbar x can be derived using:


142

p
α
Es:x−j = ka Kuex Ss:x−j (6.9)

where the allocation constant ka is determined using (6.8), and the Kuex factor is

derived from (6.4).

6.4 Examination of the Revised Voltage Unbalance Alloca-

tion Technique

The same three-bus HV test system shown in Fig. 6.1 is employed for examining the

alternative voltage unbalance allocation approach suggested in Section 6.3 above.

6.4.1 Calculation of Individual Emission Limits

The calculation procedure, under the revised allocation policy, of the emission limits

to the three aggregated loads supplied by the test system is described in the follow-

ing steps:

Allocation Constant ka

Table 6.6 gives the values18 , corresponding to each of the three busbars, of the right

hand side (RHS) of (6.8) in relation to the two cases listed above. Choosing the

smallest value of the RHS of (6.8), which corresponds to busbar 2 in each of the

cases, the allocation constant ka for the two cases can be identified as given in Table

6.7. That is, the allocation constant is chosen such that the resulting emission level

at busbar 219 is constrained at the allowed global emission level of 1.4%.


18
α = 1.4, Ug/hv = 1.4%, apparent power values Shv:x as given in Table 6.2, and influence
lines
coefficients as given in Table 6.1, and emissions Ug/hv:x = 0 for Case 1 and as given in Table 6.3 for
Case 2 are used in the calculation.
19
Note that this is the busbar at which the resulting emission level was observed to be above the
set planning level under the IEC/TR 61000-3-13 allocation procedure.
143

Table 6.6: Values of the RHS of (6.8) in relation to the test system shown in Fig. 6.1
Busbar (x) RHS of (6.8) RHS of (6.8)
for Case 1 for Case 2
1 0.110 0.124
2 0.088 0.089
3 0.096 0.108

Table 6.7: ka for the test system shown in Fig. 6.1


Case 1 Case 2
ka 0.088 0.089

Individual Emission Limits

The Kuex factors calculated20 using (6.4), and the busbar emission limits Ehv:x or

the emission limits to the three aggregated loads derived21 from (6.9) for the two

different cases are given in Table 6.8. Figs. 6.4: I - II provides a comparison of the

individual emission limits derived according to IEC/TR 61000-3-13 and the revised

allocation method for Cases 1 and 2 respectively. This illustrates that the allocated

limits under IEC/TR 61000-3-13 are greater, specially at busbar 1, than that under

the revised method.

Table 6.8: Kuex and Ehv:x according to the revised allocation method for the test
system shown in Fig. 6.1
Busbar Kuex for Kuex for Ehv:x for Ehv:x for
(x) Case 1 Case 2 Case 1 (%) Case 2 (%)
1 1 1 0.75 0.76
2 1 0.60 0.75 0.53
3 1 0.96 0.46 0.45

20
α = 1.4, allocation constant ka as given in Table 6.7, apparent power values Shv:x as given in
lines
Table 6.2, and emissions Ug/hv:x = 0 for Case 1 and as given in Table 6.3 for Case 2 are used in the
calculation.
21
α = 1.4, allocation constant ka as given in Table 6.7, apparent power values Shv:x as given in
Table 6.2, and the Kuex factors as given in column 1 of this table are used in the calculation.
144

1.0 I - Case 1 1.0 II - Case 2

0.8 0.8 Revised method


IEC/TR 61000-3-13

Ehv:x (%)
Ehv:x (%)

0.6 0.6

0.4 0.4

0.2 0.2

0.0 0.0
1 2 3 1 2 3
Busbar (x) Busbar (x)

Figure 6.4: Comparison of the busbar emission limits Ehv:x derived according to
IEC/TR 61000-3-13 and the revised method for the test system: I - for Case 1,
II - for Case 2

6.4.2 Resulting Busbar Emission Levels and Examination Remarks

The busbar emission levels which result in when the three loads are injecting at the

above allocated limits (Table 6.8) are derived22 using (6.1) for the two cases, and given

in Table 6.9. Figs. 6.5: I - II provides a comparison of the resulting busbar emission

levels under IEC/TR 61000-3-13 and the revised allocation procedure for Cases 1

and 2 respectively. This illustrates that the resulting busbar emission levels under

IEC/TR 61000-3-13 are greater, at all three busbars, than that under the revised

method.

reult
Table 6.9: Ug/hv:x arising as a result of the revised allocation procedure for the test
system shown in Fig. 6.1
reult reult
Busbar (x) Ug/hv:x for Case 1 (%) Ug/hv:x for Case 2 (%)
1 1.12 1.04
2 1.40 1.40
3 1.28 1.18

22 lines
α = 1.4, influence coefficients as given in Table 6.1, and emissions Ug/hv:x = 0 for Case 1 and
as given in Table 6.3 for Case 2 are used in the calculation.
145

2.0 I - Case 1 2.0 II - Case 2 Revised method


Planning level = 1.4% IEC/TR 61000-3-13
1.6 1.6
/ hv:x (%)

/ hv:x (%)
1.2 1.2
U glines

U glines
0.8 0.8

0.4 0.4

0.0
0.0
1 2 3
1 2 3
Busbar (x) Busbar (x)

reult
Figure 6.5: Comparison of the resulting emission levels Ug/hv:x derived according to
IEC/TR 61000-3-13 and the revised method for the test system: I - for Case 1,
II - for Case 2

As expected, the revised allocation technique restricts the resulting emission level

at the constrained busbar (e.g. busbar 2 of the test system) to the allowed global

emission level (= 1.4% for the test case) while maintaining the emission levels at other

busbars (e.g. busbars 1 and 3 of the test system) below the set limit. That is, this

proposed alternative technique allows a robust allocation in the sense that it satisfies

all four key allocation objectives23 :

• The allocation constant ka determined using (6.8) ensures that the resulting

emission levels at all network busbars are maintained at or below the set plan-

ning level.

• The revised allocation policy given by (6.9) ensures that customer installations

of equal MVA demand (whether connected at the same busbar or a different

busbars) receive identical emission limits, and also that larger customer instal-

lations (in MVA demand terms) to receive larger emission levels than smaller

installations.
23
See Section 2.9.
146

• In determining the allocation constant ka using (6.8), the resulting emission level

at a busbar is constrained at the allowed global emission level. This implies that

the network absorption capacity is fully utilised.

6.5 Chapter Summary

This chapter has examined the recently introduced IEC/TR 61000-3-13 voltage unbal-

ance allocation approach. It was seen, as in the case of the counterpart IEC harmonics

and flicker allocation methods, the prescribed voltage unbalance allocation procedure

also leads to planning levels are being exceeded even when no customer exceeds the

allocated limit.

Closely aligning with the IEC/TR 61000-3-13 guidelines, a revised voltage un-

balance allocation policy based on the CBV harmonics/flicker allocation principles,

whereby the emission levels at network busbars are explicitly forced to be at or below

the set planning levels when all loads inject their limits derived under the new ap-

proach, has been proposed. This revised allocation technique was seen to restrict the

resulting emission level at the constrained busbar to the allowed global emission level

while maintaining the emission levels at other busbars below the set limit. Further,

this proposed alternative technique allows a robust allocation in the sense that it

satisfies all four key allocation objectives.


Chapter 7

Analysis of the Problem of Voltage

Unbalance in Interconnected

Power Systems

7.1 Introduction

The presence of voltage unbalance in electricity transmission and distribution net-

works has continued to be a power quality issue of concern primarily due to difficul-

ties found by some network service providers in maintaining acceptable levels. As

an example, consider the 66kV sub-transmission interconnected system (this will be

referred to as ‘study system’ hereinafter) shown in Fig. 7.1, which is owned and

operated by an Australian (Victoria) utility. This is connected to the EHV trans-

mission system at S1 (bulk supply point: BSP) where the level of voltage unbalance

has been measured to be negligible. The system is divided into three sub-parts: up-

stream (US), central part (CP) and downstream (DS) for convenience in presenting.

Some of the sub-transmission lines are longer than 50km, and are not systematically

147
148

ZS1 : bulk supply point

A
C B

D F
ZS2
ZS3

E
ZS4

I G

N
ZS6

PV generator :
ZS5 operates
Local only in limited time periods
generator:
J K operates only in limited time periods

Local generator: ZS7


PV generator :
operates
operatescontinuously
continuously

L
Voltage regulators

ZS8 Loads

Capacitor bank s

ZS9

Figure 7.1: 66kV sub-transmission interconnected system under study


149
2.0

1.6

VUF (%) 1.2

0.8

0.4

0.0
S2 S3 S4 S5 S6 S7 S8 S9
Busbar

Figure 7.2: Measured nodal VUF values for the study system

transposed1 . Despite the applicable voltage unbalance limit of 1% according to the

Victorian electricity distribution code2 [73], the measured voltage unbalance levels3

during the peak demand periods at S6, S7 and S8 have been seen to exceed the limit

in addition to significant levels (0.6% - 0.8%) at the upstream busbars S2 and S4.

These measured values corresponding to a selected time stamp4 that lies within the

system peak are illustrated in Fig. 7.2. Although effort has been put to address

this issue by balancing loads at some of the busbars, no significant improvements

have been noted indicating that asymmetries associated with the sub-transmission

lines can play a vital role in leading up to these excessive voltage unbalance levels.

However, due to insufficient knowledge at present in systematically evaluating such a

problem, especially in meshed network environments, the system operator/owner has

found challenges in better managing the network.


1
Line transposition is not a common practice at this voltage level due to economic reasons.
2
See Section 2.7.
3
These measurements, provided by the system operator/owner, are not based on a well defined
measurement and evaluation procedure such as described in the standard IEC 61000-4-30. Therefore,
the accuracy of these VUF values are uncertain, however they can be considered to provide a general
sense of the problem. Measurements are not available for the busbars S3, S5 and S9.
4
Operating conditions at this time stamp are given in Appendix N.
150

Objectives of the work presented in this chapter are:

• To carry out deterministic studies5 which are required to develop an insight into

the problem experienced by the study system.

• To develop systematic approaches which would facilitate the evaluation of such

a problem including the identification of major contributors to voltage unbal-

ance levels, and transposition options that can effectively reduce the emission

arising as a result of line asymmetries. These approaches would provide addi-

tional assistance in the application of the IEC/TR 61000-3-13 voltage unbalance

allocation methodologies.

Sections 7.2 and 7.3 present deterministic studies and outcomes, together with

the verification of theoretical findings, in relation to line and load asymmetries re-

spectively. The overall system behaviour or the combined influence of line and load

asymmetries is addressed in Section 7.4. The chapter is summarised in Section 7.5.

7.2 Voltage Unbalance Behaviour of Line Asymmetries

7.2.1 Impact of the Line Asymmetries of the Study System on the

Voltage Unbalance Problem

The impact of the interaction of all asymmetrical lines of the study system on the

problem of voltage unbalance is established in terms of the VUF values at the various

busbars using unbalanced load flow analysis. This is accomplished by synthesising the

system operation6 , however under balanced loading7 conditions. Fig. 7.3 illustrates
5
These studies are supported by the developed unbalanced load flow program which is described
in Appendix O.
6
As revealed by measurements, voltage at the BSP is taken as balanced. This eliminates the
contribution made by the upstream EHV system to voltage unbalance in the study system.
7
Constant power loads are assumed.
151

these results in comparison to the measured voltage unbalance levels8 (Fig. 7.2)

with respect to the selected time stamp. These results demonstrate that the line

asymmetries themselves introduce excessive voltage unbalance levels at the busbars

located in the central part (S6 and S7 - 1%) and in the downstream part (S8 and S9 -

1.4%) of the system, in addition to the considerable levels at the upstream busbars (S2

and S4 - 0.5%). This indicates the importance of proper line transposition practices

even at sub-transmission voltage levels, which is currently seen not to be a common

practice. However, the transposition of each single line is not economically viable

and/or practically feasible, and thus further knowledge is required in order to assist

a careful selection of lines for the transposition.

2.0
Line asymmetries
Measured values
1.6
VUF (%)

1.2

0.8

0.4

0.0
S2 S3 S4 S5 S6 S7 S8 S9
Busbar

Figure 7.3: Nodal VUF values (load flow results) which arise as a result of the line
asymmetries, in comparison to the measured values

8
Note that these measured levels arise as a result of the interaction of both the line and load
asymmetries which exist in the actual system.
152

7.2.2 Voltage Unbalance Behaviour of the Individual Lines of the

Study System - as Standalone Lines

When an asymmetrical line supplying balanced passive loads operates as a standalone

line9 which is energised by balanced supply voltages, as evident from Chapters 3 - 410 ,
t
the negative sequence voltage V−:rec at the receiving end busbar of the line (t) can be

written for operating scenarios most likely to occur in practice as11 :

t
V−:rec ≈ −Z−+:t I+:t (7.1)

where,

Z−+:t - negative-positive sequence coupling impedance of the line

I+:t - positive sequence current in the line

t
That is, the magnitude of V−:rec is given by the term |Z−+:t I+:t |. Thus, for a given

line, as illustrated in Fig. 7.4 which shows the results obtained using unbalanced

load flow analysis for each of the lines12 of the study system, the variation of |V−:rec
t
|

with |I+:t | is approximately linear with a gradient that is being equal to |Z−+:t |. In
t
addition, the phase angle θV−:rec
t of the negative sequence voltage V−:rec : θV−:rec
t ≈

1800 + θZ−+:t + θI+:t 13 , where θZ−+:t , θI+:t are phase angles of the impedance Z−+:t

and the current I+:t respectively. Fig. 7.5 illustrates the variation of θV−:rec
t with |I+:t |

obtained using unbalanced load flow analysis for each of the lines, justifying this re-
9
i.e. without introducing any effect of interactions which exist in interconnected network envi-
ronments, and of other sources of unbalance.
10
The impact of the negative sequence current I−:t/t in the line arising as a result of the line
itself (or of the term Z++:t I−:t/t where Z++:t is the positive sequence impedance of the line) on the
t
negative sequence voltage V−:rec is negligible when the line supplies passive loads.
11
The study system is three-wired. Thus, the impact of zero sequence variables on the negative
t
sequence voltage V−:rec can be ignored.
12
Descriptive data (lengths and impedances) of the lines is given in Appendix N (Table N.7).
13
Under high load power factor conditions, this can be further simplified as: θV−:rec
t ≈ 1800 +θZ−+:t .
153

lationship of the phase angle θV−:rec


t . For a given line, the angle θV−:rec
t remains almost

constant for various |I+:t | values, where this distinct angle is determined by the angle

θZ−+:t as the angle θI+:t is governed by the load power factor14 .

600 A
B
C
500
D
E
400 F
G
|V -:rec| (V)

H
300
I
t

J
200 K
L
M
100
N

0
0 50 100 150 200 250 300
|I+:t| (A)

t
Figure 7.4: Variation of |V−:rec | with |I+:t | for the individual lines

In summary, the term −Z−+:t I+:t (which is a vector) governs the voltage unbal-

ance behaviour exhibited by a standalone asymmetrical line. The impedance element

|Z−+:t | which is the principal intrinsic parameter behind the voltage unbalance emis-

sion can be used as a measure for assessing the degree of asymmetry associated with

a line. Employing this, a rank for the lines of the study system can be assigned as

given in Table 7.115 .


14
The behaviour associated with line L (in both Figs. 7.4 and 7.5) is slightly out of ordinary at
higher |I+:t | values. This arises as a result of its relatively very large positive sequence impedance
(Table N.7) which leads to a considerable level of impact from the neglected term Z++:t I−:t/t on
t
the negative sequence voltage V−:rec . However, these |I+:t | values at which the line exhibits an out
of ordinary behaviour are unlikely to arise in practice (e.g. voltage regulation of the line > 20% for
|I+:t | > 200A).
15
The purpose of this ranking is to give an assessment of the degree of asymmetry associated with
the lines. The overall influence of a line on voltage unbalance levels, which is determined by other
154

200
A
150 B
C
100
D
E
(deg .)

50
F
0
G
−:rec

-50 H
θV t

I
-100 J
-150
K
L
-200 M
25 75 125 175 225 275
N
|I+:t| (A)

Figure 7.5: Variation of θV−:rec


t with |I+:t | for the individual lines

Table 7.1: Ranking of the sub-transmission lines based on the associated degree of
asymmetry
Line/s |Z−+:t | (Ω) Rank
M ∼2.0 Extraordinarily High
F, I ∼1.3 Very high
A, D ∼0.65 High
B, N ∼0.5 Moderate
J, L, E, C ∼0.3 Low
G, K, H ∼0.1 Very low
155

7.2.3 Voltage Unbalance Behaviour of the Individual Lines of the

Study System - as Elements in the Interconnected Network

The voltage unbalance behaviour exhibited by each of the lines, when operating in the

interconnected network, is established employing the concept of voltage unbalance as a

vector16 (referred to as ‘voltage unbalance emission vector’ [1]) using unbalanced load

flow analysis. This is accomplished by synthesising the system operation, however

parameterising such that:

• The line of which the behaviour is to be observed (referred to as ‘line under

observation’, and labelled as t) to represent its actual construction.

• All other lines are ideally transposed.

• All loads are balanced.

This leaves the line under observation as the only source of unbalance that exists in

the system.

A line under observation (t) introduces voltage unbalance at various busbars where:

t
• The negative sequence voltage V−:rec−t at its receiving end busbar has to satisfy

the following relationship:

t t
V−:rec−t = V−:send−t − Z−+:t I+:t − Z++:t I−:t/t (7.2)

t
where, V−:send−t is the negative sequence voltage introduced by the line under

observation at its sending end busbar.


additional parameters such as the line loading level (|I+:t |), will be discussed in Section 7.2.3 when
the line operates in the interconnected network environment.
16
See Section 2.2.
156
t
• The negative sequence voltage V−:rec−tany
at the receiving end busbar of any

other ideally transposed line (tany ) has to satisfy the following relationship17 :

t t
V−:rec−tany
= V−:send−tany
− Z++:tany I−:t/tany (7.3)

where,
t
V−:send−tany
- negative sequence voltage introduced by the line under observation

at the sending end busbar of any other ideally transposed line tany

Z++:tany - positive sequence impedance of any other ideally transposed line tany

I−:t/tany - negative sequence current in any other ideally transposed line tany ,

which arises as a result of the line under observation t

10
As evident from Chapters 3 - 4 (see footnote above), the influence of the terms

Z++:t I−:t/t in (7.2) and Z++:tany I−:t/tany in (7.3) on the negative sequence voltages

is insignificant compared to that of the term Z−+:t I+:t . That is, as in the case of a

standalone line, the voltage unbalance behaviour exhibited by an asymmetrical line at

various busbars of an interconnected network is also governed by the term −Z−+:t I+:t ,

however dependant on the location of the line in the network18 .

Fig. 7.6 illustrates the VUF values at the various busbars (S2 - S9 of Fig. 7.1)

corresponding to the selected time stamp obtained by applying each of the lines

(one at a time) as a line under observation. Table 7.2 gives further details of the

lines including commentaries on the values of |Z−+:t | (columns 2), |I+:t | (columns 3)

and |Z−+:t I+:t | (columns 4), and also on the location in the network (column 5) by

assigning into the three sub-parts: US, CP and DS (Fig. 7.1). Commentary on the

location also indicates whether or not a line is in the direct path connecting the BSP
17
Note that the negative-positive coupling impedance Z−+:tany of any other line is zero, as it is
taken as ideally transposed.
18
Further explanation on this influence of the location of a line is given in Appendix N.
157
0.8 S2
S3
0.7
S4
0.6 S5
0.5
S6
VUF (%)

S7
0.4 S8
0.3 S9

0.2

0.1

0.0
A B C D E F G H I J K L M N
Line under observation

Figure 7.6: Nodal VUF values arising as a result of the individual lines

and CP and/or DS of the network. Depending on the emission levels which arise as a

result of the individual lines at the various busbars, a rank19 is assigned to the lines as

given in column 6. Referring to the entries in Table 7.2, the following can be derived:

• Line M which carries a relatively low level of positive sequence current is not

seen to be a major source of emission, although it is the most asymmetrical

line of the network. Line J which carries the highest level of positive sequence

current is seen to be a major source of emission, although it has a relatively

low level of asymmetry. This indicates, as implied by (7.2) and (7.3), that the

product |Z−+:t I+:t | is the term of concern when assessing the level of emission

introduced by an asymmetrical line as an element in an interconnected network.

