Technical Brochure: Network Modelling For Harmonic Studies
Technical Brochure: Network Modelling For Harmonic Studies
Technical Brochure: Network Modelling For Harmonic Studies
Technical Brochure
Contributing Members
Reviewers
J. GING IE M. HALPIN US
R. A. WALLING US M. BOLLEN SE
J. A. R. MONTEIRO UK R. BERES DK
I. IYODA JP P. WANG CA
C. LETH BAK DK
Copyright © 2019
“All rights to this Technical Brochure are retained by CIGRE. It is strictly prohibited to reproduce or provide this publication in any
form or by any means to any third party. Only CIGRE Collective Members companies are allowed to store their copy on their
internal intranet or other company network provided access is restricted to their own employees. No part of this publication may
be reproduced or utilized without permission from CIGRE”.
Disclaimer notice
“CIGRE gives no warranty or assurance about the contents of this publication, nor does it accept any responsibility, as to the
accuracy or exhaustiveness of the information. All implied warranties and conditions are excluded to the maximum extent permitted
by law”.
WG XX.XXpany network provided access is restricted to their own employees. No part of this publication may be
reproduced or utilized without permission from CIGRE”.
Disclaimer notice
“CIGRE gives no warranty or assuranceISBN : 978-2-85873-468-9
about the contents of this publication, nor does it accept any
responsibility, as to the accuracy or exhaustiveness of the information. All implied warranties and
conditions are excluded to the maximum extent permitted by law”.
TB 766 - Network modelling for harmonic studies
ISBN : 978-2-85873-468-9
4
TB 766 - Network modelling for harmonic studies
Executive summary
The issue of harmonic distortion in power systems is not new, with publications dating as early as 1916
dealing with distorted waveforms in transmission lines and their effects on transformers, rotating
machines and telephone interference. Significant efforts were devoted then to investigate, understand
and mitigate their negative effects. Standardisation was introduced to control and limit the amount of
harmonic distortion present in the system. The focus was on large non-linear installations, such as
Electric Arc Furnaces (EAF), smelters, industrial converters, SVCs or HVDC systems, which were
subject to harmonic limits and mitigated their emissions using filters where necessary. As a
consequence, harmonic distortion was not a source of practical problems in most transmission power
systems for many years.
This picture is changing. Power systems globally are experiencing a transition towards decarbonisation
of electricity production through large-scale deployment of renewable energy sources (RES), which are
gradually displacing conventional thermal plant. The connection of RES to the electricity network is
mostly achieved through the use of Power Electronic (PE) converters, which are sources of harmonic
distortion.
It has been observed that many power systems are experiencing an increase in harmonic distortion.
Drivers for this trend include not only the integration of RES, but also increased connection of FACTS
devices, HVDC systems, HVAC cables and general proliferation of PE converters in the demand (e.g.
electric vehicles and domestic appliances). Power quality issues associated with harmonics in power
systems are therefore becoming more pronounced and are driving a new focus towards the need to
undertake detailed analysis at the planning stages in order to ensure adherence to statutory limits.
Performing harmonics studies requires a more advanced skill set and more detailed system models and
simulation tools than those required for traditional planning studies (e.g. load flow and short-circuit
analysis). In general, there is some lack of knowledge when it comes to conducting meaningful harmonic
distortion assessments in modern power systems. Most of the issues stem from the lack of practical
information on modelling electrical plant equipment for such studies. Availability of information and
guidance for such modelling requirements are either scarce or in scattered form, mostly delegated to
appendices of various documents (examples are the short articles in Electra 164 [1] and 167 [2],
published in 1996). Both Electra articles are authoritative in nature but are becoming outdated and
inaccessible. Previous CIGRE publications have tried to bridge the gap to a certain extent but the need
for such studies has increased tremendously in most parts of the world, and hence the need for up-to-
date information on the topic of modelling. In response to this increased interest and to review state-of-
the-art modelling and simulation practices, CIGRE Study Committees C4 and B4 established the Joint
Working Group (JWG) C4/B4.38: “Network Modelling for Harmonic Studies” in late 2014.