• In contrary to the above, lines B and N introduce only a low level of emission

although their |Z−+:t I+:t | is equal to that of line J, a behaviour which can be

attributed to the location of these lines in the network20 . Lines B and N are not
19
Taking the highest emission level, i.e. the tallest bar of the group of bars of a line under
observation (Fig. 7.6), as the reference.
20
See Appendix N.
158

in the path connecting the BSP, CP and/or DS, and thus their contributions to

voltage unbalance levels at the remaining busbars is reduced compared to that

of line J.

200
A
B
Phase angle (deg.)

100 C
D
F
0
I
J
-100 L
M
N
-200
S2 S4 S6 S7 S8 S9
Busbar

Figure 7.7: Phase angles of the nodal negative sequence voltages introduced by the
individual lines

Fig. 7.7 illustrates the phase angles of the negative sequence voltages at S2, S4 and

S6 - S921 caused by lines A - D, F, I, J and L - N22 individually. This demonstrates

that each of the above individual lines leads to a unique and nearly constant phase

angle across all busbars. Comparison of Figs. 7.5 (behaviour as standalone lines)

and 7.7 (behaviour as elements in the interconnected network) clearly demonstrates

the similarity of the voltage unbalance behaviour of a line as a standalone line and

as an element in an interconnected network. This further justifies the governance of

the behaviour exhibited by a line as an element in an interconnected network by the

term −Z−+:t I+:t .


21
Which are affected significantly by the asymmetrical network as noted in Fig. 7.3.
22
Which introduce considerable levels of emission as noted in Table 7.2.
159

Table 7.2: Parameters, operating features and emission levels of the individual lines

Line/s |Z−+:t | |I+:t | |Z−+:t I+:t | Location Emission


level
M Extra large Very low Very low DS: an element Low
(∼2Ω) (∼15A) (.30V) in the BSP-CP-DS
connection
F Very large High Very high US: tie line Very high
(∼1.3Ω) (∼150A) (∼230V)
I Very large High Very high US: an element Very high
in the BSP-CP-DS
connection
A, D Large Very High High US: elements in High
(∼0.65Ω) (∼200A) (∼150V) the BSP-CP-DS
connection
B, N Moderate High Moderate US: not elements Low
(∼0.5Ω) (∼65V) in the BSP-CP-DS
connection
J Small Extra High Moderate CP: an element High
(∼0.3Ω) (∼250A) in the BSP-CP-DS
connection
L Small Low Very low CP: an element Low
(∼75A) in the BSP-CP-DS
connection
E Small Low Very low US: tie line Very low
C Small Very high Low US: not an element Low
(∼40V) in the BSP-CP-DS
connection
G, H Very Small High Very low US: elements in Very Low
(∼0.1Ω) the BSP-CP-DS
connection
K Very Small High Very low CP: element Very Low
in the BSP-DS
connection
160

7.2.4 General Outcomes - Representation of the Voltage Unbalance

Behaviour of an Asymmetrical Line as an Element in an Inter-

connected Network

An asymmetrical line of an interconnected network exhibits a voltage unbalance be-

haviour which is vectorial in nature. This behaviour, in a global sense, can be ascer-

tained by a single vector (referred to as ‘global emission vector’) of which:

• The magnitude, as summarised in Table 7.2 - column 6 for the lines of the study

system, can be approximately assessed by referring to the product |Z−+:t I+:t | of

the line and its location23 in the network.

• The phase angle, as illustrated in Fig. 7.7 for the lines of the study system, can

be approximately derived using the term −Z−+:t I+:t 24 .

The global emission vectors of the individual lines25 A - D, F, I, J and L - N of

the study system are illustrated in Fig. 7.8.

7.2.5 General Outcomes - Representation of the Interaction of All

Asymmetrical Lines

The behaviour of negative sequence variables is known to be linear [1]. That is, a

resultant negative sequence voltage at a busbar, which arises as a result of the inter-

action of various sources of unbalance, is equal to the vector summation of negative

sequence voltage components caused by individual sources at the considered busbar26 .


23
The work presented in this thesis does not cover an absolute measure of the impact of the
location of the line on the magnitude of the vector rather than suggesting a relative assessment.
This is possibly a subject of interest for further research.
24
Further simplified term −Z−+:t can be used under high load power factor conditions, which is
generally the case in higher voltage systems.
25
Which introduce considerable emission levels.
26
This is further demonstrated in Appendix N employing results obtained using unbalanced load
flow analysis in relation to the study system.
161

I C
N
M
D
B L
J
A
F

Figure 7.8: Global emission vectors of the individual lines (drawn approximately to
a scale)

Thus, the integration of global emission vectors of individual lines, as illustrated in

Fig. 7.8 for the lines of study system, establishes a basis that provides a comprehensive

understanding of the manner in which various asymmetrical lines interact with each

other to form the resultant influence. This basis can be used to derive the following:

• The resultant influence of the interaction of all lines in terms of a single vector

which can be established by the summation of individual global emission vectors.

This, for the study system, is illustrated by the vector Rlines in Fig. 7.9.

• Major contributing lines to the resultant influence.

– Assessments on the study system using the proposed technique:

Referring to Fig. 7.9, lines A and F can be identified as dominant contrib-

utors as the respective global emission vectors lie in close proximity to the

resultant vector. Although the global emission vector of line I is displaced

slightly away from the resultant vector, it, being the line which introduces

the highest level of emission on its own, can also make a significant contri-
162

I C
N
M
D
B L
J
A
F

Rlines

Figure 7.9: Resultant influence of the interaction of all asymmetrical lines (drawn
approximately to a scale)
163

bution. The phase deviation close to 900 of the vector of line J with respect

to the resultant vector can make it a less of a contributor. The positioning

of the vector of line D with respect to the resultant vector suggests that it

can make a negative contribution, or in other words it assists in counter

balancing some of the emissions caused by the other lines.

– Validation of the above assessments using unbalanced load flow analysis:

Fig. 7.10 illustrates the contributions, quantified using (7.4)27 while em-

ploying the results (presented in Figs. 7.6 and 7.7) obtained using unbal-

anced load flow analysis, made by each of the lines to the resultant voltage

unbalance levels (presented in Fig. 7.3) at the various busbars (S2, S4 and

S6 - S928 ).
t
|V−:Si |cos(θV−:Si
t − θV−:Si
lines )
Ct/Si = lines
× 100% (7.4)
|V−:Si |

where,

Ct/Si - contribution made by any line t to the resultant voltage unbalance

level at any busbar Si


t
V−:Si - negative sequence voltage caused by the line t at the busbar Si
lines
V−:Si - resultant negative sequence voltage at the busbar Si
t
θV−:Si
t , θV−:Si
lines - phase angles of the negative sequence voltages V−:Si and

lines
V−:Si respectively

Fig. 7.10 clearly demonstrates the dominant contributions made by lines

A, F and I as identified above using the proposed technique. Further, the

assessments made on the role of lines J and D above as a less of a con-


27
This quantifies the fraction of the negative sequence voltage caused by any line t which is in-
phase with the resultant negative sequence voltage, as a ratio to the magnitude of the resultant
negative sequence voltage.
28
Significantly affected busbars.
164

tributor and as a negative contributor respectively are also seen to be in

agreement with the results presented in Fig. 7.10.

50
S2
40 S4
S6
30 S7
S8
Contribution (%)

20 S9

10

-10 S8
A B S6
C D Busbar
E F G H S2
I J K L
Line M N

Figure 7.10: Nodal contributions made by the individual lines to the resultant voltage
unbalance levels

• Transposition options for better managing the emission arising as a result of

line asymmetries.

– Assessments on the study system using the proposed technique:

Referring to Fig. 7.11 (I), which is deduced using Fig. 7.8 by representing

the smaller global emission vectors of lines B, C, L, M and N using a single

vector (labelled ‘B+C+L+M+N’), the most effective option to correct the

network through the transposition of a single line is seen to be the selection


165

I I I

D D D
B+C+L+M+N B+C+L+M+N B+C+L+M+N
J J J
F A A
Resultant vector after
transposing lines F and A
Resultant vector after
transposing line F

Rlines
(I) (II) (III)

Figure 7.11: (I) Deduced from Fig. 7.8 (II) Effect of the transposition of line F only
(III) Effect of the transposition of lines A and F together (drawn approximately to a
scale)

of line F as it is being represented by the largest vector among the group

of vectors (i.e. B+C+L+M+N, A and F) clustered together. Further

correction can be introduced to the network effectively by transposing lines

A and F. This results in the vectors B+C+L+M+N, D, I and J to remain,

out of which the vectors B+C+L+M+N and I lie in close proximity. This

is the case with the vectors D and J as well. The phase difference close

to 900 between these two groups (i.e. B+C+L+M+N and I, and D and

J) suggests that the emissions arising as a result of lines B, C, I and L

- N assist in counter balancing some of the emissions of lines D and J.

The effects of the transposition of line F only and lines A and F together

are illustrated in Figs. 7.11 (II) and (III) respectively. These demonstrate

that the transposition of line F can introduce approximately 30% reduction

in the resultant influence, whereas approximately 50% reduction in the

resultant influence can be expected by transposing both lines A and F.

– Validation of the above assessments using unbalanced load flow analysis:

Fig. 7.12 illustrates the effects, in terms of the residual VUF values at the
166

various busbars obtained using unbalanced load flow analysis29 , of the two

line transposition options identified above. This also shows the voltage

unbalance levels which arise as a result of the existing line asymmetries

(reproduction of Fig. 7.3). Fig. 7.12 demonstrates, as identified above

using the proposed technique, that the transposition of line F introduces

approximately 30% reduction in the voltage unbalance levels caused by the

existing line asymmetries at S2, S4 and S6 - S9 (busbars which are signifi-

cantly affected by the line asymmetries as noted in Fig. 7.3), whereas this

reduction is approximately 50% when both lines A and F are transposed.

1.6 Exisitng line asymmetries


After transposing line F
1.2 After transposing lines A and F
VUF (%)

0.8

0.4

0.0
S2 S3 S4 S5 S6 S7 S8 S9

Busbar

Figure 7.12: Effects, obtained using unbalanced load flow analysis, of the transposi-
tion of line F only, and lines A and F together

29
Operating scenario corresponds to the selected time stamp.
167

7.3 Voltage Unbalance Behaviour of Load Asymmetries

7.3.1 Impact of the Load Asymmetries of the Study System on the

Voltage Unbalance Problem

It has been noted, based on measurements, that the loads supplied by the study

system exhibit unbalance with regard to their active power components, whereas the

reactive power components have been seen to be reasonably balanced30 . Table 7.3

gives the distribution of the active (Pa , Pb and Pc ) and reactive (Qa , Qb and Qc )

power across the three phases at each of the load busbars, which corresponds to the

selected time stamp. It also gives the degree of asymmetry associated with the three

active power components, in terms of the standard deviation (µP ), at each of the

busbars.

Table 7.3: Distribution of the active and reactive power across the three phases at
each of the load busbars of the study system

Load busbar S2 S3 S4 S7 S8 S9
Pa (MW) 6.32 6.24 3.87 11.3 2.04 0.57
Pb (MW) 5.87 6.08 3.43 11.3 1.89 0.55
Pc (MW) 5.87 5.96 3.37 11.38 2.04 0.54
Qa ≈ Qb ≈ Qc 0.12 2.37 1.705 4.84 0.63 0.08
(MVAr)
µP 0.26 0.14 0.27 0.05 0.09 0.02

Similar to the case of the line asymmetries presented in Section 7.2.1, the impact

of the interaction of all unbalanced loads of the study system on the problem of volt-
30
This can be considered as a general case for higher voltage systems which are usually associated
with loads with a near unity power factor. That is, in comparison to the level of active power,
systems supply only a low level of reactive power at a load point, of which the level of unbalance
can be considered to be insignificant compared to that of the reactive power.
168

age unbalance is established in terms of the VUF values at the various busbars using

unbalanced load flow analysis. This is accomplished by synthesising the system oper-

ation31 while parameterising the loads as given in Table 7.3, and the lines assuming

they are ideally transposed. Fig. 7.13 illustrates these results, in comparison to the

emission levels arising as a result of the line asymmetries (reproduction of Fig. 7.3).

This reveals that the emission which arises due to the loads themselves is also consid-

erable (highest VUF: 1% at S9), although the line asymmetries are seen to dominate

the problem.

1.5
Load asymmetries
Line asymmetries
1.2
VUF (%)

0.9

0.6

0.3

0.0
S2 S3 S4 S5 S6 S7 S8 S9
Busbar

Figure 7.13: Nodal VUF values which arise as a result of the load asymmetries, in
comparison to that of the line asymmetries

31
Which corresponds to the selected time stamp.
169

7.3.2 Voltage Unbalance Behaviour of the Individual Loads of the

Study System - as Elements in the Interconnected Network

Based on a similar approach taken to observe the voltage unbalance behaviour of

the individual lines as elements in the interconnected network32 , this for the indi-

vidual loads is obtained using unbalanced load flow analysis by considering a single

load (referred to as ‘load under observation’) at a time, where it is applied as de-

fined in Table 7.3 while parameterising all other loads and all lines to represent a

balanced behaviour.

Fig. 7.14 illustrates the VUF values which arise as a result of each of the loads

as a load under observation at the various busbars. These results are summarised

in Table 7.4 by giving a rank to each of the loads based on its emission level in

a global sense33 (column 5). For brevity, the information given in Table 7.3 are

repeated together with commentaries on the degree of asymmetry associated with

the real power components (column 2), and the three-phase loading level (column 3).

A commentary on the location (i.e. US, CP, DS) of each of the loads in the system

is also given in column 4. Referring to the entries in Table 7.4, the following can

be derived:

• Three-phase loading level is not a factor which governs the level of emission

introduced by an unbalanced load34 .

• As it is known, the degree of asymmetry associated with a load is the primary

factor that needs consideration when assessing the level of emission introduced

by the load35 .
32
See Section 7.2.3.
33
Taking the highest emission level as the reference, as in the case of the line asymmetries.
34
e.g. compare the emission levels introduced by the loads at S3 and S8 referring to their three-
phase loading levels (column 3).
35
e.g. compare the emission levels introduced by the loads at S9 and S8 referring to their µP .
170

• In contrary to the above, the load supplied by S8 (µP = 0.09) located in the

downstream part of the system is seen to introduce the highest level of emission,

although the associated degree of asymmetry is considerably lower than that of

the loads supplied by S2 and S4 (µP ≈ 0.3) located in the upstream part. It

is therefore evident that, for a similar degree of asymmetry, the loads supplied

by the busbars located in the downstream part of the system tend to introduce

increased emission levels compared to those located in the upstream.

0.6 S2
S3
0.5 S4
S5
0.4
S6
VUF (%)

S7
0.3
S8
S9
0.2

0.1

0.0
S2 S3 S4 S7 S8 S9
Load under observation

Figure 7.14: Nodal VUF values which arise as a result of the individual loads

Fig. 7.15 illustrates the phase angles of the negative sequence voltages which arise

as a result of the individual loads supplied by S2, S4, S7 - S936 at the various busbars

(S2 - S9). This demonstrates that, as in the case of the line asymmetries (Fig. 7.7),

the above individual loads yield a unique and nearly constant phase angle across all

the busbars.
36
Which introduce considerable emission levels as noted in Table 7.4.
171

Table 7.4: Operating features and emission levels of the individual loads of the study
system

Load µP Three-phase loading Location in Emission


busbar level (MW/MVAr) the network level
S4 High High US High
(0.27) (11/5.115)
S2 High Very high US High
(0.26) (18/0.36)
S3 Low Very high US: a remote Negligible
(0.14) (18/7.11) busbar
S8 Low Low DS Very high
(0.09) (6/1.89)
S7 Very low Extra high CP Low
(0.05) (34/14.52)
S9 Very low Very low DS Low
(0.02) (2/0.08)

250

200
Phase angle (deg.)

Load at S2
150 Load at S4
Load at S7
100 Load at S8
Load at S9
50

0
S2 S3 S4 S5 S6 S7 S8 S9

Busbar

Figure 7.15: Phase angles of the nodal negative sequence voltages introduced by the
individual loads
172

Arising as a result of a load under observation (LSi ), the negative sequence volt-
LSi
age V−:rec−tany
at the receiving end busbar of any line tany 37 (in other words, at any

busbar) has to satisfy:

LSi LSi
V−:rec−tany
= V−:send−tany
− Z++:tany I−:LSi /tany (7.5)

where,
LSi
V−:send−tany
- negative sequence voltage introduced by the load under observation at

the sending end busbar of any line tany

I−:LSi /tany - negative sequence current in the line tany , which arises as a result of the

load under observation

Equation (7.5) provides an explanation to the results presented in Figs. 7.15 and 7.16:

• The negative sequence currents I−:LSi /tany flow through the lines (i.e. through

the impedances Z++:tany ) are drawn by the load under observation. The level

of the negative sequence current drawn by this load is governed by the degree

of its asymmetry regardless of the three-phase loading level. That is, as seen

in Fig. 7.15, the principal intrinsic parameter that determines the negative se-
LSi
quence voltage V−:rec−tany
is the degree of asymmetry associated with the load

under observation. The location of this load in the system determines the part

of the network in which the negative sequence currents I−:LSi /tany flow38 , making
LSi
the negative sequence voltage V−:rec−tany
dependant secondarily on the location

of the load under observation as indicated by the results presented in Fig. 7.15.

37
Note that lines are taken as transposed for this case of observing the voltage unbalance behaviour
of individual loads. Thus, the impedance Z−+:tany = 0.
38
e.g. when the load is supplied by a upstream busbar, the negative sequence currents I−:LSi /tany
mainly flow through the lines located in the upstream part of the network.
173

• The phase angles of the negative sequence voltages which arise as a result of the

load under observation at the various busbars should be associated with the term

−Z++:tany I−:LSi /tany . The phase angle of the impedance Z++:tany depends only

on the X/R ratio of the line which has been noted to be nearly identical39 for

all lines of the study network. The phase angle of the negative sequence current

I−:LSi /tany is governed by the order of the distribution of the three active power

components of the load under observation across the three phases. Thus, the

phase angles of the negative sequence voltages at the various busbars remain

similar and unique for a particular load under observation as seen in Fig. 7.15.

Load at S8

Load at S7

Load at S2

Load at S9

Load at S4

Figure 7.16: Global emission vectors of the individual loads (drawn approximately to
a scale)

39
This can be considered as a general case.
174

7.3.3 General Outcomes

As for an asymmetrical line40 , the voltage unbalance behaviour exhibited by an unbal-

anced load of an interconnected network can be represented using a global emission

vector of which:

• The magnitude, as summarised in Table 7.4 - column 5 for the loads of the study

system, can be approximately assessed by referring to the degree of asymmetry

associated with the load, and the location of the load in the network.

• The phase angle, as illustrated in Fig. 7.15 for the loads of the study system,

can be approximately derived by referring to the order of the distribution of

power of the load under observation across the three phases, and the X/R ratio

associated with lines of the network.

The global emission vectors of the individual loads41 S2, S4 and S7 - S9 of the

study system are illustrated in Fig. 7.16.

Employing the linearity of negative sequence variables, as for line asymmetries42 ,

the global emission vectors of individual loads can be integrated to establish a basis

that provides a comprehensive understanding of the manner in which various un-

balanced loads interact with each other to form the resultant influence. Fig. 7.17

illustrates this resultant influence for the study system using a single vector (Rloads )

which is obtained by the summation of the individual global emission vectors shown

in Fig. 7.16.
40
See Section 7.2.4.
41
Which introduce considerable emission levels.
42
See Section 7.2.5.
175

Rloads Load at S8

Load at S7

Load at S2

Load at S9

Load at S4

Figure 7.17: Resultant influence of the interaction of all unbalanced loads (drawn
approximately to a scale)
176

7.4 Combined Voltage Unbalance Behaviour of Line and Load

Asymmetries

7.4.1 Combined Impact of the Line and Load Asymmetries of the

Study System on the Voltage Unbalance Problem

When both the line and load asymmetries exist concurrently in the study system43 ,

their impact on the problem of voltage unbalance is established in terms of the VUF

values at the various busbars using unbalanced load flow analysis. This is accom-

plished by synthesising the system operation44 while parameterising the loads as given

in Table 7.3, and the lines based on their actual construction. Fig. 7.18 illustrates

these results, in comparison to the emission levels which arise as a result of the line

asymmetries alone (reproduction of Fig. 7.3) and the load asymmetries alone (re-

production of Fig. 7.13). The measured voltage unbalance levels (reproduction of

Fig. 7.2) are also shown in Fig. 7.18. This reveals that the interaction of the various

line and load asymmetries causes voltage unbalance levels up to 1.8% at S8 and S9

(downstream), 1.2% at S6 and S7 (central part) and 0.6% at S2 and S4 (upstream),

which are seen to be in close agreement with the measured values.

7.4.2 Representation of the Voltage Unbalance Behaviour of the Entire

System

Employing the linearity of negative sequence variables, the global emission vectors

of individual sources of unbalance (i.e. individual lines and loads) can be integrated

to establish a basis that provides a comprehensive understanding of the manner in


43
Which represents the real operating scenario, where voltage unbalance levels arise as a result of
the interaction of both the line and load asymmetries.
44
Which corresponds to the selected time stamp.
177

2.0 Load asymmetries


Line asymmetries
1.6 Both line and load asymmetries
Measured values
1.2
VUF (%)

0.8

0.4

0.0
S2 S3 S4 S5 S6 S7 S8 S9
Busbar

Figure 7.18: Nodal VUF values which arise as a result of both the line and load asym-
metries, in comparison to that of the line asymmetries alone, and the load asymmetries
alone, and also to the measured values

which various untransposed lines and unbalanced loads interact with each other to

form the overall influence. Fig. 7.19 illustrates this overall influence for the study

system using a single vector (Rsystem ) which is obtained by the summation of the

individual global emission vectors shown in Figs. 7.9 (for lines) and 7.16 (for loads).

Fig. 7.19 also shows the vectors Rlines (Fig. 7.10) and Rloads (Fig. 7.17) where

Rsystem ≈ Rlines + Rloads .

This can be used to make the following assessments on the problem of voltage

unbalance, which are also confirmed using unbalanced load flow analysis:

• Contributions made by the line and load asymmetries to the overall voltage

unbalance levels.

– Using the proposed technique:

Referring to the vectors Rsystem , Rlines and Rloads , the component of Rlines

which is in-phase with Rsystem accounts approximately for 60% of the mag-
178

Rload

Rsystem

Rlines

Figure 7.19: Resultant influence of the interaction of all lines and loads (drawn ap-
proximately to a scale)
179

nitude of Rsystem , whereas that of Rloads is approximately 30%. That is,

the asymmetry associated with the lines is the dominant contributor to

the problem, whereas the load asymmetries play only a secondary role.

– Using unbalanced load flow analysis:

Fig. 7.20 illustrates the contributions, quantified45 employing the results

obtained using unbalanced load flow analysis, made by the line and load

asymmetries to the overall voltage unbalance levels (presented in Fig. 7.18)

at the various busbars (S2, S4 and S6 - S946 ). Fig. 7.20 clearly demon-

strates that, as identified above using the proposed technique, the line

asymmetries contribute approximately by 60% - 70% to the overall volt-

age unbalance levels at S2, S4 and S6 - S9, whereas it is approximately

25% - 30% by the load asymmetries.

80
Line asymmetries
Load asymmetries

60
Contribution (%)

40

20

0
S2 S4 S6 S7 S8 S9
Busbar

Figure 7.20: Nodal contributions made by the line and load asymmetries to the overall
voltage unbalance levels

45
Based on the approach given by (7.4).
46
Significantly affected busbars.
180

• Contributions made by the individual sources of unbalance to the overall voltage

unbalance levels.

– Using the proposed technique:

Observation of the global emission vectors of the individual lines and loads

illustrated in Figs. 7.9 and 7.16 respectively together with the vector

Rsystem (Fig. 7.19) suggests that, among all sources of unbalance, lines

F and I, these being represented by the largest and the closest vectors to

Rsystem , can be identified as the major contributors to the overall voltage

unbalance levels. In addition, the vectors of line A and the loads sup-

plied by S2 and S4, having relatively large magnitudes and being closer

to Rsystem , can contribute significantly to the problem supporting the two

major contributors (i.e. lines F and I). The phase deviation close to 900

of the vector of the load supplied by S8 with respect to Rsystem can make

it a less of a contributor. The positioning of the vectors of lines D and J

and the load supplied by S7 with respect to Rsystem suggests that they can

make negative contributions.

– Using unbalanced load flow analysis:

Fig. 7.21 illustrates the contributions, quantified employing the results

(presented in Figs. 7.6, 7.7, 7.14 and 7.15) obtained using unbalanced load

flow analysis, made by each of the sources of unbalance to the overall

voltage unbalance levels (presented in Fig. 7.18) at the various busbars

(S2, S4 and S6 - S9). This clearly demonstrates the dominant contributions

made by lines F and I as identified above using the proposed technique.

Further, the assessments made on the role of line A and the loads supplied

by S2 and S4 as significant contributors, the load supplied by S8 as a less

of a contributor, and lines D and J and the load supplied by S7 as negative


181

contributors are also seen to be in agreement with the results presented in

Fig. 7.21.

S2
50
S4
S6
40
S7
S8
30
S9
Contribution (%)

20

10

-10

-20
Load at S2
Load at S3
Load at S4
Load at S7
Load at S8

S7
Load at S9

B
A

C
D
E
F

S2 Busbar
G
H
I
J
K
L
M
N

Source of unbalance

Figure 7.21: Nodal contributions made by the individual sources of unbalance to the
overall voltage unbalance levels

7.5 Chapter Summary

This chapter has established theoretical bases to broaden the understanding of the

voltage unbalance behaviour exhibited by various sources that exist in interconnected

network environments. These would provide additional assistance in the application

of the IEC/TR 61000-3-13 voltage unbalance allocation guidelines.


182

Employing a 66kV interconnected sub-transmission system as the study case, de-

terministic studies were carried out with the view to develop insights into the in-

fluences made by line and load asymmetries on the problem in a systematic man-

ner considering each of the asymmetrical elements. The following can be drawn

from the study:

• The voltage unbalance behaviour (in terms of the level of voltage unbalance

and its phase angle) exhibited by an asymmetrical line at various busbars of

an interconnected network is primarily governed by the term −Z−+:t I+:t , and

secondarily dependant on the location of the line in the network. This be-

haviour, in a global sense, can be ascertained by a single vector referred to as

‘global emission vector’ of which the magnitude can be approximately assessed

by referring to the product |Z−+:t I+:t | of the line and the location of the line in

the network, and the phase angle can be approximately derived using the term

−Z−+:t I+:t .

• As for an asymmetrical line, the voltage unbalance behaviour exhibited by an

unbalanced load of an interconnected network can also be represented using

a global emission vector. Of this global emission vector, the magnitude can

be approximately assessed by referring to the degree of asymmetry associated

with the load and the location of the load in the network. The phase angle

can be approximately derived by referring to the order of the distribution of

power of the load under observation across the three phases, and the X/R ratio

associated with lines of the network.

• The global emission vectors of individual sources of unbalance (i.e. of individual

lines and loads) can be integrated to establish a basis that provides a compre-

hensive understanding of the manner in which various untransposed lines and


183

unbalanced loads interact with each other to form the overall system behaviour.

That is, the overall influence of various sources of unbalance that exist in an in-

terconnected network can be represented, in a global sense, using a single vector.

This basis can be used to assess the interconnected networks on the problem of

voltage unbalance in different ways, e.g. principal contributors, most favorable

corrective options. This proposed technique was applied to the study system

to identify the major contributors to voltage unbalance levels and the effective

line transposition options, which were also confirmed using unbalanced load

flow analysis.
Chapter 8

Conclusions and Recommendations

for Future Work

8.1 Conclusions

This thesis has focused on making contributions for further development of the re-

cently released Technical Report IEC/TR 61000-3-13 which provides guiding prin-

ciples for coordinating voltage unbalance between various voltage levels of a power

system through the allocation of emission limits to installations. A number of related

aspects of this voltage unbalance allocation procedure such as the global emission

arising as a result of system inherent asymmetries, the propagation of voltage unbal-

ance and associated deficiencies have been addressed. Furthermore, theoretical bases

to broaden the understanding of the influence made by various sources of unbalance

that exist in interconnected networks on voltage unbalance, which would provide ad-

ditional assistance in the application of these voltage unbalance allocation guidelines

have been established.

184
185

As an essential tool required for specific applications of the work presented in this

thesis, an unbalanced load flow algorithm based on the phase coordinate reference

frame was developed. The component level load flow constraints and the three-phase

modelling of system components were incorporated. The major task of this work

was to develop models for the representation of three-phase induction motors, which

overcome the limitations associated with existing models.

Preliminary studies carried out on the global emission which arises as a result

of system inherent asymmetries revealed that this emission has some degree of load

dependency. The approach given in IEC/TR 61000-3-13 to evaluate the influence

of an asymmetrical radial line on the global emission does not consider this load

dependency, and thus it can be applied only when the line supplies passive loads. This

IEC approach was noted to be conservative when the line supplies large proportions

of three-phase induction motors. In this case of motor loads, the negative sequence

currents arising as a result of line asymmetries made a significant influence on the

global emission, whereas it was seen to be insignificant for passive loads. The degree

of this influence was seen to be dependant primarily on the motor proportion, and

secondarily on the system and motor characteristics. In essence, the global emission

levels in HV power systems, in the presence of considerable proportions of induction

motors, were seen to arise not only as a result of the local HV lines but also as a

result of the downstream MV line asymmetries.

Systematic approaches, covering radial and interconnected networks, for evaluat-

ing the global emission which arises due to line asymmetries at nodal level, taking

the load dependency of this emission into account were proposed. These were verified

using unbalanced load flow analysis in relation to simple test systems. These test

results revealed that the presence of considerable proportions of induction motors

increases the emission, compared to emission levels when only passive loads exist, at
186

the busbar which is directly connected to the upstream system. At all other network

busbars, induction motor loads were seen to reduce the emission levels which arise due

to the local line asymmetries, compared to that when only passive loads exist. In HV

power system, in the presence of motor loads, the influence of the downstream MV

line asymmetries was noted to either decrease or increase the resultant emission lev-

els with respect to the local emission levels depending on the impedance/admittance

characteristics of the downstream lines relative to the local HV lines. The proposed

approach was further verified using the IEEE 14-bus test system taking it as an HV

system supplying passive loads at the HV level itself.

The propagation of voltage unbalance from higher voltage to lower voltage systems

in terms of transfer coefficients was initially examined in the presence of four basic

load types (i.e. constant impedance, constant current, constant power and three-

phase induction motor loads).

• It was seen that the value of unity which has been assumed in IEC/TR 61000-

3-13 for the MV to LV transfer coefficient in the presence of passive loads in

general is conservative in relation to some passive load types such as commonly

prevailing constant power loads. Transfer coefficients for passive loads were
τ
1
seen to be scaled up by the factor 1 + j ksc−s ∠θpf :s relative to unity, where

τ = 0, 1, 2 ∼ 3 for constant impedance, constant current and constant power

loads respectively1 . Noting that the factor ksc−s can be a value in the range

of 5 to 25 for various systems and the exponent τ varies in the range of 0 to

3 for different load types, a uniform behaviour or the behaviour of constant

impedance loads as assumed in the IEC approach cannot be used to represent

all load types with a high degree of accuracy. That is, voltage unbalance can get
1
ksc−s - ratio between the short-circuit capacity (in MVA) at the busbar S under evaluation and
the total load (in MVA) supplied by the busbar S, and θpf :s - pf angle (− and + for lagging and
leading conditions respectively) at the busbar S.
187

significantly amplified, especially for constant power loads, as it propagates from

higher voltage to lower voltage levels. As the values of ksc−s for MV systems

are usually smaller than those for LV systems, this amplification in the case of

the HV to MV propagation is greater than that in the MV to LV propagation.

• As suggested in IEC/TR 61000-3-13 in relation to the MV to LV transfer coeffi-

cient, voltage unbalance was seen to get significantly attenuated as it propagates

from higher voltage to lower voltage levels in the presence of induction motor

loads. This attenuation, relative to unity, can be given by the scaling factors
1 01
„ « and 2 13 in the cases of the MV to LV and HV
1+ k ks
“ ”
41+ 1 ks
sc−lv ksc−mv
@
ks
A5
1+
ksc−lv
agg
to MV propagation respectively2 . As ksc−mv for MV systems < ksc−lv for LV

systems in practice, a higher degree of this attenuation can be expected in the

HV to MV propagation than that in the MV to LV propagation. Noting that

5 < ks < 7, the degree of this reduction of the transfer coefficients relative to

unity is significant compared to the increment introduced by passive loads.

Incorporating the above, systematic methods for evaluating the MV to LV and HV to

MV transfer coefficients were developed. These were verified using unbalanced load

flow analysis for various scenarios. It was demonstrated that, compared to the IEC

method, the proposed new formulation for assessing the MV to LV transfer coefficient

gives a more accurate estimation, particularly for load bases which are dominated by

passive elements. The MV to LV and HV to MV transfer coefficients were noted to

vary in the ranges of 0.6 to 1.1 and 0.5 to 1.4 respectively depending on the system

and load characteristics and the downstream load composition.

2
Motor loads are considered to be supplied at the LV level. ks - ratio between the positive and
negative sequence impedances of the aggregated motor load supplied by the LV system, ksc−lvagg -
ratio between the short-circuit capacity (in MVA) at, and the total load (in MVA) supplied by, the
aggregated LV busbar.
188

Preliminary studies carried out on the propagation of voltage unbalance from one

busbar to other neighbouring busbars of a sub-system in terms of influence coefficients

revealed that, for a simple two-bus radial sub-system, the influence coefficients can

be approximated to unity in the presence of passive loads in general. However, these

influence coefficients were seen to be considerably smaller than unity (in other words,

than that in the case of passive loads) for induction motor loads. The conclusion of

this observation is that the negative sequence currents which arise as a result of the

voltage unbalance that exists at a particular busbar introduce a considerable impact

on influence coefficients for motor loads, whereas it is insignificant for passive loads.

Employing this, a systematic method for evaluating influence coefficients for inter-

connected network environments was developed. This was verified using unbalanced

load flow analysis in relation to a simple test system, and also to the IEEE 14-bus

test system supplying local passive loads.

The IEC/TR 61000-3-13 voltage unbalance allocation approach was examined

employing a simple test system with and without the inclusion of the influence of

system inherent asymmetries. For both of the cases stated above, the IEC/TR 61000-

3-13 procedure was noted to lead to situations where the set planning level are being

exceeded even when no customer exceeds the allocated limit. Closely aligning with the

IEC/TR 61000-3-13 guidelines, a revised voltage unbalance allocation policy based

on the constrained bus voltage (CBV) harmonics/flicker allocation principles was

proposed. In this CBV allocation technique, emission levels at network busbars are

explicitly forced, introducing a new factor referred to as ‘allocation constant’, to be at

or below the set planning levels when all loads inject their limits derived under the new

approach. The issue of the emission arising as a result of system inherent asymmetries

involved with the case of voltage unbalance was taken into account according to

IEC/TR 61000-3-13 using the factor Kue. This revised allocation technique was
189

examined employing the above mentioned test system, and it was seen to restrict the

resulting emission level at all network busbars at or below the set limit. Further, this

proposed alternative technique was noted to be able to satisfy all four key allocation

objectives.

With the view of developing theoretical bases which describe the voltage unbal-

ance behaviour exhibited by various sources that exist in interconnected network

environments, deterministic studies supported by unbalanced load flow analysis were

carried employing a 66kV interconnected sub-transmission system. This study net-

work has been monitored to experience voltage unbalance levels above the applicable

regulatory limit of 1%.

• The influence made by each of the untransposed lines, as a standalone line

and also as an element in the interconnected network, on voltage unbalance

was established employing a new concept termed ‘voltage unbalance emission

vector’, and extensively analysed. The voltage unbalance behaviour (in terms of

the magnitude and the phase angle) exhibited by an asymmetrical line at various

busbars of an interconnected network was seen to be primarily governed by the

term −Z−+:t I+:t as in the case of a standalone line, and secondarily dependant

on the location of this line in the network. This behaviour, in a global sense, was

ascertained by a single vector referred to as ‘global emission vector’ of which the

magnitude can be approximately assessed by referring to the product |Z−+:t I+:t |

of the line and the location of this line in the network. The phase angle of this

vector can be approximately derived using the term −Z−+:t I+:t .

• Following an analysis of the voltage unbalance emission vectors introduced by

each of the unbalanced loads as an element in the interconnected system, it was

seen, as for an asymmetrical line, that the voltage unbalance behaviour exhib-
190

ited by an unbalanced load of an interconnected network can also be represented

using a global emission vector. Of this global emission vector, the magnitude

can be approximately assessed by referring to the degree of asymmetry associ-

ated with the load and the location of this load in the network, and the phase

angle can be approximately derived by referring to the order of the distribution

of power of the load under observation across the three phases and the X/R

ratio associated with lines of the network.

• It was seen that the voltage unbalance behaviour exhibited by individual lines

and loads can be ascertained, in a global sense, in terms of the global emission

vectors. Based on the linearity of negative sequence variables, these individual

global emission vectors were integrated to establish a basis that provides a com-

prehensive understanding of the manner in which various untransposed lines and

unbalanced loads interact with each other to establish the overall system be-

haviour. That is, the overall influence of various sources of unbalance that exist

in an interconnected network can be represented, in a global sense, using a single

vector. This basis can be used to assess interconnected networks in relation to

the problem of voltage unbalance in different ways, e.g. principal contributors,

most favourable corrective options. Further, this knowledge which facilitates

the identification of contributions made by individual unbalanced sources forms

a platform for developing techniques to assess the compliance against emission

limits, which is another subject of relevance to future editions of IEC/TR 61000-

3-13. This proposed technique was applied to the study system to identify the

major contributors to voltage unbalance levels and the effective line transposi-

tion options, which were also confirmed using unbalanced load flow analysis.
191

• In the study system, the line asymmetries were seen to contribute approxi-

mately by 60% - 70% to the overall voltage unbalance levels, whereas the con-

tribution made by the load asymmetries was only 25% - 30%. This indicates

the requirement of proper line transposition practices even at relatively low

sub-transmission voltage levels, which is currently not seen to be a common

practice.

8.2 Recommendations for Future Work

The Technical Report IEC/TR 61000-3-13, which will possibly be adopted as an

Australian standard in the future, provides valuable techniques for managing voltage

unbalance in public power systems. In the cases of the counterpart Australian/New

Zealand harmonics (AS/NZS 61000-3-6 [75]) and flicker (AS/NZS 61000-3-6 [76])

standards, which are essentially based on the respective IEC technical reports, amend-

ments has been shown to require in order to allow accurate and straightforward

implementation by utilities and customers. Consequently, Standards Australia com-

missioned the writing of the handbook HB-264-2003 [4] which gives more prescriptive

procedures for the use of these harmonics and flicker standards. Similarly, in adopting

IEC/TR 61000-3-13 as an Australian/New Zealand standard, further work is seen to

require in order to assist the application of such a standard to complex systems such

as MV distribution systems with relatively long feeders.

From a utility point of view, a significant improvement to the voltage allocation

process would be the development of software modules which are able to produce

various factors (e.g. Kue factor) and coefficients (e.g. transfer and influence coeffi-

cients) directly from network databases. Such software modules would also benefit

customers, as the emission limits can be evaluated without substantial involvement

or effort on the part of the responsible utility.


192

Although the Technical Report IEC/TR 61000-3-13 provides the guidance for eval-

uating customer emission limits, it does not discuss the assessment of the compliance

against these limits. Development of systematic techniques to carry out this would sig-

nificantly assist in further developing IEC/TR 61000313 such that a complete voltage

unbalance management procedure is prescribed. In fact, the compliance assessment

has been noted to be a subject of interest to a currently active CIGRE/CIRED work-

ing group which is responsible for updating IEC/TR 61000-3-13. The work presented

in Chapter 7 of this thesis provides a fundamental basis towards the development of

such techniques, yet a substantial amount of further work is required for completing

the task.

Variation in network topologies and loads makes the setting of prescriptive generic

planning levels and voltage unbalance allocation procedures impractical. At present,

the planning levels proposed in IEC/TR 61000-3-13 are given as indicative values. The

significant differences between networks mean that there are likely to be situations

in which it may suit utilities to choose alternative planning levels. Guidance on how

this might best be accomplished would be a useful advancement.

The Technical Report IEC/TR 61000-3-13, as well as this thesis, has focused

mainly on Stage 2 (i.e. allocation of emission limits to individual customers) of the

voltage unbalance management process. Prescriptive Methodologies governing the

third and final stage, or as to how a customer who would fail to comply with the

Stage 2 emission limit to be treated are yet to be covered.