This Technical Brochure has been compiled drawing expertise from the JWG members and provides
comprehensive guidelines for practising power system engineers when they need to perform harmonic
distortion assessments. The document covers the modelling of the most common network components
and discusses key features that need to be considered in the assessments. The focus is on practical
aspects of modelling for direct application in the planning process of connecting a new customer to the
power system, or when introducing a change to the system as part of asset replacement or system
expansion. As such, the guidelines are concentrated on frequency-domain modelling for steady-state
AC harmonic analysis in transmission and distribution networks, typically in the range from power
frequency up to the 50th harmonic (2.5 kHz in 50 Hz systems or 3 kHz in 60 Hz systems), consistent
with typical power quality assessments. The approach and modelling guidelines provided are
reasonably valid up to the 100th harmonic if specialized studies are required. These guidelines will be
valuable in the definition of harmonic performance specifications for new HVDC converters, FACTS
devices or other non-linear installations. They will also assist connectees when modelling their
installation to assess or demonstrate compliance with the emission limits provided by the System
Owner/Operator and to investigate and specify mitigation measures such as harmonic filters.
Furthermore, this Technical Brochure can also be used post-commissioning for any incident
investigation or to assist resolution of customer complaints via modelling and analysis.
This Technical Brochure has summarised the state-of-the-art and reflects best practice in modelling for
harmonics studies in power systems. Areas where further research work is required have also been
identified. These include: (i) improved load modelling with focus on PE-based load, e.g. load
characterisation and aggregation; (ii) validation of power converter models with field measurements,
e.g. methods to separate the effect of system harmonic impedance and background distortion from the
converter emission; (iii) summation of harmonic sources, i.e. gather enough information, based on field
5
TB 766 - Network modelling for harmonic studies
measurements and theoretical analysis, to propose robust and realistic alternatives to improve the IEC
summation law currently adopted in most harmonics analysis; (iv) improved methods for accurate
aggregation of wind farm components (wind turbine generators, transformers, cables, etc) into a single
frequency dependent equivalent; and (v) develop practical methods for the estimation of background
distortion in meshed network topologies where availability of measurements is limited.
6
TB 766 - Network modelling for harmonic studies
Table of contents
Executive summary ............................................................................................................. 5
1. Introduction.............................................................................................................. 17
1.1 Background.............................................................................................................................................. 17
1.2 Scope ........................................................................................................................................................ 18
1.3 Structure................................................................................................................................................... 19
7
TB 766 - Network modelling for harmonic studies
6. Conclusions ............................................................................................................191
7. Bibliography/references .........................................................................................195
8
TB 766 - Network modelling for harmonic studies
9
TB 766 - Network modelling for harmonic studies
Figure 3-43 Downstream Harmonic Component Load Model Representation at Supply Bus GSP as used by
Scottish Power TSO .............................................................................................................................................. 76
Figure 3-44 Test Case .......................................................................................................................................... 77
Figure 3-45 Modelling of 25kV Feeder with Various Load Levels ......................................................................... 79
Figure 3-46 Model Comparison: Feeder with Heavy Load ................................................................................... 80
Figure 3-47 Model Comparison: Feeder with Light Load ...................................................................................... 81
Figure 3-48 Frequency Response as seen from the Wind Farm (120kV) ............................................................. 82
Figure 3-49 Model Comparison as seen from the Wind Farm (120kV) ................................................................. 83
Figure 3-50 Effect of Load Proximity to Point of Interconnection .......................................................................... 84
Figure 3-51 Extended CIGRE 14-Bus System for Comparison of Harmonic Load Models .................................. 85
Figure 3-52 Impedance at SB4 220kV – Winter Load .......................................................................................... 86
Figure 3-53 Impedance at SB4 220kV – Autumn Load ........................................................................................ 87
Figure 3-54 Impedance at SB4 220kV – Summer Load ....................................................................................... 87
Figure 3-55 Impedance at Bus1 110kV – Winter Load ......................................................................................... 88
Figure 3-56 Impedance at Bus1 110kV – Autumn Load ....................................................................................... 88