Verification of the methodologies proposed in Chapters 3 - 5 through laboratory

level experiments or field measurements is encouraged. To confirm the outcomes pre-

sented in Chapter 7, extended work such as the carrying out suitable field measure-

ments and the examination of the proposed techniques in relation to other networks
193

with different topologies are recommended. Chapter 7, which can be considered as an

initial step towards a broad research area, requires more deterministic, analytical and

theoretical work such that the techniques are further developed to be more definitive.

As an example, Chapter 7 only covers a relative assessment of the influence and the

location of a source of unbalance that exists in an interconnected network, for which

absolute assessment techniques can be developed.


References

[1] IEC/TR 61000-3-13: Electromagnetic compatibility (EMC) - limits - assessment

of emission limits for the connection of unbalanced installations to MV, HV

and EHV power systems, Ed. 1. Technical report, International Electrotechnical

Commission, 2008.

[2] IEC/TR 61000-3-6: Electromagnetic compatibility (EMC) - limits - assessment

of emission limits for distorting loads in MV and HV power systems, Ed. 1.

Technical report, International Electrotechnical Commission, 1996.

[3] IEC/TR 61000-3-7: Electromagnetic compatibility (EMC) - limits - assessment

of emission limits for fluctuating loads in MV and HV power systems, Ed. 1.

Technical report, International Electrotechnical Commission, 1996.

[4] HB 264: Power quality - recommendation for the application of

AS/NZS 61000.3.6 and AS/NZS 61000.3.7. Technical report, Standards Aus-

tralia/New Zealand, 2003.

[5] Timothy James Browne. Harmonic management in transmission networks. PhD

thesis, School of Electrical, Computer and Telecommunications Engineering, Uni-

versity of Wollongong, Wollongong, Australia, March 2008.

[6] M.B. Hughes. Revenue metering error caused by induced voltage from adjacent

transmission lines. IEEE Trans. on Industry Applications, 7(2):741–745, April

1992.

[7] M. Tavakoli Bina and E. Pasha Javid. A critical overview on zero sequence

component compensation in distorted and unbalanced three-phase four-wire sys-

tems. In Proc. International Power Engineering Conference (IPEC 2007), pages

1167–1172, December 2007.


194
195

[8] J.E. Parton and Y.K. Chant. The three-limbed phase transformer with controlled

zero sequence effect. IEEE Trans. on Power Apparatus and Systems, 90(5):2019–

2029, September 1971.

[9] EN 50160: Voltage characteristics of electricity supplied by public distribution

networks. Technical report, CENELEC, 2007.

[10] NRS 048-2: Electricity supply - quality of supply, part 2 - voltage characteristics,

compatibility levels, limits and assessment methods, Ed. 3. Technical report,

Standards South Africa, 2007.

[11] National Electricity Code Australia, Version 1.0 - Amendment 9.0. Technical

report, National Electricity Code Administrator Limited, Australia, 2004.

[12] CIGRE/CIRED Joint Working Group C4.07 (formerly GIGRE WG 36.07).

Power quality indices and objectives. Technical report, January 2004.

[13] CIGRE/CIRED Joint Working Group C4.103 (formerly C4.06). Assessment of

emission limits for the connection of disturbing installations to power systems,

Final report. Technical report, June 2007.

[14] CIGRE/CIRED Joint Working Group C4.103. Assessment of emission limits for

the connection of unbalanced installations to MV, HV and EHV power systems

(proposed future IEC 61000-3-13). Technical report, April 2006.

[15] UIE guide to quality of electrical supply for industrial installations - part 1: gen-

eral introduction to electromagnetic compatibility (EMC) types of disturbances

and relevant standards, Ed. 1. Technical report, 1994.

[16] UIE guide to quality of electrical supply for industrial installations - part 4:

voltage unbalance, Ed. 1. Technical report, 1998.


196

[17] A.K. Singh, G.K. Singh, and R. Mitra. Some observations on definitions of

voltage unbalance. In Proc. 39th North American Power Symposium (NAPS

’07), September-October 2007.

[18] P. Pillay and M. Manyage. Definitions of voltage unbalance. IEEE Power Engi-

neering Review, 21(5):49–51, May 2001.

[19] IEC 61000-4-30: Electromagnetic compatibility (EMC) - part 4-30 - environment

- testing and measurement techniques - power quality measurement methods, Ed.

1. Technical report, International Electrotechnical Commission, 2003.

[20] NEMA MG1: Motors and generators. Technical report, National Electricity

Manufacturer’s Association, 1993.

[21] IEEE 112: Standard test procedures for polyphase induction motors and gener-

ators. Technical report, Institute of Electrical and Electronics Engineers, 1991.

[22] IEEE 100: standard dictionary of electrical and electronics terms. Technical

report, Institute of Electrical and Electronics Engineers, 1996.

[23] IEEE 1159: IEEE recommended practice for monitoring electric power quality.

Technical report, Institute of Electrical and Electronics Engineers, 1995.

[24] M.H.J. Bollen. Definitions of voltage unbalance. IEEE Power Engineering Re-

view, 22(11):49–50, November 2002.

[25] Math H.J. Bollen and Irene Gu. Signal Processing of Power Quality Disturbances.

John Wiley and Sons, 2006.

[26] Tsai-Hsiang Chen. Criteria to estimate the voltage unbalances due to high-speed

railway demands. IEEE Trans. on Power Systems, 9(3):1672–1678, August 1994.


197

[27] Tsai-Hsiang Chen and Hung-Yuan Kuo. Analysis on the voltage unbalances

due to high-speed railway demands. In Proc. International Conference on En-

ergy Management and Power Delivery (EMPD ’95), volume 2, pages 657–661,

November 1995.

[28] M. Cao, P.P. Biringer, and V. Bulat. Eddy current losses and unbalance in

electrode arms. IEEE Trans. on Magnetics, 24(6):2670–2672, November 1988.

[29] Eric T.B. Gross and S.W. Nelson. Electromagnetic unbalance of untransposed

transmission lines. AIEE Trans. on Power Apparatus and Systems, 74(3):887–

893, 1955.

[30] R.H. Brierley, A.S. Morched, and T.E. Grainger. Compact right-of-ways with

multi-voltage towers. IEEE Trans. on Power Delivery, 6(4):1682–1689, October

1991.

[31] Z. Emin and D.S. Crisford. Negative phase-sequence voltages on E&W trans-

mission system. IEEE Trans. on Power Delivery, 21(3):1607–1612, July 2006.

[32] A. Ametani, D. Van Dommelen, and I. Utsumi. Study of super-bundle and low

reactance phasings on untransposed twin-circuit lines. In IEE Proc., volume

137-4, pages 245–253, July 1990.

[33] W.R. Bullard, H.L. Lowe, and H.W. Wahlquist. Calculation of unbalanced volt-

age in distribution circuits with particular reference to multi-grounded neutrals.

AIEE Trans. on Electrical Engineering, 63:145–148, 1944.

[34] Paulo Vinicius Santos Valois, Carlos Mhrcio Vieira Than, Nelson Kagan, and

Hector Arango. Voltage unbalance in low voltage distribution networks. In Proc.

International Conference on Electricity Distribution (CIRED2001), June 2001.


198

[35] E.T.B. Gross and Wing Chin. Electrostatic unbalance of untransposed single cir-

cuit lines. AIEE Trans. on Power Apparatus and Systems, 87(1):24–34, January

1968.

[36] R. Gutman and L.B. Wagenaar. EHV transformer bank unbalance: practical

issues and solutions. IEEE Trans. on Power Delivery, 11(4):1830–1835, October

1996.

[37] J. C. Neupauer. Unbalanced open-wye open-delta transformer banks. AIEE

Trans. on Power Apparatus and Systems, 75(3):570–572, January 1956.

[38] J.W. Williams. Operation of 3 phase induction motors on unbalanced voltages.

AIEE Trans. on Power Apparatus and Systems, 73:125–133, April 1954.

[39] Gafford W.C. Duesterhoeft and C.C. Mosher. Heating of induction motors on

unbalanced voltages. AIEE Trans. on Power Apparatus and Systems, 78:282–

297, June 1959.

[40] C.Y. Lee. Effects of unbalanced voltage on the operation performance of a three-

phase induction motor. IEEE Trans. on Energy Conversion, 14(2):202–208, June

1999.

[41] AS 1359.31: Rotating electrical machines - general requirements - three-phase in-

duction motors - operation on unbalanced voltages. Technical report, Standards

Australia, 1997.

[42] D.P. Manjure and E.B. Makram. Impact of unbalance on power system harmon-

ics. In Proc. 10th International Conference on Harmonics and Quality of Power

(ICHQP 2002), volume 1, pages 328–333, 2002.

[43] A. von Jouanne and B. Banerjee. Assessment of voltage unbalance. IEEE Trans.

on Power Delivery, 16(4):782–790, October 2001.


199

[44] Math H.J. Bollen. Understanding Power Quality Problems - Voltage Sags and

Interruptions. Institute of Electrical and Electronics Engineers, 2000.

[45] L. Moran, P.D. Ziogas, and G. Joos. Design aspects of synchronous PWM

rectifier-inverter systems under unbalanced input voltage conditions. IEEE

Trans. on Industry Applications, 28(6):1286–1293, November-December 1992.

[46] V.O. Zambrano, E.B. Makram, and R.G. Harley. Stability of a synchronous ma-

chine due to an unsymmetrical fault in unbalanced power systems. In Proc. 20th

Southeastern Symposium on System Theory, pages 231–235, March 1988.

[47] E.B. Makram, V.O. Zambrano, and R.G. Harley. Stability of a synchronous

machine due to multiple faults in unbalanced power systems. In Proc. 20th

Southeastern Symposium on System Theory, pages 226–230, March 1988.

[48] Edited by Angelo Baggini. Handbook of Power Quality. John Wiley and Sons,

2008.

[49] AS 1359.101: Rotating electrical machines - general requirements - rating and

performance. Technical report, Standards Australia, 1997.

[50] T.H. Chen. Evaluation of line loss under load unbalance using the complex un-

balance factor. In IEE Proc. Generation, Transmission and Distribution, volume

142-2, pages 173–178, March 1995.

[51] J. Kuang and S.A. Boggs. Pipe-type cable losses for balanced and unbalanced

currents. IEEE Trans. on Power Delivery, 17(2):313–317, April 2002.

[52] A.E. Emanuel. On the definition of power factor and apparent power in unbal-

anced polyphase circuits with sinusoidal voltage and currents. IEEE Trans. on

Power Delivery, 8(3):841–852, July 1993.


200

[53] L.S. Czarnecki. Power related phenomena in three-phase unbalanced systems.

IEEE Trans. on Power Delivery, 10(3):1168–1176, July 1995.

[54] J.S. Wu, K.L. Tomsovic, and C.S. Chen. A heuristic search approach to feeder

switching operations for overload, faults, unbalanced flow and maintenance.

IEEE Trans. on Power Delivery, 6(3):1579–1586, October 1995.

[55] M.W. Siti, D.V. Nicolae, A.A. Jimoh, and A. Ukil. Reconfiguration and load

balancing in the LV and MV distribution networks for optimal performance.

IEEE Trans. on Power Delivery, 22(4):2534–2540, October 2007.

[56] Tsai-Hsiang Chen and Jeng-Tyan Cherng. Optimal phase arrangement of dis-

tribution transformers connected to a primary feeder for system unbalance im-

provement and loss reduction using a genetic algorithm. IEEE Trans. on Power

Delivery, 15(3):994–1000, August 2000.

[57] Paul M. Anderson. Analysis of Faulted Power Systems. John Wiley and Sons,

1995.

[58] G.V. Moodley, D. Dama, and R. Vajeth. Consideration of electromagnetic induc-

tion during transposition studies. In Proc. 7th AFRICON Conference (AFRICON

2004), volume 2, pages 1065–1070, Africa, 2004.

[59] Jen-Hung Chen, Wei-Jen Lee, and Mo-Shing Chen. Using a static Var compen-

sator to balance a distribution system. IEEE Trans. on Industry Applications,

35(2):298–304, March-April 1999.

[60] Kuang Li, Jinjun Liu, Biao Wei, and Zhaoan Wang. Comparison of two control

approaches of static Var generators for compensating source voltage unbalance

in three-phase three-wire systems. In Proc. 4th International Conference on


201

Power Electronics and Motion Control (IPEMC 2004), volume 3, pages 1213–

1218, August 2004.

[61] Tong XiangQian, Xi Keqing, Shen Ming, and Ma Xianhong. Reactive power

and unbalance compensation with DSTATCOM. In Proc. 8th International

Conference on Electrical Machines and Systems (ICEMS 2005), volume 2, pages

1181–1184, September 2005.

[62] A. Campos, G. Joos, P.D. Ziogas, and J.F. Lindsay. Analysis and design of a

series voltage unbalance compensator based on a three-phase VSI operating with

unbalanced switching functions. IEEE Trans. on Power Electronics, 9(3):269–

274, May 1994.

[63] A. Campos, G. Joos, P. Ziogas, and J. Lindsay. Analysis and design of a series

voltage compensator for three-phase unbalanced sources. IEEE Trans. on Power

Electronics, 39(2):159–167, April 1992.

[64] V.B. Bhavaraju and P.N. Enjeti. An active line conditioner to balance voltages

in a three-phase system. IEEE Trans. on Industry Applications, 32(2):287–292,

March-April 1992.

[65] Hong-Ju Jung, In-Young Suh, Byung-Seob Kim, Rae-Young Kim, See-Young

Choi, and Jong-Hwhan Song. A study on DVR control for unbalanced voltage

compensation. In Proc. 17th Annual IEEE Conference and Exposition on Applied

Power Electronics (APEC 2002), volume 2, pages 1068–1073, March 2002.

[66] Gong Maozhong, Liu Hankui, Gu Hanjun, and Xu Dianguo. Active voltage

regulator based on novel synchronization method for unbalance and fluctuation

compensation. In Proc. 28th Annual IEEE Conference of the Industrial Electron-

ics Society (IECON 02), volume 2, pages 1374–1379, November 2002.


202

[67] D. Graovac, V. Katic, and A. Rufer. Power quality problems compensation with

universal power quality conditioning system. IEEE Trans. on Power Delivery,

22(2):968–976, April 2007.

[68] J. G. Pinto, R. Pregitzer, Luis. F. C. Monteiro, Carlos Couto, and Joao. L.

Afonso. A combined series active filter and passive filters for harmonics, un-

balances and flicker compensation. In Proc. International Conference on Power

Engineering, Energy and Electrical Drives (POWERENG 2007), volume 2, pages

54–59, April 2007.

[69] A. Moreno-Munoz. Power Quality - Mitigation Technologies in a Distributed

Environment. Springer.

[70] IEC 61000-2-2: Electromagnetic compatibility (EMC) - part 2.2 - environment

- compatibility levels for low-frequency conducted disturbances and signalling in

public low-voltage power supply systems, Ed. 2. Technical report, International

Electrotechnical Commission, 2002.

[71] IEC 61000-2-12: Electromagnetic compatibility (EMC) - part 2.2 - environment

- compatibility levels for low-frequency conducted disturbances and signalling in

public medium-voltage power supply systems, Ed. 1. Technical report, Interna-

tional Electrotechnical Commission, 2003.

[72] A. Robert and J. Marquet (on behalf of Working Group 36.05). Assessing volt-

age quality with relation to harmonics, flicker and unbalance. In Proc. CIGRE

conference, paper 36-203, Paris, August-September 1992.

[73] Electricity distribution code, Victoria. Technical report, January 2006.

[74] Robert Koch, Germain Beaulieu, Luc Berthet, and Mark Halpin. Interna-

tional survey of unbalance levels in LV, MV, HV and EHV power systems:
203

CIGRE/CIRED JWG C4.103 results. In Proc. 19th International Conference

on Electricity Distribution (CIRED 2007), paper 0892, Vienna, May 2007.

[75] AS/NZS 61000.3.6: Electromagnetic compatibility (EMC) - limits - assessment

of emission limits for distorting loads in MV and HV power systems. Technical

report, Standards Australia/New Zealand, 2001.

[76] AS/NZS 61000.3.7: Electromagnetic compatibility (EMC) - limits - assessment

of emission limits for fluctuating loads in MV and HV power systems. Technical

report, Standards Australia/New Zealand, 2001.

[77] J. Duncan Glover and Gareth Digby. Software Manual - Power System Analysis

and Design. PWS Publishing Company, 1994.

[78] R.C. Dugan. Induction machine modeling for distribution system analysis -

test case description. In Proc. Transmission and Distribution Conference and

Exhibition 2005/2006 IEEE PES, pages 578–582, 21-24 May 2006.

[79] Paul C. Krause, Oleg Wasynczuk, and Scott D. Sudhoff. Analysis of Electric

Machinery and Drive Systems. John Wiley and Sons Inc., 2002.

[80] R.C. Dugan, M.F. McGranaghan, and H.W. Beaty. Electrical Power Systems

Quality. New York: McGraw-Hill, 1996.


Appendix A

Derivation of (3.5)

Equation (3.4) in relation to the considered radial MV-LV network shown in Fig. 3.2

is replicated here by (A.1):

t
|V−:g/mv:rec | = |Z−+:t I+:t + Z++:t I−:t/t | (A.1)

When the loads Smv:rec−local and Smv:rec−ds represent constant impedance loads, the

currents I+:t and I−:t/t can be written1 respectively as:

V+:send
I+:t = (A.2)
Z++:send
I−:t/t = Y−+:send V+:send (A.3)

where,

V+:send - positive sequence voltage at the sending end busbar (labelled ‘send’)

Z++:send , Y−+:send - as defined below when xr = send

1
Note the voltage at the sending end busbar is balanced.
204
205

The downstream negative-positive sequence coupling admittance Y−+:xr seen at a

busbar xr of a radial network can be generally expressed2 using various impedance

elements seen at the busbar as:

−Z−+:xr
Y−+:xr ≈ (A.4)
Z++:xr Z−−:xr

Alternatively, the downstream negative-positive sequence coupling impedance Z−+:xr

seen at the busbar xr can be expressed using various admittance elements seen at the

busbar as:
−Y−+:xr
Z−+:xr ≈ (A.5)
Y++:xr Y−−:xr

where,

Z++:xr , Z−−:xr - downstream positive and negative sequence impedances respectively

seen at the busbar xr

Y++:xr , Y−−:xr - downstream positive and negative sequence admittances respectively

seen at the busbar xr

For the considered system where the loads are balanced (i.e. decoupled sequence

impedances) and the MV line is untransposed: Z−+:send = Z−+:t . Thus, the admit-

tance Y−+:send can be given by:

−Z−+:t
Y−+:send ≈ (A.6)
Z++:send Z−−:send

Using (A.2), (A.3) and (A.6), (A.1) can be developed as:

 
t
Z++:t
|V−:g/mv:rec | ≈ Z−+:t I+:t 1 −
(A.7)
Z−−:send
2
Through the inversion of the impedance matrix and the simplification.
206

Noting that Z−−:send = Z−−:rec + Z++:t , (A.8) can be rewritten as:

 
t
Z −−:rec
|V−:g/mv:rec | ≈ Z−+:t I+:t (A.8)
Z−−:send
Appendix B

Radial MV-LV Test System

(Fig. 3.2)

This test system is designed mainly based on the data given in [78].

• System details - 60Hz, three-wire, MV - 12.47kV , LV - 460V

• MV line -

– Line details: 3.2187km, untransposed

– Tower construction details: 1.143m flat and horizontal,

phase positioning1 : a b c

– Conductor data:

geometric mean radius = 7.7724mm

resistance = 0.19014Ω/km

earth resistivity = 100Ωm

1
This shows the considered arrangement of the three phase conductors (a, b and c) of the hori-
zontal tower.
207
208

– Relevant impedance/admittance data:

positive sequence impedance (Ω/km) = 0.1901 + j0.3937 (0.4372∠640 )

negative-positive sequence coupling impedance (Ω/km) = 0.0302+j0.0174

(0.0349∠300 )

positive sequence admittance (Skm) = 1.0098 − j2.0630 (2.2969∠−640 )

negative-positive sequence coupling admittance (Skm) = 0.0258 + j0.1821

(0.1839∠820 )

• MV-LV coupling transformer - aggregated representation of fully loaded 1MVA,

12.47kV/460V transformers with winding resistance = 1%, leakage reactance

= 5%, and secondary tap setting = 1.03pu. This gives a value of 19 for the

factor ksc−lvagg .

• Induction motor load - aggregated representation of 50hp, 460V, 1705rpm mo-

tors [15] supplying the rated mechanical load.

– Parameters per motor:

stator resistance = 0.087Ω

rotor resistance (referred to stator side) = 0.228Ω

stator, rotor (referred to stator side) leakage reactance = 0.302Ω

magnetising reactance = 13.08Ω

– Other relevant details: input power = 50kVA at 0.9 lagging pf and ks = 6.7

under rated operating conditions

• Passive loads - aggregated representation of passive components with a rated

power of 50kVA at 0.9 lagging pf which consist of equal shares constant power

and constant impedance elements.

• Sending end positive sequence voltage = 7.78∠00 kV (1.08pu)


Appendix C

Derivation of (3.14)

Equation (3.13) in relation to the considered radial MV-LV network shown in Fig. 3.2

is replicated here by (C.1):

 
t
Z−−:rec−im
|V−:g/mv:rec | ≈ Z−+:t I+:t
(C.1)
Z−−:send−im

Noting Z−−:send−im = Z−−:rec−im + Z++:t and the inductive nature of the associated

impedance elements1 , (C.1) can be rewritten as:

 
t 1
|V−:g/mv:rec | ≈ |Z−+:t I+:t | 
Z++:t Z++:rec
 (C.2)
1 + Z++:rec Z−−:rec−im

where, Z++:rec - downstream equivalent positive sequence impedance seen at the re-

ceiving end busbar of the MV line. The ratio ZZ++:rec
++:t
can be expressed as:


Z++:t V Rt
Z++:rec ≈ 1 − V Rt (C.3)

1
e.g. the impedance Z−−:rec−im is made up of the negative sequence impedances of induction
motors and the MV-LV transformer.

209
210

Substituting Z−−:rec−im = (nml )2 (Z−−:im + Z++:ml−lv ) and rearranging, the ratio



Z++:rec
Z−−:rec−im can be written as:

! !
Z++:rec 1 1
Z−−:rec−im = (C.4)

(nml )2 |Z−−:im +Z++:ml−lv | |V+:rec |2
|V+:rec |2 |Z++:rec |

where,

nml - operating turns ratio of the MV-LV coupling transformer

Z−−:im - negative sequence impedance of the aggregated motor load

|V+:rec |2
The term |Z++:rec |
represents the total MVA load (i.e. Smv:rec−local + Smv:rec−ds ) sup-

plied by the MV busbar. Using the relationships |V+:rec | = nml |V+:lv | and |Z−−:im +

Z++:ml−lv | ≈ |Z−−:im | + |Z++:ml−lv |, (C.4) can be further rearranged as:

 

Z++:rec 1 1
Z−−:rec−im ≈ (Smv:rec−local + Smv:rec−ds ) Z−−:im Z++:im |Z++:lv |
 
|Z++:ml−lv |
Z++:im Z++:lv |V+:lv |2 + |V+:lv |2
(C.5)

where,

V+:lv - positive sequence voltage at the LV busbar

Z++:im - positive sequence impedance of the aggregated motor load

Z++:lv - equivalent downstream positive sequence impedance seen at the LV busbar

Noting that:

Z−−:im 1 Z++:im 1 |Z++:lv | 1
Z++:im = ks
, Z++:lv = km
, |V+:lv |2
= Smv:rec−ds

|Z++:ml−lv | 1 Smv:rec−ds Ssc−lvagg


|V+:lv |2
= Ssc−lvagg
, Smv:rec−local +Smv:rec−ds
= klv , Smv:rec−ds
= ksc−lvagg
211

(C.5) can be written as:


Z++:rec klv
Z−−:rec−im ≈ (C.6)

1 1
k s km
+ ksc−lv
agg

Substitution of (C.3) and (C.6) in (C.2) gives:

t 1
|V−:g/mv:rec | ≈ |Z−+:t I+:t |    (C.7)
V Rt klv
1+ 1−V Rt 1
+k 1
ks km sc−lvagg
Appendix D

Y−−:x−im for an MV Network

The admittance Y−−:x−im can be generally written noting the inductive nature of the

respective impedance Z−−:x−im as:


Z++:x
Y−−:x−im ≈ −j |Y++:x |
(D.1)
Z−−:x−im

where, Y++:x , Z++:x - downstream equivalent positive sequence admittance and imped-

ance respectively seen at the busbar x. Noting that the system representation of the

MV busbar x shown in Fig. 3.6 and that of the receiving end busbar of the MV line

considered in Fig. 3.2 are similar, the ratio ZZ−−:x
++:x
can be given according to (C.6) as:


Z++:x klv:x
Z−−:x ≈ (D.2)

1
ks:x km:x
+ ksc−lv1
agg :x

The absolute value of the admittance Y++:x can be expressed in terms of the system

voltage and the nodal current as:


3 |I+:x |
|Y++:x | ≈ (D.3)
Vn−mv

212
213

Substitution of (D.2) and (D.3) in (D.1) gives:


! √ !
klv:x 3 |I+:x |
Y−−:x−im ≈ −j 1 (D.4)
ksc−lvagg :x
+ ks:x1km:x Vn−mv
Appendix E

Application of the Methodology

Given by (3.25) to the Three-bus

MV Test System (Fig. 3.7)

The matrix equation (3.25) which gives the proposed methodology for MV networks

is replicated here by (E.1):

0
lines
[V−:g/mv ]n×1 ≈ −[Y++ ]−1
n×n [Y−+ ]n×n [V+ ]n×1 (E.1)

where,
0
Y++:xy ≈ Y++:xy + Y−−:x−im for x = y
0
Y++:xy = Y++:xy for x 6= y

214
215

The matrix [V+ ]1 for the test system is:

 
0
 1.05∠−4.07   
  12.47
 0.98∠−6.74  ×
[V+ ] =  0  √ kV (E.2)
  3
1.01∠−5.570

Noting that the positive sequence admittance per km of the lines = 1.0098−j2.0630S

and also that the impedance Y++:11 has to be incorporated with the HV-MV trans-

former impedance, the matrix [Y++ ] is:

 
 0.3722 − j1.3119 −0.1010 + j0.2063 −0.2020 + j0.4126 
 
[Y++ ] = 
 −0.1010 + j0.2063 0.2020 − j0.4126 −0.1010 + j0.2063
S
 (E.3)
 
−0.2020 + j0.4126 −0.1010 + j0.2063 0.3029 − j0.6189

Noting that the negative-positive sequence coupling admittance per km of the lines

= 0.0258 + j0.1821S, the matrix [Y−+ ] is:

 
 0.0078 + j0.0546 −0.0026 − j0.0182 −0.0052 − j0.0364 
 
[Y−+ ] = 
 −0.0026 − j0.0182 0.0052 + j0.0364 −0.0026 − j0.0182
S
 (E.4)
 
−0.0052 − j0.0364 −0.0026 − j0.0182 0.0078 + j0.0546

As busbars 1 and 3 supply passive loads (i.e. km:1 = 0, km:3 = 0), the admittances

Y−−:1−im ≈ 0 and Y−−:3−im ≈ 0.

• Case 1 - busbar 2 supplies passive loads or km:2 = 0. That is, the admittance
0
Y−−:2−im ≈ 0 implying that [Y++ ] ≈ [Y++ ]. Then, the substitution of (E.2),

(E.3) and (E.4) in (E.1) gives:


1
This is obtained using load flow analysis.
216

 
 0 
lines
 
[V−:g/mv ]=
 47.38 V
 (E.5)
 
26.64

Expressing (E.5) as VUFs:

 
 0 
lines
 
[Ug/mv ]=
 0.67 %
 (E.6)
 
0.37

• Case 2 - busbar 2 supplies motor loads at the LV level or km:2 = 1. Equation

(3.24) gives the admittance Y−−:2−im for klv:2 = 1, ksc−lvagg :2 = 18, ks:2 = 6.7,

I+:2 = 194A and Vn−mv = 12.47kV as:

Y−−:2−im ≈ −j0.1316S (E.7)

0
Then, the matrix [Y++ ] can be obtained as:

 
 0 0 0 
0
 
[Y++ ] ≈ [Y++ ] + 
 0 −j0.1316 0 S
 (E.8)
 
0 0 0

Substitution of (E.2), (E.3) and (E.8) in (E.1) gives:

 
 5.94 
lines
 
[V−:g/mv ]=
 31.46  V
 (E.9)
 
17.38
217

Expressing (J.11) as VUFs:

 
 0.08 
lines
 
[Ug/mv ]=
 0.45  %
 (E.10)
 
0.24
Appendix F

Derivation of (4.7)

t
Equation (4.5) which gives the negative sequence voltage V−:g/hv:rec caused by the

local HV line t of the considered radial HV-MV-LV network shown in Fig. 4.2 is

replicated here by (F.1):

t
V−:g/hv:rec = −(Z−+:t I+:t + Z++:t I−:t/t ) (F.1)

Similar to (A.8) in the case of the radial MV-LV network considered in Fig. 3.2, (F.1)

can be written noting that the local HV line is the element which gives rise to the

negative-positive sequence coupling impedance seen at the sending end busbar of the

HV line as1 :
 
t Z++:t
V−:g/hv:rec = −Z−+:t I+:t 1 − (F.2)
Z−−:send−im
Z++:t
Due to the inductive nature of the impedances Z++:t and Z−−:send−IM , Z−−:send−im


Z++:t
Z−−:send−im = µ. Thus, (F.2) can be rewritten as:

t
V−:g/hv:rec = −Z−+:t I+:t (1 − µ) (F.3)
1
The downstream MV line is treated as balanced in this case.

218
219
td
Equation (4.6) which gives the negative sequence voltage V−:g/hv:rec caused by the

downstream MV line td is replicated here by (F.4):

td
V−:g/hv:rec = −Z++:t I−:td /t (F.4)

The negative sequence current I−:td /t = Y−+:send V+:send where, according to (A.4),
−Z−+:send 2
Y−+:send = Z++:send Z−−:send−im
. Noting that the impedance Z−+:send arises as a result

of the downstream MV line in this case3 , it is not simply equal to the impedance

Z−+:td . Employing (A.4) and (A.5), Z−+:send can be established as4 :

Z−+:send ≈ (nhm )2 ktd kn Z−+:td (F.5)

where, nhm - operating turns ratio of the HV-MV coupling transformer. Then, (F.4)

can be written as:

  
td 2 Z++:t V+:send
V−:g/hv:rec ≈ (nhm ) ktd kn Z−+:td (F.6)
Z−−:send−im Z++:send

Noting that:

Z++:t V+:send (nhm )2 Z−+:td


Z−−:send−im
= µ, Z++:send
= I+:t , Z−+:t

(F.6) can be rearranged as:

td
V−:g/hv:rec ≈ (µktd kn σ)Z−+:t I+:t = ζZ−+:t I+:t (F.7)

2
New notations which have been used here are as defined in Appendix A while using the subscript
‘send’ with reference to the sending end busbar of the local HV line.
3
The local HV line is treated as balanced in this case.
4
A uniform pf across the network is assumed.
220
t+td
Summation of (F.3) and (F.7) gives the negative sequence voltage V−:g/hv:rec which

arises as a result of both the HV and MV lines as:

t+td
|V−:g/hv:rec | ≈ |Z−+:t I+:t (1 − µ − ζ)| (F.8)
Appendix G

Derivation of (4.9)

New notations used in this appendix, unless defined here or in Chapter 4 (Section 4.2),

are as per Appendix C while using the subscripts ‘send’ and ‘rec’ with reference to the

sending and receiving end busbars respectively of the local HV line of the considered

HV-MV-LV radial network shown in Fig. 4.2.

The factor µ is defined as:


Z++:t
µ = (G.1)
Z−−:send−im

As Z−−:send−im = Z−−:rec−im + Z++:t , and the associated impedances are inductive in

nature, (G.1) can be written as:

1
µ≈
Z−−:rec−im Z++:rec (G.2)
1 + Z++:rec Z++:t


The ratio ZZ++:rec can be expressed in terms of the voltage regulation of the line t as:

++:t


Z++:rec 1 − V Rt
Z++:t ≈ V Rt (G.3)

221
222

Noting that Z−−:rec−im = (nml nhm )2 (Z−−:im +Z++:mlr −lv +Z++:td −lv +Z++:hm−lv ) when

motor loads are supplied only by the MV line at the busbar LVr , and the inductive

Z−−:rec−im
nature of the associated impedances, the ratio Z++:rec can be given as:

!
2
 
≈ |V+:rec | |Z−−:im | |Z++:mlr −lv + Z++:td −lv + Z++:hm−lv |
Z−−:rec−im

|V+:rec |2
+ |V+:rec |2
Z++:rec |Z++:rec |
(nml nhm )2 (nml nhm )2
(G.4)

|V+:rec |2
The term |Z++:rec |
represents the total load Srec (in MVA) supplied by the HV busbar
nml nhm |V+:lv |
rec. Further, |V+:rec | ≈ (1−V Rtd )
. Substitution of these and rearrangement of (G.4)

gives:

 
≈ Srec (1 − V Rt )2 |Z−−:im | + |Z++:mlr −lv + Z++:td −lv + Z++:hm−lv |
Z−−:rec−im

d
Z++:rec |V+:lv |2 |V+:lv |2
(G.5)

Noting that:
  
|Z−−:im | 1 1
|V+:lv | 2 = ksr kmr Slvr
(see (C.5) - (C.6))
agg

|Z++:mlr −lv +Z++:td −lv +Z++:hm−lv | 1


|V+:lv |2
= Ssc−lvragg

Srec 1
Slvragg
= klvr

Slvragg 1
Ssc−lvragg
= ksc−lvragg

(G.5) can be given as:

Z−−:rec−im (1 − V Rtd )2
 
≈ 1 1

Z++:rec + (G.6)
klvr ksr kmr ksc−lvragg
223

Substitution of (G.3) and (G.6) in (G.2) gives:

1
µr ≈ „ « (G.7)
1 1
(1−V Rt )(1−V Rtd )2 ksr kmr
+k
sc−lvragg
1+ V Rt klvr
Appendix H

Test Case Description of the Radial

HV-MV-LV System (Fig. 4.2)

• System details - 60Hz, three-wire, HV - 66kV , MV - 12.47kV , LV - 460V

• System loads -

– Shv = 0

– Smvsagg - represents passive loads

– Smvragg = 0 (i.e. ktd = klvr )

– Slvsagg = 0 (i.e. kn = 1 for the case where kmr = 1)

– Slvragg - represents passive loads and induction motors for the cases where

kmr = 0 and kmr = 1 respectively

• HV line -

– Line details: 15km, untransposed

– Tower construction details: 1.44m flat and horizontal,

phase positioning1 : a b c
1
This shows the considered arrangement of the phases a, b and c of the horizontal tower.
224
225

– Conductor data:

geometric mean radius = 9mm

resistance = 0.105Ω/km

earth resistivity = 100Ωm

– Relevant impedance/admittance data:

positive sequence impedance (Ω/km) = 0.1050 + j0.4001 (0.4136∠750 )

negative-positive sequence coupling impedance (Ω/km) = 0.0302+j0.0174

(0.0349∠300 )

positive sequence admittance (Skm) = 0.6265 − j2.3517 (2.4337∠−750 )

negative-positive sequence coupling admittance (Skm) = 0.1040 + j0.1779

(0.2061∠600 )

• HV-MV coupling transformer - aggregated representation of equally loaded

(10MVA per transformer under the considered operating scenario) 12MVA,

66kV/12.47kV transformers with winding resistance and leakage reactance of

1% and 10% respectively [78].

• MV line - aggregated representation of equally loaded (10MVA per line under

the considered operating scenario) 12.47kV, 3.2187km lines of which the details

are given in Appendix B.

• MV-LV coupling transformer - aggregated representation of fully loaded 1MVA,

12.47kV/460V transformers with winding resistance and leakage reactance of

1% and 5% respectively [78].

• Induction motor and passive loads - as described in Appendix B.


226

• Considered operating scenario -

– Positive sequence voltage at the sending end busbar of the HV line =

41.916kV ∠00 (1.1pu)

– Total load supplied by the HV line = 50MVA at 0.9 lagging pf

– Secondary tap setting of the HV-MV coupling transformer = 1.1pu

– Secondary tap setting of the MV-LV coupling transformer = 1.05pu

– Resulting system conditions2 :

I+:t ≈ 490A , V Rt ≈ 7% , V Rtd ≈ 9% , |V+:rec | ≈ 39.6kV (1.04pu)

• Number of 12.47kV, 3.2187km lines and the values of the factors ksc−lvragg and

σ corresponding to various klvr are given in Table H.1 below:

Table H.1: Values of ksc−lvragg and σ for various klvr


klvr No. of 12.47kV , ksc−lvragg σ
3.2187km lines
0.0 0 N/A N/A
0.2 1 5.7 5.0 ∠00
0.4 2 5.0 2.5 ∠00
0.6 3 4.5 1.7 ∠00
0.8 4 4.1 1.3 ∠00
1.0 5 3.8 1.0 ∠00

2
Which are obtained using load flow analysis.
Appendix I

Y−+:x for an HV Network

According to (A.4), the admittance Y−+:x which arises as a result of line asymmetries

that exist in the downstream system supplied by the busbar x can be written as1 :

−Z−+:x
Y−+:x ≈ ≈ j Z−+:x Y++:x |Y−−:x−im | (I.1)
Z++:x Z−−:x−im

Noting that:

• the impedance Z−+:send given by (F.5) in the case where the untransposed MV

line gives rise to unbalance in the considered radial HV-MV-LV network shown

in Fig. 4.2, is equal to the respective impedance Z−+:rec that is seen at the

receiving end busbar of the HV line, as the local HV line is treated as balanced

in this case, and

• the system representation of any HV busbar x shown in Fig. 4.5 and that of the

receiving end busbar of the radial HV-MV-LV network considered in Fig. 4.2

are similar,
1
Note that the impedance Z−−:x−im is inductive.

227
228

the impedance Z−+:x can be expressed as:

Z−+:x ≈ (nhm:x )2 ktd:x kn:x Z−+:td:x (I.2)

where, nhm:x - operating turns ratio of the HV-MV coupling transformer supplied by

the busbar x. Similar to (D.3), the admittance Y++:x can be expressed as:

√ !
3 |I+:x |
Y++:x ≈ ∠θpf :x (I.3)
Vn−hv

Substitution of (I.2) and (I.3) in (I.1) gives:

√ !
3 |I+:x |
|Y−+:x | ≈ ktd:x kn:x |Y−−:x−im Z−+:td:x −hv | (I.4)
Vn−hv
θY−+:x ≈ 900 + θZ−+:td:x + θpf :x (I.5)
Appendix J

Application of the Methodology

Given by (3.22) to the Three-bus

HV Test System (Fig. 4.6)

The matrix equation (3.22) which gives the proposed methodology for HV networks

is replicated here by (J.1):

0
lines
[V−:g/hv ]n×1 ≈ −[Y++ ]−1
n×n [Y−+ ]n×n [V+ ]n×1 (J.1)

where,
0
Y++:xy ≈ Y++:xy + Y−−:x−im for x = y
0
Y−+:xy ≈ Y−+:xy + Y−+:x for x = y
0 xy
Y++:xy = Y++ for x 6= y
0 xy
Y−+:xy = Y−+ for x 6= y

229
230

The matrices [V+ ]1 , [Y++ ] and [Y−+ ] for the three-bus HV test system shown in

Fig. 4.6 are:  


0
 1.07∠−8.3   
  66
 × √ kV
[V+ ] = 
 1.01∠−10.7
0 (J.2)


 3
1.04∠−9.60
 
 0.0467 − j0.2214 −0.0125 + j0.0470 −0.0313 + j0.1176 
 
[Y++ ] = 
 −0.0125 + j0.0470 0.0282 − j0.1058 −0.0157 + j0.0588
S
 (J.3)
 
−0.0313 + j0.1176 −0.0157 + j0.0588 0.0470 − j0.1764
 
 0.0073 + j0.0124 −0.0021 − j0.0036 −0.0052 − j0.0089 
 
[Y−+ ] = 
 −0.0021 − j0.0036 0.0047 + j0.0080 −0.0026 − j0.0044
S
 (J.4)
 
−0.0052 − j0.0089 −0.0026 − j0.0044 0.0078 + j0.0133

As busbars 1 and 3 supply passive loads (i.e. km:1 = 0, km:3 = 0), the admittances

Y−−:x−im ≈ 0 and Y−+:x ≈ 0 for x = 1, 3.