Figure 3-57 Impedance at Bus1 110kV – Summer Load ...................................................................................... 88
Figure 3-58 Impedance at Bus2 33kV – Winter Load ........................................................................................... 89
Figure 3-59 Impedance at Bus2 33kV – Autumn Load ......................................................................................... 89
Figure 3-60 Impedance at Bus2 33kV – Summer Load ........................................................................................ 90
Figure 3-61 Impedance at Bus3 33kV – Winter Load ........................................................................................... 90
Figure 3-62 Impedance at Bus3 33kV – Autumn Load ......................................................................................... 91
Figure 3-63 Impedance at Bus3 33kV – Summer Load ........................................................................................ 91
Figure 3-64 Impedance at Bus4 33kV – Winter Load ........................................................................................... 92
Figure 3-65 Impedance at Bus4 33kV – Autumn Load ......................................................................................... 92
Figure 3-66 Impedance at Bus4 33kV – Summer Load ........................................................................................ 92
Figure 3-67 Impedance at Bus5 33kV – Winter Load ........................................................................................... 93
Figure 3-68 Impedance at Bus5 33kV – Autumn Load ......................................................................................... 93
Figure 3-69 Impedance at Bus5 33kV – Summer Load ........................................................................................ 94
Figure 3-70 Synchronous Generator Harmonic Impedance ................................................................................. 95
Figure 3-71 Sample 370 MVA Synchronous Generator Harmonic Equivalent Resistance, R genh, and Reactance,
Xgenh ....................................................................................................................................................................... 96
Figure 3-72 Model 1: IEEE ................................................................................................................................... 96
Figure 3-73 Model 2: Electra 167 ......................................................................................................................... 97
Figure 3-74 Model 3: HQ ...................................................................................................................................... 97
Figure 3-75 Effect of Detailed Synchronous Generator Models on Harmonic Impedance in a Real Transmission
System (a) Close to Generation and (b) Far from Generation ............................................................................... 98
Figure 3-76 Maximum Damping Factor (with Respect to Base Case) Provided by each Synchronous Generator
Model ..................................................................................................................................................................... 99
Figure 3-77 Detuned Shunt Capacitors .............................................................................................................. 100
Figure 3-78 Effect of connecting Shunt Capacitors in a Transmission Network ................................................. 101
Figure 3-79 Examples of Typical Filter Configurations ....................................................................................... 102
Figure 4-1 Norton/Thévenin Harmonic Model of Power Electronic Converter (source [62]) ............................... 105
Figure 4-2 Doubly-Fed Induction Generator [65] ............................................................................................... 108
Figure 4-3 Example Type 4 Wind Turbine Generator ......................................................................................... 109
Figure 4-4 Example Comparison of WT Harmonic Impedance and Smoothing Reactor of Converter ............... 111
Figure 4-5 WTG Modelling Approaches ............................................................................................................. 112
Figure 4-6 Frequency Scan of a Wind Farm with different WTG Harmonic Models ........................................... 113
Figure 4-7 Typical PV System Overview ............................................................................................................ 113
Figure 4-8 AC Bank Switching Steps following HVDC LCC Converter Q Curve................................................. 116
Figure 4-9 Illustrative Variation of Characteristic Harmonics Magnitude vs Load (HVDC LCC) ......................... 117
Figure 4-10 Filter Bank, Sub-Bank and Branch Definition .................................................................................. 118
Figure 4-11 Harmonic Filter Impedance - Minimum DC Power Configuration .................................................... 119
Figure 4-12 Harmonic Filter Impedance - Triple-Tuned Filter ............................................................................. 119
Figure 4-13 Harmonic Filter Impedance - Maximum DC Power Configuration ................................................... 120
Figure 4-14 Simplified Equivalent Circuit for Harmonic Emission Assessment at PCC (HVDC LCC) ................ 122
Figure 4-15 Simplified Circuit to Assess Magnification (or Damping) of Background Voltage Distortion at PCC
(HVDC LCC) ........................................................................................................................................................ 123
Figure 4-16 HVDC LCC Model for System-Wide Studies ................................................................................... 124
Figure 4-17 a) MMC VSC Configuration, b) Half-Bridge Module (c) Full-Bridge Module .................................... 125
Figure 4-18 Simplified Equivalent Circuit for Harmonic Emission Assessment at PCC (HVDC VSC) ................ 127