• Case 1 - busbar 2 supplies passive loads or km:2 = 0. That is, the admittances
0 0
Y−−:2−im ≈ 0 and Y−+:2 ≈ 0 implying that [Y++ ] ≈ [Y++ ] and [Y−+ ] ≈ [Y−+ ]

respectively. Then, the substitution of (J.2), (J.3) and (J.4) in (J.1) gives:

 
 0 
lines
 
[V−:g/hv ]=
 226.89 V
 (J.5)
 
119.02
1
This is obtained using load flow analysis.
231

Expressing (J.5) as VUFs:

 
 0 
lines
 
[Ug/hv ]=
 0.59  %
 (J.6)
 
0.30

• Case 2 - busbar 2 supplies motor loads at the LV level or km:2 = 1. Equation

(4.13) gives the admittance Y−−:2−im for klvr:2 = 1, ksc−lvragg :2 = 3.6, ksr:2 = 6.7,

I+:2 = 208A, V Rtd:2 = 9% and Vn−hv = 66kV as:

Y−−:2−im ≈ −j0.0154S (J.7)

Equation (4.14) gives the admittance Y−+:2 for ktd:2 = 1, kn:x = 1, I+:2 = 208A,

Vn−hv = 66kV , Y−−:2−im ≈ −j0.0154S, Z−+:td:x −hv = 1.3∠300 Ω, and θpf :x =

−260 as:

Y−+:2 ≈ (−0.0077 + j0.1093) × 10−3 S (J.8)

0 0
Then, the matrices [Y++ ] and [Y−+ ] can be obtained as:

 
 0 0 0 
0
 
[Y++ ] ≈ [Y++ ] + 
 0 −j0.0154 0 S
 (J.9)
 
0 0 0
 
 0 0 0 
0  × 10−3 S
 
[Y−+ ] ≈ [Y−+ ] + 
 0 −0.0077 + j0.1093 0  (J.10)
 
0 0 0
232

Substitution of (J.2), (J.9) and (J.10) in (J.1) gives:

 
 94.19 
lines
 
[V−:g/hv ]=
 73.52  V
 (J.11)
 
5.51

Expressing (J.11) as VUFs:

 
 0.23 
lines
 
[Ug/hv ]=
 0.19 %
 (J.12)
 
0.01
Appendix K

Data of the IEEE 14-bus Test

System (Fig. 4.9)

Note that the impedance/admittance values given in pu are based on a 100M V A

base.

Table K.1: Voltage controlled bus data


Bus Voltage Minimum Maximum
number magnitude (pu) MVAr capability MVAr capability
2 1.045 40 50
3 1.010 0 40
6 1.070 6 24
8 1.090 6 24

Table K.2: Static capacitor data: susceptances


Bus Susceptance
number (pu)
9 0.19

233
234

Table K.3: Generator and load bus data: three-phase MW and MVAr values
Bus Generation Load
number MW MVAr MW MVAr
1 (reference bus) 0 0 0 0
2 40 0 21.7 12.7
3 0 0 94.2 19.0
4 0 0 47.8 3.9
5 0 0 7.6 1.6
6 0 0 11.2 7.5
7 0 0 0 0
8 0 0 0 0
9 0 0 29.5 16.6
10 0 0 9.0 5.8
11 0 0 3.5 1.8
12 0 0 6.1 1.6
13 0 0 13.5 5.8
14 0 0 14.9 5.0

Table K.4: Transformer data: impedances and secondary tap settings (1st and 2nd
bus numbers refer to the primary and the secondary respectively)
Transformer Impedance Secondary tap
number (pu) setting
4-7 j0.20912 1.022
4-9 j0.55618 1.032
5-6 j.25202 1.073
7-8 j0.17615 1
7-9 j0.11001 1
235

Table K.5: Nodal positive sequence voltages


Bus number Magnitude (pu) Phase angle (deg.)
1 1.0600 0
2 1.0450 −5.00
3 1.0100 −12.60
4 1.0137 −10.19
5 1.0158 −8.64
6 1.0700 −14.67
7 1.0605 −13.55
8 1.0900 −13.55
9 1.0558 −15.21
10 1.0517 −15.45
11 1.0579 −15.23
12 1.0579 −15.61
13 1.0541 −15.76
14 1.0409 −16.61
236

Table K.6: Transmission line data: lengths and impedances


Line Length Positive sequence
(km) impedance (pu)
1-2 6.56 0.0158 + j0.0602
1-5 24.17 0.0583 + j0.2220
2-3 21.43 0.0517 + j0.1968
2-4 19.55 0.0471 + j0.1796
2-5 19.27 0.0464 + j0.1770
3-4 19.34 0.0466 + j0.1777
4-5 4.65 0.0112 + j0.0427
6-11 23.21 0.0560 + j0.2132
6-12 29.89 0.0720 + j0.2745
6-13 15.39 0.0371 + j0.1413
9-10 9.51 0.0229 + j0.0873
9-14 31.46 0.0758 + j0.2890
10-11 22.00 0.0530 + j0.2020
12-13 31.37 0.0756 + j0.2882
13-14 40.83 0.0984 + j0.3750
Appendix L

Derivation of (5.18)

Consider a series of constant impedance loads L1 , L2 , ..., Ln (balanced: decoupled and

equal positive, negative and zero sequence impedances) is supplied by the LV sys-

tem shown in Fig. 5.2, where the busbars US and S represent MV and LV systems

respectively. For this, (5.5) can be written as:


V−:Umv /lv−mv 1
=   (L.1)
V−:mv 1
1 + Z++:ml−lv Z++:L

+ 1
+ ... + 1
1
Z++:L2 Z++:Ln

where, Z++:Li - positive sequence impedance of any load Li (i = 1, 2, ..., n). Employing

(5.3), (L.1) can be expressed in terms of the system and load characteristics and the

load composition as:


V−:Umv /lv−mv 1
= (L.2)
V−:mv k L k L kLn
1 + j ksc−lv ∠θpf :L1 + j ksc−lv ∠θpf :L2 + ... + j ksc−lv ∠θpf :Ln
1 2

where,

kLi - ratio between the load Li (in MVA) and the total load (in MVA) supplied by

the LV system

237
238

θpf :Li - power factor angle1 of the load Li

kLi kLj kLi 2 kLi kLj


Noting that (ksc−lv )2
 1, ksc−lv , (L.2) can be written, neglecting (ksc−lv )2
and higher

order terms3 , as:


V−:Umv /lv−mv 1

V−:mv k L k L k Ln
1 + j ksc−lv ∠θpf :L1 1 + j ksc−lv ∠θpf :L2 ... 1 + j ksc−lv ∠θpf :Ln
1 2

(L.3)

Alternatively, the negative sequence voltage |V−:Umv /lv−mv | can be written in an ex-

panded form by decomposing the negative sequence current I−:Uus /tf (I−:Uus /tf =

I−:L1 + I−:L2 + ... + I−:Ln ) for the above considered case as:

n
X
|V−:Umv /lv−mv | = V−:mv − (Z++:ml−mv I−:Li ) (L.4)


i=1

where, I−:Li - negative sequence current (referred to MV) in the load Li , which

arises as a result of the MV unbalance. Noting that the influence of the term

V
Z++:tf −mv I−:Umv /tf on the ratio −:UVmv /lv−mv
−:mv
has been replaced by the factor (referred
1
to as ‘replacement factor’) ˛ ˛ for an aggregated constant impedance
˛1+j 1 ∠θpf :lv ˛
˛ ˛
˛ ksc−lv ˛
load4 , the comparison of (L.3) and (L.4) indicates, for the series of loads, that the in-

V
fluence of each of the Z++:ml−mv I−:Li components on −:UVmv /lv−mv
−:mv
has been replaced
1
by the factor ˛
k
˛ which involves an additional factor kLi . This observa-
˛1+j Li ∠θpf :L ˛
˛ ˛
˛ ksc−lv i˛

tion suggests that the impact of a share I−:Li of the total negative sequence current

I−:Uus /tf (or of a share Z++:ml−mv I−:Li of the total Z++:ml−mv I−:Uus /tf ) with a unique

behaviour, which is determined by the load type and its characteristics, on the prop-

agation of the negative sequence voltage from higher voltage to lower voltage systems
1
− and + for lagging and leading conditions respectively.
2
As kLi , kLj < 1 and 10 < ksc−lv < 25.
3 k L kL
i j
kL k L
i j
i.e. (ksc−lv )3 , (ksc−lv )4 etc.
4
Refer to (5.5) and (5.6).
239

can be represented in the form of (L.3). Equation (L.3) is a product of a number of

terms where an individual term accounts for an unique I−:Li or a load element Li .

Generalising the above outcome, the influence of a series of load elements L1 ,

L2 ,..., Ln 5 on the propagation of the negative sequence voltage from higher voltage

to lower voltage systems can be represented as a product of a number of terms in the

form given by (L.3). An individual term of this product, which corresponds to any

load element Li , can be obtained by modifying the replacement factor for its load
 
1
type6 where ksc−lv is multiplied by the load proportion kLi . Applying this to a mix

of constant impedance (Z), constant current (I), constant power (P Q) and induction

motor (IM ) loads (i.e. Li = z, i, pq, im), the individual terms to form the product

can be obtained as given in Table L.1.

Table L.1: Replacement factors for a mix of various load types


Load type Replacement factor

1
Z ˛ ˛
˛1+j kz ∠θpf :z ˛
˛ ˛
˛ ksc−lv ˛

I 1

1
PQ ˛ ˛β
˛1+j kpq ∠θpf :pq ˛
˛ ˛
˛ ksc−lv ˛

1
IM „ «
1+ kkm ks
sc−lv

5
Li can represent any type (i.e. constant impedance, constant current, constant power and
induction motor loads) of load element.
6 1 ˛γ where γ = 1, 0, −2 ∼ −1 for constant impedance, constant current and
i.e. ˛˛ 1
∠θpf :lv ˛
˛
˛1+j k
sc−lv
1
constant power loads respectively, and “
ks
” for induction motor loads.
1+ k
sc−lv
Appendix M

Application of the Methodology

Given by (5.37) to the Three-bus

MV Test System (Fig. 5.16)

The matrix equation (5.37) which gives the proposed methodology for evaluating influ-

ence coefficients [ki−x ] between any busbar i and other busbars x (x = 1, 2, ..., n x 6= i)

is reproduced here by (M.1):


0 −1
[ki−x ](n−1)×1 ≈ [Y++:xz ](n−1)×(n−1) [Y++:xi ](n−1)×1 (M.1)

where,
0
Y++:xz ≈ Y++:xz + Y−−:x−im for x = z
0
Y++:xz = Y++:xz for x 6= z

For the purpose of evaluating the influence coefficients k1−2 and k1−3 between

busbar 1 (i.e. i = 2) and busbars 2 and 3 (i.e. x, z = 2, 3) of the three-bus MV test

240
241

system shown in Fig. 5.16, the matrices [Y++:xz ] and [Y++:xi ] are:

 
 0.2020 − j0.4126 −0.1010 + j0.2063 
[Y++:xz ] =  S (M.2)
−0.1010 + j0.2063 0.3029 − j0.6189

 
 −0.1010 + j0.2063 
[Y++:xi ] =  S (M.3)
−0.2020 + j0.4126

As busbar 3 supplies passive loads (i.e. km:3 = 0), the admittance Y−−:3−im ≈ 0.

• Case 1 - busbar 2 supplies passive loads or km:2 = 0. That is, the admittance
0
Y−−:2−im ≈ 0 implying that [Y++:xz ] ≈ [Y++:xz ]. Then, the substitution of (M.2)

and (M.3) in (M.1) gives:    


 k1−2   1 
 =  (M.4)
k1−3 1

• Case 2 - busbar 2 supplies motor loads at the LV level or km:2 = 1. Equation

(5.36) gives the admittance Y−−:2−im for klv:2 = 1, ksc−lvagg :2 = 18, ks:2 = 6.7,

I+:2 = 194A and Vn−mv = 12.47kV as:

Y−−:2−im ≈ −j0.1316S (M.5)

0
Then, the matrix [Y++:xz ] can be obtained as:

 
0  0 0
[Y++:xz ] ≈ [Y++:xz ] +  S (M.6)

0 −j0.1316
242

Substitution of (M.3) and (M.6) in (M.1) gives:

   
 k1−2   0.75 
 =  (M.7)
k1−3 0.91
Appendix N

66kV Sub-transmission

Interconnected Study System

(Fig. 7.1) - Additional

Data/Information

N.1 Operating Conditions at the Considered Time Stamp

Note that the impedance values given in pu are based on a 100M V A base.

Table N.1: System details


Nominal voltage 66kV (line-line)
Nominal frequency 60Hz
Connection type three-wire

243
244

Table N.2: Voltage controlled bus data


Busbar Voltage
magnitude (pu)
S1 1.023
S5 0.975

Table N.3: Generator and load bus data: three-phase MW and MVAr values
Busbar Generation Load
MW MVAr MW MVAr
S1 (reference bus) 0 0 0 0
S2 0 0 18.06 0.36
S3 0 0 18.27 7.11
S4 0 0 10.68 5.13
S5 13.50 0 0 0
S61 0 0 0.09 3.99
S7 0 0 33.99 14.52
S8 0 0 5.97 1.89
S9 0 0 1.65 0.24
1
This represent the operation (balanced) of a static Var compensator.
245

Table N.4: Voltage regulator data: impedances and secondary tap settings
Busbar Impedance Secondary tap
(pu) setting
S2 0.0014 + j0.0192 1.058
S7 0.0014 + j0.0192 1.111

Table N.5: Static capacitor data: susceptances


Busbar Susceptance
(pu)
S2 0.0218
S3 0.0597
S4 0.0409
S7 0.1277
S8 0.0130

Table N.6: Generator impedance data


Busbar Sequence impedances (pu)
Zero Positive Negative
S1 0 0 0
S5 j0.5 j0.6 j0.6
246

N.2 Line Data

Notations:

Z++ - positive sequence impedance of a line

Z−+ - negative-positive sequence coupling impedance of a line

Table N.7: Lengths and impedances (Z−+ and Z−+ ) of the sub-transmission lines
Line Length Z−+ Z++
(km) (Ω) (Ω)
A 67.65 0.62∠840 24.09∠730
B 19.16 0.52∠300 8.36∠530
C 17.83 0.25∠3360 6.37∠720
D 71.49 0.74∠1780 24.81∠710
E 19.59 0.27∠650 9.63∠480
F 45.37 1.25∠300 22.81∠450
G 66.29 0.14∠340 32.50∠450
H 56.46 0.03∠470 28.80∠450
I 55.32 1.40∠300 18.98∠720
J 11.40 0.31∠1480 4.01∠720
K 15.57 0.08∠1210 5.47∠730
L 80.65 0.30∠1310 41.00∠450
M 83.20 1.91∠550 29.90∠640
N 21.16 0.45∠230 9.23∠590

N.3 An Explanation on the Influence of the Location of an

Asymmetrical Line of an Interconnected Network on the

Voltage Unbalance Behaviour of the Line

• When a line (e.g. line J of the study system) is located in the downstream part

of its network, it introduces voltage unbalance primarily at the downstream


247
t
busbars as the negative sequence voltage V−:rec−tany
at any other upstream bus-

bar (e.g. consider S2 and S6 in the case of line J) has to satisfy the relationship
t
given by (7.3) where V−:send−tany
is approximately zero1 .

• Alternatively, a line (e.g. line A of the study system) which is located in the

upstream part of its network can introduce voltage unbalance at the upstream

busbars as well as the downstream busbars (e.g. consider the upstream busbar

S2 and the downstream busbar S7 in the case of line J) as the line is in a position
t
to take part of V−:send−tany
for all other lines.

• When a line (e.g. line B of the study system) is isolated from the rest of

the network, its influence at other main busbars, compared to that of a line

(e.g. line J of the study system) which is located in the main part of the network,
t
is somewhat reduced. Because, the negative sequence voltage V−:rec−tany
at any

other busbar (e.g. consider S2 in the case of line B) has to satisfy the relationship

(7.3) where there can exist several connections (e.g. lines E and A for S2) to
t t
the busbar resulting in several V−:send−tany
(e.g. V−:send−E ≈ Z−+:B I+:B , and
t
V−:send−A = 0 for S2) out of which at least one or more can be approximately

zero.

N.4 A Demonstration of the Linearity of Negative Sequence

Voltages

Table N.8 gives the negative sequence voltages2 (V−:S2


t
) introduced by the individual

lines A - N of the study system at the busbar S2. Table N.9 gives a comparison of
1
Because the voltage at the bulk supply is balanced, and the term Z++:tany I−:t/tany in (7.3) is
inconsiderable.
2
These, which are obtained using unbalanced load flow analysis, correspond to the selected time
stamp.
248

the resultant negative sequence voltage3 (V−:S2


lines
) at S2 obtained using the summation

of the individual vectors given in Table N.8, and unbalance load flow analysis.

t
Table N.8: Negative sequence voltages V−:S2 caused by the individual lines A - N
at the busbar S2
t
Line (t) V−:S2 (V )
A 58.560∠−127.1760
B 14.691∠−120.6320
C 12.124∠151.7370
D 45.203∠−24.2310
E 4.401∠−139.6360
F 95.222∠−137.7970
G 0.950∠−134.2140
H 0.257∠−119.4290
I 54.234∠176.4530
J 14.402∠−47.7590
K 1.850∠−75.2630
L 1.409∠−68.1720
M 0.361∠−170.8340
N 13.425∠176.6590

lines
Table N.9: Resultant negative sequence voltage V−:S2 at the busbar S2
lines
V−:S2 (V )
Vector summation 207.735∠−133.9680
Unbalanced load flow analysis 205.703∠−134.3440

3
i.e. as a result of the interaction of all lines.
Appendix O

Development of a Method for

Unbalanced Load Flow Analysis

O.1 Introduction

Unbalanced load flow is an essential tool for analysing steady-state unbalanced power

system problems (e.g. voltage unbalance). Due to the need for careful representation

of power system components and the lack of widespread availability of comprehensive

commercial packages, a generaliased unbalanced load flow program which is based on

M AT LAB R is developed. This appendix gives a detailed description on the program

including the background information (Sections O.2 and O.3), and the representation

of power system components (Section O.4) employed for the formulation of load flow

equations. The key section, Section O.4.5, discusses existing models of three-phase

induction motors that have been used for unbalanced load flow studies, and proposes

two types of models which overcome the limitations associated with the existing

models.

249
250

O.2 Symmetrical Component Versus Phase Coordinate Ref-

erence Frames for Unbalanced Load Flow Analysis

References [1, 2, 3, 4, 5] propose two basic approaches for unbalanced load flow analy-

sis, which are based on the symmetrical component and the phase coordinate reference

frames respectively. Analysis of unbalanced power system problems has been tradi-

tionally based on the symmetrical component reference frame due to the advantages

of the availability of sequence impedances for power system components, and the

decoupled nature of most power system components in the symmetrical component

reference frame [1]. However, the use of the phase coordinate reference frame has been

identified as the best way to represent three-phase power system components, as it

facilitates the maintenance of the initial physical identity of the system with regard

to line parameters and variables such as nodal voltages and line currents [2, 4, 5].

The only drawback of this approach is that the size of the problem is significantly

large compared to that based on the symmetrical component reference frame [3].

Disregarding the computational advantages associated with the symmetrical com-

ponent approach, an unbalanced load flow program that is based on the phase coor-

dinate reference frame is developed for specific applications of the work presented in

this thesis.

O.3 Special Considerations in Developing an Unbalanced Load

Flow Program

Compared to balanced load flow algorithms, a number of additional issues are to be

addressed in developing an unbalanced load flow method:

• First is the question on how to formulate unbalanced load flow equations in


251

a generalised manner so that the incorporation of numerous component con-

nections (e.g. single-phase and dual-phase loads) can be easily achieved. The

concept of specifying load flow constraints for each bus or each phase of a bus,

as used in the traditional balanced or unbalanced load flow methods, cannot

take various component connections into account. In view of the fact that the

power constraints such as the specified power generation and consumption are

properties of system components instead of buses, [4, 5] proposes a new concept

that is referred to as ‘component level power flow constraints’ which allows the

incorporation of numerous component connections.

• Second is the question on how to represent the behaviour of system components

when subjected to unbalanced voltage/current excitations. As an example, a

three-phase induction motor responds differently to the positive and negative se-

quence applied voltages, whereas a typical residential load would exhibit nearly

identical behaviours with both the positive and negative sequence voltages.

The developed unbalanced load flow program incorporates the suggested concept

of the component level power flow constraints, and takes into account the second

requirement in the modelling of system components.