Figure 4-19 Simplified Circuit to Assess Magnification (or Damping) of Background Voltage Distortion at PCC (VSC
HVDC) ................................................................................................................................................................. 128
Figure 4-20 HVDC VSC Model for System-Wide Studies ................................................................................... 129
Figure 4-21 HVDC Modelling Guidelines for Harmonic Studies .......................................................................... 130
Figure 4-22 SVC Typical Configuration .............................................................................................................. 132
Figure 4-23 Typical Harmonic Currents Generated by a TCR [87] ..................................................................... 133
Figure 4-24 Simplified circuit for Harmonic Emission Assessment at PCC (SVC) .............................................. 134
Figure 4-25 Simplified Circuit to Assess Magnification (or Damping) of Background Voltage Distortion at PCC (SVC)
............................................................................................................................................................................ 135
10
TB 766 - Network modelling for harmonic studies
App Figure A.1 Phase Amplification as Function of the Harmonic Order and (a) the Relative Permittivity and (b) the
Cable System Length .......................................................................................................................................... 204
App Figure A.2 Phase and Sequence Voltage in an Inter-Sequence Coupled Meshed Transmission Network . 205
App Figure A.3 Frequency Scans in Inter-Sequence Coupled Meshed Transmission Network at Two Unique
Substations .......................................................................................................................................................... 205
App Figure B.1 220/110 kV Transmission Network Topology............................................................................. 207
App Figure C.1 Sequential Steps followed in the 220 kV Network Model Development ..................................... 209
App Figure C.2 Harmonic Impedance at Bus-5 under Incremental 220kV Network Model Development .......... 211
App Figure C.3 Setup for Calculation of Harmonic Impedance at Bus 5 including 110 kV System .................... 212
App Figure C.4 Harmonic Impedance at Bus-5 with 220kV and 110kV Network Models ................................... 213
App Figure D.1 Cable Layers ............................................................................................................................. 215
11
TB 766 - Network modelling for harmonic studies
Tables
Table 2-1 Summary of Study Domains and their Applications ............................................................................... 28
Table 3-1 Corrections for Skin Effect in OHLs ....................................................................................................... 34
Table 3-2 Variation of Earth Resistivity with Depth ............................................................................................... 35
Table 3-3 Summary of Sensitivity Analysis: Impacts on Harmonic Impedance ..................................................... 43
Table 3-4 Summary of Sensitivity Analysis in 0: Impacts on Harmonic Impedance............................................... 50
Table 3-5 Values for Coefficients 𝐚𝟎, 𝐚𝟏 𝐚𝟐 and 𝐛 (requirement: 𝐚𝟎 + 𝐚𝟏 + 𝐚𝟐 = 𝟏) ............................................ 54
Table 3-6 Coefficients for Transformer Model 5 [50] ............................................................................................. 55
Table 3-7 Example 1: Transformer Model Parameters ......................................................................................... 56
Table 3-8 Example 2: Transformer Model Parameters ......................................................................................... 59
Table 3-9 Typical values of RLV and PFC *courtesy National Grid ....................................................................... 75
Table 3-10 Typical BLV (Power Factor Correction (PFC) + Cable Capacitance) and Equivalent XT Transformer
Reactance *courtesy National Grid ........................................................................................................................ 75
Table 3-11 Typical Cable Capacitance, BHV ......................................................................................................... 76
Table 3-12 Typical values of RLV for Different Load Categories .......................................................................... 76
Table 3-13 Composition of 25 kV Load Feeder 235 ............................................................................................. 77
Table 3-14 Residential-Commercial Assumed Load Compositions ....................................................................... 78
Table 3-15 Distribution-Level Circuits .................................................................................................................... 86
Table 3-16 Load Profiles Modelled ........................................................................................................................ 86
12
TB 766 - Network modelling for harmonic studies
Table 4-1 Example Representation/Template of the Harmonic Voltage/Current Source ..................................... 106
Table 4-2 Example Representation/Template of the Harmonic Equivalent Impedance ....................................... 106
Table 4-3 Comparison of Harmonic Performance in Typical FACTS Devices .................................................... 131
Table 4-4 Typical Harmonic Mitigation Solutions in FACTS Devices .................................................................. 131
Table 4-5 Modelling Considerations for Line-Commutated FACTS Devices ...................................................... 138