O.4 Representation of System Components

O.4.1 Synchronous Generators

The synchronous generator model used in the developed program is based on [4,

5], which takes the different machine responses for the positive, negative and zero

sequence current injections into account. As illustrated in Fig. O.1 and given by

(O.1), this is a positive sequence voltage source behind the generator admittance

matrix [Yg ] with no pre-determined connection form.


252

a Eg

k b a2Eg m
[Yg]
c aEg

Figure O.1: Synchronous generator model

[Ikm:g ] = [Yg ]([Vk ] − [Vm ] − [Eg ]) (O.1)

where,

[Vk ] = [Va:k Vb:k Vc:k ]t , matrix of phase voltages on side k

[Vm ] = [Va:m Vb:m Vc:m ]t , matrix of phase voltages on side m

[Eg ] = [Eg a2 Eg aEg ]t , matrix of the generator internal phase voltages (balanced)
¯ ¯
[Ikm:g ] = [Ikm:g−a Ikm:g−b Ikm:g−c ]t , matrix of the generator currents flow from side

k to side m

[Yg ] = T [Yg:seq ]T −1

 
 Yg:0 0 0 
 
[Yg:seq ] = 
 0 Yg:+ 0 

 
0 0 Yg:−
 
 1 1 1 
 
T = 2
 1 ¯a a¯


 
1 a a2
¯ ¯

Yg:0 , Yg:+ , Yg:− - zero, positive and negative sequence admittances respectively of

the generator
253

superscript t - represents the transpose of a matrix

a = 1∠1200
¯

If the generator operates as a slack machine, the magnitude and the phase angle

of the positive sequence voltage at the generator terminals are constrained:

[T+ ]([Vk ] − [Vm ]) = Vg:spec ∠θg:spec (O.2)

where,

[T+ ] = 13 [1 a a2 ]
¯ ¯
|Vg:spec |, θg:spec - specified magnitude and phase angle respectively of the positive se-

quence voltage at the generator terminals

If the generator operates as a PV machine1 , the three-phase active power and the

magnitude of the positive sequence voltage at the generator terminals are constrained:

Real − [Ikm:g ]c ([Vk ] − [Vm ]) = Pg:spec



(O.3)

|[T+ ]([Vk ] − [Vm ])| = |Vg:spec | (O.4)

where,

Pspec - specified generation of the three-phase active power


c
superscript - represents the conjugate of a vector
1
This refers to a power and voltage controlled generator.
254

O.4.2 Passive Loads

The exponential load model [6] which takes the voltage dependency of the active

and reactive power into account is employed. As illustrated in Fig. O.2, this is

represented as a branch between two nodes allowing the incorporation of different

load configurations.
 λp
|Vk − Vm |
P = P0 (O.5)
V0
 λq
|Vk − Vm |
Q = Q0 (O.6)
V0

where,

P , Q - active power and reactive power respectively consumed by the load

P0 , Q0 - P and Q respectively under the rated voltage

V0 - rated voltage

λp , λq - voltage coefficients of active power and reactive power respectively, where

λp = λq = 2 for constant impedance loads , λp = λq = 1 for constant current loads,

λp = λq = 0 for constant power loads, and different values can be chosen to represent

the aggregated effects of various load mixes

The active and reactive power consumed by this load are constrained:

c
Ikm:l (Vk − Vm ) = Pl:spec + jQl:spec (O.7)

where,

Ikm:l - load current flows from side k to side m

Pl:spec , Ql:spec - specified consumption of the active and reactive power respectively of

the load
255

k m
P + jQ

Figure O.2: Load model

O.4.3 Overhead Lines

Overhead lines are modelled as electromagnetically coupled2 impedance matrices in



the phase coordinate reference frame. The phase impedance matrix [Zt:xy ](3×3) for

a three-wire line3 with the earth return is derived using [8]:

 !
De
Zt:xy = Rearth +Rconductor +j k ln in Ω/m, when x = y, i.e. self impedance
Dxy
(O.8)
 !
De
Zt:xy = Rearth + j k ln in Ω/m, when x 6= y, i.e. mutual impedance
Dxy
(O.9)

where,

Rearth = 9.869 × 10−7 f in Ω/m, earth resistance


q
De = 658.376 × rfo in m

k = 2× 10−7 in H/m


 conductor geometric mean radius (m), when x = y




Dpq − geometric mean distance between the conductor x and the conductor y (m),



 when x 6= y

Rconductor - ac resistance of the conductor x (Ω/m)

f - operating frequency (Hz)

ro - earth resistivity (Ωm)

x, y - represents the three phases a, b and c


2
Capacitive effects are ignored.
3
Single-circuit.
256

O.4.4 Capacitor Banks

Capacitor banks are represented as constant impedance elements which allow their

reactive power injection to be determined according to the prevailing voltage condi-

tion.

O.4.5 Three-phase Voltage Regulators/Transformers

Voltage regulators/transformers are represented using three single-phase star con-

nected units. Each unit is modelled as an impedance in series with an ideal trans-

former having taps on the secondary. The equivalent circuit is shown in Fig. O.3

where [Avr:x ], [Bvr:x ] and [Cvr:x ] are 3 × 3 diagonal matrices of which the elements

Avr:x , Bvr:x and Cvr:x respectively are given by [7]:

Avr:x = tapx yvr:x (O.10)

Bvr:x = tapx (tapx − 1) yvr:x (O.11)

Cvr:x = (1 − tapx ) yvr:x (O.12)

i, j - represent the busbars at which the primary and the secondary respectively are

connected

tapxy - tap position, as a ratio of the primary voltage and the secondary voltage, of

any phase x (= a, b, c)

yvr:x - series impedance, referred to the secondary, of the phase x

O.4.6 Three-phase Induction Motors

A number of different models to represent the behaviour of an induction motor in

unbalanced load flow studies has been proposed [4, 5, 11, 13]. A simple model based

on the sequence equivalent circuits is given in [4, 5], where the positive sequence
257

Busbar i Busbar j

[Avr:x]

[Bvr:x] [Cvr:x]

Figure O.3: Equivalent circuit of a voltage regulator/transformer

impedance is considered as an unknown variable which is to be determined so that

the total real and reactive power drawn by the motor are equal to specified values. In

this model, the unknown positive sequence impedance is represented by an unknown

voltage source behind the known negative sequence impedance, and the voltage source

is solved to meet the power flow constraints stated above. The phase domain model

of the above, which is established by transforming the sequence elements into the

phase domain, is illustrated in Fig. O.4 and expressed by (O.13). The considered

power flow constraints are given by (O.14).

Eim
a

k b a2Eim m
[Yim]
c aEim

Figure O.4: Three-phase induction motor model proposed in [4, 5]


258

[Ikm:im ] = [Yim ]([Vk ] − [Vm ] − [Eim ]) (O.13)

−[Ikm:im ]c ([Vk ] − [Vm ]) = Pim:spec + jQim:spec (O.14)

where,
−1
[Yim ] = T [Yim:seq
 ]T 
 Yim:0 0 0 
 
[Ysequence ] = 
 0 Yim:− 0 

 
0 0 Yim:−

Yim:0 , Yim:− - zero and negative sequence admittances respectively of the induction mo-

tor

[Ikm:im ] = [Ikm:im−a Ikm:im−b Ikm:im−c ]t , matrix of the induction motor currents flow

from side k to side m

Eim - unknown voltage

Pim:spec , Qim:spec - specified motor input active and reactive power respectively

Reference [10] reports that although the active power drawn by a three-phase

induction motor is nearly independent of the supply voltage level until the point of

stalling, the reactive power is more sensitive to the voltage level. Fig. O.5, which

is reproduced using [10], illustrates the variation of the real (P) and reactive (Q)

power drawn by a typical induction motor with the voltage level justifying the above

statement4 . Hence, the power flow constraints used in the model proposed in [4, 5],

which controls the reactive power drawn by the motor at a specified value, does not

represent the actual motor behaviour resulting erroneous outcomes.

The subject of the modelling of three-phase induction motors for unbalanced load

flow studies has received increased attention recently by the IEEE Distribution System
4
The rated mechanical load and balanced supply voltages are assumed.
259

2.6
2.4 P
2.2 Q
2.0
1.8
P, Q (pu) 1.6
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.55 0.60 0.65 0.70 0.75 0.80 0.85 0.90 0.95 1.00
Voltage (pu)

Figure O.5: Variation of the real (P) and reactive (Q) power with the supply voltage
level for a typical three-phase induction motor

Analysis Sub-committee with the view to provide reference test cases for developers

to validate the induction machine models under unbalanced conditions [11, 12, 13].

The model discussed in [11, 12] is also based on the sequence equivalent circuits,

and considers both the positive and negative sequence impedances as functions of

the motor slip which is taken as a new state variable to be determined such that the

input real power is equal to a specified value. The reactive power is then adjusted by

the load flow algorithm according to the existing supply voltage condition. This is a

valid representation only if the specified power corresponds to the actual operating

speed which is determined by the characteristics of both the electromechanical torque

developed by the motor and the mechanical torque demanded by the load. However,

the nominal input real power corresponding to the rated motor speed has been used

as the specified power in [11, 12]. In other words, the rated mechanical load which

lead the motor to operate at the rated speed has been assumed. Although the real

power drawn by an induction motor exhibits a relatively low dependency on supply

voltage conditions until the point of stalling (see Fig. O.5), it is highly sensitive to
260

the changes in mechanical loading conditions. As an example, Figs. O.6 and O.7

illustrate the variation5 of the real (P) and reactive (Q) power and the motor speed

respectively of a 2250hp motor with various characteristics6 of pump systems, when

the motor is excited at the rated voltage (balanced). These demonstrate that the

changes in mechanical loading conditions have a high degree of influence on the real

and reactive power, whereas such changes have only a minor influence on the motor

speed. The rated mechanical loading condition assumed in the model discussed in

[11, 12] may not always arise in practice, and hence this model also has limitations

in generalised applications.

2.5

P kp:rated = 0.93
2.0 Q
P (MW), Q (MVAr)

1.5

1.0

0.5

0.0
kp 0.3 0.5 0.7 0.9 1.1
Loading
33 54 76 97 118
level (%)

Figure O.6: Variation of the real (P) and reactive (Q) power with kp (motor loading
levels corresponding to various kp is also given as a percentage to the rated output
power) for a 2250hp induction motor

In the remaining part of this section, two types of induction motor models which

overcome the above stated limitations are proposed:


5
Obtained using PSCAD/EMTDC.
6
Various characteristics are represented using kp , where Torque = kp × speed2 . kp:rated represents
the rated load.
261

• Impedance type - the operation of an induction motor is represented using

impedance/admittance elements ([Yim ]) as shown in Fig. O.8, together with a

power flow constraint.

• PQ type - the operation of an induction motor is represented using active and

reactive power components as shown in Fig. O.97 , together with a power flow

constraint.

1.2

1.0
Motor speed (pu)

0.8

0.6
kp:rated = 0.93

0.4

0.2

0.0
kp 0.3 0.5 0.7 0.9 1.1
Loading
33 54 76 97 118
level (%)

Figure O.7: Variation of the speed with kp (motor loading levels corresponding to
various kp is also given as a percentage to the rated output power) for a 2250hp
induction motor

Busbar k
a

b [Yim]
c

Figure O.8: Impedance type induction motor model

7
Where, Pim:a , Pim:b , Pim:c - active power drawn by the phases a, b and c respectively of the
motor, Qim:a , Qim:b , Qim:c - reactive power drawn by the phases a, b and c respectively of the motor.
262

Busbar k
a Pim:a + jQim:a
b Pim:b + jQim:b
c Pim:c + jQim:c

Figure O.9: PQ type induction motor model

Impedance Type Induction Motor Model

This is an extension to the model given in [11, 12], which is based on the sequence

equivalent circuits. The positive and negative sequence equivalent circuits of an in-

duction motor are shown in Fig. O.10. The positive (Zim:+ ) and negative (Zim:− ) se-

quence impedances, as functions of the motor speed, are given by (O.15) and (O.16)

respectively. The zero sequence impedance is irrelevant because of the three-wire

connection associated with motors.

1
Zim:+ = rst + jxst + 1 1 (O.15)
jxmag
+ rrt ωsyn
+jxrt
ωsyn −ωrt

1
Zim:− = rst + jxst + 1 1 (O.16)
jxmag
+ rrt ωsyn
+jxrt
ωsyn +ωrt

where,

rst , xst - stator resistance and leakage reactance respectively

rrt , xrt - rotor resistance and leakage reactance respectively (referred to the stator

side)

xmag - magnetizing reactance

ωsyn - synchronous speed

ωrt - rotor angular speed


263

rrt ω syn
rst jxst ω syn − ω rt jxrt

jxmag

I
rrt ω syn
rst jxst ω syn + ω rt jxrt

jxmag

II

Figure O.10: Sequence equivalent circuits of a three-phase induction motor: I - posi-


tive sequence, II - negative sequence
264

In the sequence domain, a three-phase induction motor can be represented as:

f1 = V0:k Yim:0 − I0:im = 0 (O.17)

f2 = V+:k Yim:+ − I+:im = 0 (O.18)

f3 = V−:k Yim:− − I−:im = 0 (O.19)

where,

Yim:0 = 0 (noting the three-wire connection)


1
Yim:+ = Zim:+
1
Yim:− = Zim:−

V0:k , V+:k , V−:k - zero, positive and negative sequence voltages respectively at the

motor terminals (busbar k)

I0:im , I+:im , I−:im - zero, positive and negative sequence currents respectively drawn

by the motor

As the load flow method is based on the phase coordinate reference frame, the se-

quence domain equations (O.17) - (O.19) are transformed into the phase domain. Thus:

f1 = f (Iim:a , Iim:b , Iim:c , Va:k , Vb:k , Vc:k )

f2 = f (Iim:a , Iim:b , Iim:c , Va:k , Vb:k , Vc:k , ωrt )

f3 = f (Iim:a , Iim:b , Iim:c , Va:k , Vb:k , Vc:k , ωrt )

where,

Va:k , Vb:k , Vc:k - three phase voltages at the supply terminals of the motor

Ia:im , Ib:im , Ic:im - three phase currents drawn by the motor


265

The motor speed ωrt associated with (O.17) - (O.19) is a new variable to be de-

termined by the load flow algorithm. Thus, an additional equation is required to

complete the load flow formulation, which essentially has to be the power flow con-

straint under which the induction motor operates. The general situation where the

electromechanical torque/power developed by the motor is equal to the mechanical

torque/power demanded by the rotating load is applied as the power flow constraint

instead of constraining the power drawn by the motor at the nominal values as in the

case of the existing models.

The resultant electromechanical torque developed by an induction motor is equal

to the sum of the two torque components corresponding to the positive and negative

sequence voltage/current inputs. The positive (Tim:+ ) and negative (Tim:− ) sequence

torque components can be expressed in terms of the three phase voltages, motor

parameters and motor speed as:


 
2 Zmag 2
1 (Va:k + ¯aVb:k + ¯a Vc:k ) Zmag +Zst 
rrt 
Tim:+ = Zmag Zst
(O.20)
3 + jxrt + ωrsyn
rt ωsyn ωsyn − ωrt

Zmag +Zst −ωrt

 
2 Zmag 2
(V + a V + aV )
1 a:k ¯ b:k ¯ c:k Zmag +Zst  rrt 
Tim:− =− Zmag Zst
(O.21)
3 + jxrt + ωrrt ω+ω ωsyn + ωrt
syn

Z +Z
mag st syn rt

where,

Zst = rst + jxst

Zmag = jxmag

Considering pump systems, torque characteristics of the mechanical load (Tload ) can

be expressed as:
2
Tload = kl−1 + kl−2 ωrt (O.22)
266

where, kl−1 , kl−2 are constants for a given pump system. Then, the suggested power

flow constraint can be given by:

f4 = Tim:+ + Tim:− − Tload = 0 (O.23)

where,

f4 = f (Va:k , Vb:k , Vc:k , ωrt )

The functions f1 , f2 , f3 and f4 which give the complete load flow formulation for

a three-phase induction motor allow the load flow algorithm to determine the motor

speed and three stator phase voltages and currents thus naturally adjusting the input

real and reactive power.

PQ Type Induction Motor Model

The three-phase induction motor model proposed in this section is in line with the

exponential load model that has been used to represent induction motors in balanced

load flow studies [10, 14].

The real and reactive power drawn by each of the three phases of an induction

motor can be expressed using the stator phase voltages and the phase admittance

matrix as:
     c
 Pim:a + jQim:a   Va:k 0 0   Va:k 
     
 P = 0 V  [Yim ]c  V (O.24)
 im:b + jQim:b 0 
  b:k   b:k 
     
Pim:c + jQim:c 0 0 Vc:k Vc:k
267

where,  
 Yim:s Yim:m1 Yim:m2 
 
[Yim ] = 
 Yim:m2 Yim:s Yim:m1


 
Yim:m1 Yim:m2 Yim:s

Yim:s = Yim:0 + Yim:+ + Yim:− , self admittance

Yim:m1 = Yim:0 + aYim:+ + a2 Yim:− , a mutual admittance


¯ ¯
2
Yim:m2 = Yim:0 + a Yim:+ + aYim:− , a mutual admittance
¯ ¯

Referring to (O.24), the complex power per phase can be generally written as:

c
Pim:x + jQim:x = Yim:s |Vx:k |2 + Yim:m1
c c
Vx:k Vy:k c
+ Yim:m2 c
Vx:k Vz:k (O.25)

where, x, y, z - represent the three phases a, b and c. The real and reactive power

components in (O.25) can be separated as:

Pim:x = Pim:x−xx + Pim:x−xy + Pim:x−xz (O.26)

Qim:x = Qim:x−xx + Qim:x−xy + Qim:x−xz (O.27)

where,

Pim:x−xx = |Yim:s |cos(−θim:s )|Vx:k |2

Pim:x−xy = |Yim:m1 |cos(−θim:m1 + θx:k − θy:k )|Vx:k ||Vy:k |

Pim:x−xz = |Yim:m2 |cos(−θim:m2 + θx:k − θz:k )|Vx:k ||Vz:k |

Qim:x−xx = |Yim:s |sin(−θim:s )|Vx:k |2

Qim:x−xy = |Yim:m1 |sin(−θim:m1 + θx:k − θy:k )|Vx:k ||Vy:k |

Qim:x−xz = |Yim:m2 |sin(−θim:m2 + θx:k − θz:k )|Vx:k ||Vz:k |

θim:s , θim:m1 , θim:m2 - phase angles of Yim:s , Yim:m1 and Yim:m2 respectively

θx:k , θy:k , θz:k - phase angles of Vx:k , Vy:k and Vz:k respectively
268

For the case of balanced voltage angles8 (i.e. θx:k −θy:k = 1200 and θx:k −θz:k = −1200 ),

the power components Pim:x−xx - Qim:x−xz can be rearranged as:

 |Y
im:s |cos(−θim:s )
 |V
x:k |
0
2
n
Pim:x−xx = Pim:x−xx n n n
(O.28)
|Yim:s | cos(−θim:s ) |Vx:k |

0 n
 |Y
im:m1 |cos(−θim:m1 + 1200 )  |Vx:k ||Vy:k | 
Pim:x−xy = Pim:x−xy n n n n
(O.29)
|Yim:m1 |cos(−θim:m1 + 1200 ) |Vx:k ||Vy:k |

0 n
 |Y
im:m2 |cos(−θim:m2 − 1200 )  |Vx:k ||Vz:k | 
Pim:x−xz = Pim:x−xz n n n n
(O.30)
|Yim:m2 |cos(−θim:m2 − 1200 ) |Vx:k ||Vz:k |

 |Y
im:s |sin(−θim:s )
 |V
x:k |
0
2
Qim:x−xx = Qnim:x−xx n n n
(O.31)
|Yim:s |sin(−θim:s ) |Vx:k |

0
 |Y
im:m1 |sin(−θim:m1 + 1200 )  |Vx:k ||Vy:k | 
Qim:x−xy = Qnim:x−xy n n n n
(O.32)
|Yim:m1 |sin(−θim:m1 + 1200 ) |Vx:k ||Vy:k |

0
 |Y
im:m2 |sin(−θim:m2 − 1200 )  |Vx:k ||Vz:k | 
Qim:x−xz = Qnim:x−xz n n n n
(O.33)
|Yim:m2 |sin(−θim:m2 − 1200 ) |Vx:k ||Vz:k |

where,
0
- refers to the condition of balanced voltage angles

superscript n - refers to the parameters corresponding to the nominal conditions

(rated voltage and rated motor speed)


n n 2 n n
Pim:x−xx = |Vx:k | |Yim:s |cos(−θim:s )
n n n n n
Pim:x−xy = |Vx:k ||Vy:k ||Yim:m1 |cos(−θim:m1 + 1200 )
n n n n n
Pim:x−xz = |Vx:k ||Vz:k ||Yim:m2 |cos(−θim:m2 − 1200 )

Qnim:x−xx = |Vx:k
n 2 n
| |Yim:s n
|sin(−θim:s )
8
In most practical circumstances unbalance in phase voltages arise mainly due to the unbalance
in their magnitudes. Thus, this condition of balanced voltage angles can be considered as a general
case.
269
n
Qnim:x−xy = |Vx:k n
||Vy:k n
||Yim:m1 n
|sin(−θim:m1 + 1200 )

Qnim:x−xz = |Vx:k
n n
||Vz:k n
||Yim:m2 n
|sin(−θim:m2 − 1200 )

0 0
Referring to (O.28) - (O.33), the power components Pim:x−xx - Qim:x−xz can be written

in a generalised form as:

  
0 n
Y |Vi:k ||Vj:k |
Sim:x−ij = Sim:x−ij (O.34)
Y n |V n ||V n |
i:k j:k

where,
0 0 0
Sim:x−ij - represents any Pim:x−ij or Qim:x−ij where i, j = x, y, z
n 0
Sim:x−ij - nominal value of Sim:x−ij (i.e. at the rated voltage and the rated motor

speed)

|Y | - represents the magnitude of various admittance components which are functions

of the motor speed

|Y n | - nominal value of |Y | (i.e. at the rated motor speed)

For the purpose of illustrating the characteristics of the various power and admit-

tance elements in (O.34), a 60Hz, 3hp, 220V induction motor with the parameters
n
given in Table O.1 [15] is used. Table O.2 gives the power components Px−xx - Qnx−xz

under the nominal conditions for the above motor.



The admittance ratio YYn can be shown to be approximately equal to an exponen-
ωrt
tial form of the speed ratio n
ωrt
as given by (O.35). As examples, Figs. O.11 and O.12 il-
|Yim:s | cos(−θim:s ) 0 |Yim:m2 | sin(−θim:m2 −1200 ) 0
lustrate the variation of n | cos(−θ n
|Yim:s
of Px−xx and |Yin m:m2| sin(−θim:m2
n −1200 )
of Qx−xz
im:s )

ωrt
respectively with n
ωrt
over a range of ωrt corresponding to 125% - 25% of the rated

motor loading level.

 
Y
= ωrt
λsp:s−ij
(O.35)
Y n ωn rt
270

Table O.1: Parameters of a 60Hz, 3hp, 220V induction motor


Parameter Value
ωsync 1800rpm
n
ωrt 1710rpm
rst 0.435Ω
xst 0.754Ω
rrt 0.816Ω
xrt 0.754Ω
xmag 26.13Ω

n
Table O.2: Power components Px−xx - Qnx−xz for the 3hp, 220V motor
P n (W) Qn (VAr)
x − xx 1834.94 2968.26
x − xy 1923.00 -2483.74
x − xz -2842.67 166.41

1.10
Approximation: y = x-3.3828
Actual variation
1.05
Yim:s cos( −θ im:s )
Yimn :s cos( −θ imn :s )

1.00

0.95

0.90

0.85
0.98 0.99 1.00 1.01 1.02 1.03 1.04 1.05
ω rt
ω rtn

|Yim:s | cos(−θim:s ) 0 ωrt


Figure O.11: Variation of n | cos(−θ n
|Yim:s
of Px−xx with n for the 3hp, 220V motor
im:s ) ωrt
271

where, λsp:s−ij - speed coefficient corresponding to the power component Sim:x−ij ,

which is a constant for a given motor. Hence, (O.34) can be written as:

0
 ω λsp:s:ij  |V ||V | 
n rt i:k j:k
Sim:x−ij = Sim:x−ij n n n
(O.36)
ωrt |Vi:k ||Vj:k |

A speed range corresponding to 125% - 25% of the rated motor loading level is

usually sufficient to cover most practical circumstances, within which the accuracy

of the approximation (O.35) is seen to be acceptable. A higher degree of accuracy

can be achieved if the speed range under consideration can be further narrowed. The

speed coefficients corresponding to the power components Px−xx - Qx−xz for a range

of induction motors [15] are given in Table O.3.

1.2
Approximation: y=x-5.5278
Actual variation
Yim:m 2 cos( −θ im:m 2 − 120 0 )
Yimn :m 2 cos( −θ imn :m 2 − 120 0 )

1.1

1.0

0.9

0.8

0.7
0.98 0.99 1.00 1.01 1.02 1.03 1.04 1.05

ω rt
ω rtn

|Yim:m2 | sin(−θim:m2 −1200 ) 0 ωt


Figure O.12: Variation of n
|Yim:m2 n
| sin(−θim:m2 −1200 )
of Qx−xz with n
ωrt
for the 3hp, 220V
motor

The power components Px−xx - Qx−xz which represent the behaviour of an induc-

tion motor under both unbalanced voltage magnitudes and unbalanced voltage angles
272

Table O.3: Speed coefficients corresponding to the power components Px−xx - Qx−xz
for a range of induction motors
Motor specification
λsp:s−ij 3hp, 220V 50hp, 460V 500hp, 2.3kV 2250hp, 2.3kV
λsp:p−xx -3.3828 -5.9151 -40.658 -94.6090
λsp:p−xy -2.9136 -3.0044 -12.089 -19.8360
λsp:p−xz 2.0121 2.7340 13.9710 23.2230
λsp:q−xx -0.0848 -0.4081 -2.9381 -4.2029
λsp:q−xy 0.2243 0.6279 5.6215 8.3997
λsp:q−xz -5.5278 3.8814 10.469 11.9090

0 0
can be expressed in terms of the power components Px−xx - Qim:x−xz (which represent

the behaviour under the condition of balanced voltage angles) as:

0
Pim:x−xx = Pim:x−xx (O.37)
0 0
Pim:x−xy = Pim:x−xy cos(θxy:k ) − Qim:x−xy sin(θxy:k ) (O.38)
0 0
Pim:x−xz = Pim:x−xz cos(θxz:k ) − Qim:x−xy sin(θxz:k ) (O.39)
0
Qim:x−xx = Qim:x−xx (O.40)
0 0
Qim:x−xy = Qim:x−xy cos(θxy:k ) + Pim:x−xy sin(θxy:k ) (O.41)
0 0
Qim:x−xz = Qim:x−xz cos(θxz:k ) + Pim:x−xy sin(θxz:k ) (O.42)

where,

θxy:k = (θx:k − θy:k ) − 1200

θxz:k = (θx:k − θz:k ) + 1200

The power flow constraint considered in the impedance type model, which is given

by (O.23), can be rewritten in terms of power (= torque × ωrt ) as:

Pim:+ + Pim:− = Pload (O.43)


273

where,

Pim:+ , Pim:− - gross electromechanical power developed by the motor corresponding

to the positive and negative sequence voltage/current inputs respectively

Pload - mechanical power demanded by the load

The total gross electromechanical power developed by the motor (i.e. Pim:+ + Pim:− )

can be expressed in terms of the input active power components (given by (O.26)) in

the phase domain and the motor efficiency (ηim ) as:

Pim:+ + Pim:− = ηim (Pim:a + Pim:b + Pim:c ) (O.44)

Based on the observation that the real power drawn by an induction motor is nearly

independent of the supply voltage condition until the point of stalling (Fig. O.5),

the motor efficiency ηim can be expressed in terms of the motor parameters and the

speed as:
Pgross at V n
ηim = (O.45)
Pin at V n

where,

Pgross at V n - total gross electromechanical power developed by the motor at the rated

voltage V n (balanced)  
Z
2
V n Z mag  
mag +Zst ωrt rrt
Pgross at V n = 3 Zmag Zst

rrt ωsync
Zmag +Zst +jxrt + ωsync −ωrt ωsync −ωrt
Pin at V n - motor input real power at the rated voltage V n (balanced)
 n2 
Pin at V n = 3 real Z|Vc |
im:+

The motor efficiency ηim , as a function of the motor speed ωrt , can be shown to

be approximated using (O.46) with a reasonable degree of accuracy. As an exam-

ple, Fig. O.13 illustrates the variation of ηim with ωrt obtained using (O.45) and the
274

approximation (O.46) for the 3hp, 220V motor, where the speed range considered

corresponds to 125% - 25% of the rated motor loading level.

ηim = λef f −1 (ωrt )λef f −2 (O.46)

where, λef f −1 , λef f −2 - efficiency coefficients, which are constants for a given mo-

tor. Table O.4 gives these efficiency coefficients for the range of induction motors

considered earlier. Thus, the power flow constraint given by (O.43) can be modified

employing (O.44) and (O.46) as:

λef f −1 (ωrt )λef f −2 (Pim:a + Pim:b + Pim:c ) = Pload (O.47)

Collecting all related equations together, the complete load flow formulation rep-

resenting the behaviour of a three-phase induction motor can be given by:


c
f1 (Ia:im , Ib:im , Ic:im , Va:k , Vb:k , Vc:k , ωrt ) = Va Ia:im − (Pim:a + jQim:a ) = 0
c
f2 (Ia:im , Ib:im , Ic:im , Va:k , Vb:k , Vc:k , ωrt ) = Vb Ib:im − (Pim:b + jQim:b ) = 0
c
f3 (Ia:im , Ib:im , Ic:im , Va:k , Vb:k , Vc:k , ωrt ) = Vc Ic:im − (Pim:c + jQim:c ) = 0

f4 (Ia:im , Ib:im , Ic:im , Va:k , Vb:k , Vc:k , ωrt ) =

λef f −1 (ωrt )λef f −2 (Pim:a + Pim:b + Pim:c ) − Pload = 0

n
The model requires the nominal power components Px−xx - Qnx−xz , speed coefficients

αsp:p−xx - αsp:q−xz , efficiency coefficients λef f −1 , λef f −2 and mechanical loading char-

acteristics as inputs. Estimation of these input parameters is an additional task

involved with this model in comparison to the impedance type model. However, this

model makes the load flow formulation and the derivation of the jacobian matrix

much simpler, especially in the case where the condition of balanced voltage angles

can be applied.
275

Table O.4: Efficiency coefficients for a range of induction motors


Motor specification
3hp, 220V 50hp, 460V 500hp, 2.3kV 2250hp, 2.3kV
λef f −1 95.776 99.135 99.565 99.824
λef f −2 0.9019 1.2673 2.2260 2.1890

97 Approximation: y = 95.776x0.9019
Actual variation
95

93
ηim (%)

91

89

87

85
0.93 0.94 0.95 0.96 0.97 0.98 0.99 1.00
ωrt (pu)

Figure O.13: Variation of ηim with ωrt for the 3hp, 220V motor
276

Comparison of the Proposed Impedance Type and PQ Type

Induction Motor Models

Cases 1 - 3 listed below demonstrate that the PQ type model behaves almost similar

to the impedance type model, although it incorporates some approximations9 and

assumptions10 .

• Case 1 - demonstration of the impact of the approximations (O.35) and (O.46)

incorporated with the PQ type model

Fig. O.14 illustrates the variation of the per phase input active and reactive

power with the motor loading level (given as a % to the rated motor load) for

the 3hp, 220V motor excited at the rated voltage (balanced).

• Case 2 - demonstration of the impact of the assumption, incorporated with

the PQ type model, of the motor efficiency is being independent of the supply

voltage condition

Fig. O.15 illustrates the variation of the per phase input active and reactive

power components with the motor loading level (given as a % to the rated

motor load) for the 3hp, 220V motor excited at reduced (|V+:k | = 0.9pu) and

unbalanced (V U F = 3%) voltages11 .

• Case 3 - demonstration of the validity of the PQ type model for large motors

Fig. O.16 illustrates the variation of the per phase input active and reactive

power components with the motor loading level (given as a % to the rated motor

load) for a 2250hp, 2.3kV motor [15] excited at reduced (|V+:k | = 0.9pu) and

unbalanced (V U F = 3%) voltages12 .


9
See (O.35) and (O.46).
10
The motor efficiency is being independent of the supply voltage condition.
11
Where,Va:k = 0.84∠00 pu, Vb:k = 0.92∠ − 1200 pu, Vc:k = 0.92∠1200 pu.
12
Where, Va:k = 0.84∠20 pu, Vb:k = 0.92∠ − 1200 pu, Vc:k = 0.92∠1200 pu. Note that the supply
voltage associates also with an angle unbalance.
277

1200 Pim:a - impedance type


Pim:a - PQ type
1000

Pim:a (W), Q im:a (VAr)


Qim:a - impedance type
800 Qim:a - PQ type

600

400

200

0
20 40 60 80 100 120
Loading level ( %)

Figure O.14: Variation of the per phase input active and reactive power with the
motor loading level for the 3hp, 220V motor excited at the rated voltage (balanced)

1400

1200 Pim:a - impedance type


Pim:a - PQ type
Pim:b - impedance type
1000
Pim:x (W), Qim:x (VAr)

Pim:b - PQ type
Pim:c - impedance type
800 Pim:c - PQ type
Qim:a - impedance type
600 Qim:a - PQ type
Qim:b - impedance type
Qim:b - PQ type
400
Qim:c - impedance type
Qim:c - PQ type
200

0
20 40 60 80 100 120

Loading level (%)

Figure O.15: Variation of the per phase input active and reactive power components
with the motor loading level for the 3hp, 220V motor excited at reduced and unbal-
anced voltages
278

0.9

0.8 Pim:a - impedance type


Pim:x (MW), Qim:x (MVAr) Pim:a - PQ type
0.7
Pim:b - impedance type
0.6 Pim:b - PQ type
Pim:c - impedance type
0.5
Pim:c - PQ type
0.4 Qim:a - impedance type
Qim:a - PQ type
0.3 Qim:b - impedance type
Qim:b - PQ type
0.2
Qim:c - impedance type
0.1 Qim:c - PQ type

0.0
20 40 60 80 100 120
Loading level (%)

Figure O.16: Variation of the per phase input active and reactive power components
with the motor loading level for a 2250hp, 2.3kV motor excited at reduced and un-
balanced voltages

kp:rated = 0.87
1000

800
Pim:a (W)

600 Proposed impedance type model


Proposed PQ type model
400 Mode given in [4,5]
Mode given in [11,12]
200

0
0.15 0.4 0.65 0.9 1.15
kp

Figure O.17: Variation of Pim:a with kp for the existing and proposed induction motor
models
279

Comparison of the existing and proposed models

Figs. O.17 and O.18 illustrate the variation of the input real and reactive power

respectively for the phase a with kp of pump systems obtained employing the ex-

isting and proposed models in relation to the 3hp, 220V motor excited at reduced

(|V+:k | = 0.9pu) and unbalanced (V U F = 3%) voltages13 . These demonstrate that

the model discussed in [11, 12]14 behaves similar to the proposed models only when

the motor supplies the rated mechanical load15 . The model given in [4, 5]16 does not

represent the motor behaviour accurately even when the motor is loaded with the

rated mechanical load17 .


kp:rated = 0.87

450
Qim:a (VAr)

350

Proposed impedance type model

250 Proposed PQ type model


Model given in [4,5]
Model given in [11,12]
150
0.15 0.4 0.65 0.9 1.15
kp

Figure O.18: Variation of Qim:a with kp for the existing and proposed induction motor
models

13
Where, Va:k = 0.84∠00 pu, Vb:k = 0.92∠ − 1200 pu, Vc:k = 0.92∠1200 pu.
14
This constraints the real power drawn by the motor at the rated value, and allows the reactive
power to be adjusted by the load flow algorithm according to the existing supply voltage condition.
15
Note the deviation in both Pim:a and Qim:a associated with the model given in [11, 12] compared
to the proposed models when kp 6= kp:rated .
16
This constraints both the real and reactive power drawn by the motor at the rated values.
17
Note the deviation in Qa at kp:rated associated with the model given in [4, 5] compared to the
other models.
280

O.4.7 Network Interactions

With the representation of individual system components as described above, the

interaction between them are obtained using the component branch currents by:

[Ykm ][Vk ] + [Ik ] = 0 (O.48)

where,

[Ykm ] - matrix of nodal admittances

[Vk ] - matrix of nodal voltages

[Ik ] - matrix of nodal currents

Ykm - represents any element (a 3 × 3 matrix) of the matrix [Ykm ], where:

• for k = m, Ykm is equal to the summation of all phase admittance matrices

(3 × 3) connected to the busbar k (= m),

• for x 6= y, Yxy is equal to the negative value of the phase admittance matrix

(3 × 3) that exists between the busbars k and m

Vk = [Va:k Vb:k Vc:k ]t , any element of the matrix [Vk ]

Ik = [Ia:k Ib:k Ic:k ]t , any element of the matrix [Ik ], which is equal to the summation

of the phase current matrices (3 × 1) correspond to individual system components

(e.g. generator and load currents) connected at the busbar k

O.5 Load Flow Solution

The load flow equations are solved employing the well established Newton-Raphson

iterative technique (full version).


281

O.6 Related References

1. Xiao-Ping Zhang and Heng Chen. Sequence-Decoupled Newton-Raphson Three-

phase Load Flow. In Proc. IEEE Region 10 Conference on Computer, Com-

munication, Control and Power Engineering (TENCON ’93), volume 5, issue

0, part 50000, pages 394 - 397, 19-21 October 1993.

2. M. A. Laughton. Analysis of Unbalanced Poly-phase Networks by the Method

of Phase Co-ordinates. In IEE Proc., volume 115, no. 8, August 1968.

3. M. Abdel-Akher, K.M. Nor, and A.H.A. Rashid. Improved Three-phase Power

Flow Methods using Sequence Components. IEEE Trans. on Power Systems,

20(3):1389-1397, August 2005.

4. W. Xu, H. W. Dommel, and J. R. Marti. A Generalised Three-phase Power

Flow Method for the Initialisation of EMTP Simulations. In Proc. Interna-

tional Conference on Power System Technology (POWERCON ’98), volume 2,

pages 875 - 879, 18 - 21 August 1998.

5. W. Xu, J. R. Marti, and H. W. Dommel. A Multiphase Harmonic Load Flow

Solution Technique. IEEE Trans. on Power Systems, 6(1):174-182, Febuary

1991.

6. Prabha Kundur. Power System Stability and Control. McGraw-Hill, 1993.

7. P.A.N. Garcia, J.L.R. Pereira, and S.Jr. Carneiro. Voltage Control Devices

Models for Distribution Power Flow Analysis. IEEE Trans. on Power Systems,

16(4):586-594, November 2001.

8. Paul M. Anderson. Analysis of Faulted Power Systems. John Wiley & Sons,

1995.
282

9. J. C. Das. Power System Analysis - Short-Circuit Load Flow and Harmonics.

Marcel Dekker, 2002.

10. Yuin-Hong Liu, Wei-Jen Lee, and Mo-Shing Chen. Incorporating Induction Mo-

tor Model in a Load Flow Program for Power System Voltage Stability Study.

In Proc. IEEE International Conference on Electric Machines and Drives, vol-

ume 2, 18 - 21 May 1997.

11. D.R.R. Penido, L.R. Araujo, S. Carneiro Jr., and J.L.R. Pereira. Unbalanced

Three-phase Distribution System Load-flow Studies Including Induction Ma-

chines. In Proc. IEEE Power Engineering Society General Meeting, 18 -

22 June 2006.

12. Thomas E. McDermott. Radial Distribution Feeder and Induction Machine

Test Cases - Steady State Solutions. In Proc. IEEE Power Engineering Society

General Meeting, 18 - 22 June 2006.

13. R.C. Dugan. Induction Machine Modeling for Distribution System Analysis

- Test Case Description. In Proc. IEEE Power Engineering Society General

Meeting - Transmission and Distribution, 21 - 24 May 2006.

14. Daniel Ruiz-Vega, Tomás I. Asiaı́n Olvares, and Daniel Olguı́n Salinas. An

Approach to the Initialization of Dynamic Induction Motor Models. IEEE

Trans. on Power Systems, 22(6), June 2006.

15. Paul C. Krause, Oleg Wasynczuk, and Scott D. Sudhoff. Analysis of Electric

Machinery and Drive Systems. John Wiley and Sons, 2002.

S-ar putea să vă placă și