Table 4-6 Modelling Considerations for Self-Commutated FACTS Devices ....................................................... 142
Table 5-1. IEC 61000-3-6 Summation Exponents for Harmonics [100] ............................................................... 175
Table 5-2: Comparison of Different Statistic Values of Measurements of 7 th Harmonic of Two Different Weeks in
Germany .............................................................................................................................................................. 177
Table 5-3: Comparison of Different Statistical Values of Measurements of 7 th Harmonic for Two Different Months
in Germany .......................................................................................................................................................... 178
Table 5-4 Maximum and Minimum 95th Percentile Levels of 11th and 13th Harmonics in a 400 kV Danish Substation
over 20 Weeks..................................................................................................................................................... 179
Table 5-5 Ratios of the Maximum and Minimum 95th Percentile Levels of 11 th and 13th Harmonics in a 400 kV
Danish Substation over 20 Weeks....................................................................................................................... 179
Equations
Equation 2.1 .......................................................................................................................................................... 22
Equation 3.1 .......................................................................................................................................................... 33
Equation 3.2 .......................................................................................................................................................... 35
Equation 3.3 .......................................................................................................................................................... 41
Equation 3.4 .......................................................................................................................................................... 52
Equation 3.5 .......................................................................................................................................................... 53
Equation 3.6 .......................................................................................................................................................... 53
Equation 3.7 .......................................................................................................................................................... 53
Equation 3.8 .......................................................................................................................................................... 53
Equation 3.9 .......................................................................................................................................................... 53
Equation 3.10 ........................................................................................................................................................ 53
Equation 3.11 ........................................................................................................................................................ 53
Equation 3.12 ........................................................................................................................................................ 53
Equation 3.13 ........................................................................................................................................................ 54
Equation 3.14 ........................................................................................................................................................ 54
Equation 3.15 ........................................................................................................................................................ 54
Equation 3.16 ........................................................................................................................................................ 54
Equation 3.17 ........................................................................................................................................................ 55
Equation 3.18 ........................................................................................................................................................ 55
Equation 3.19 ........................................................................................................................................................ 67
Equation 3.20 ........................................................................................................................................................ 67
Equation 3.21 ........................................................................................................................................................ 67
Equation 3.22 ........................................................................................................................................................ 68
Equation 3.23 ........................................................................................................................................................ 68
Equation 3.24 ........................................................................................................................................................ 68
Equation 3.25 ........................................................................................................................................................ 69
Equation 3.26 ........................................................................................................................................................ 69
Equation 3.27 ........................................................................................................................................................ 69
Equation 3.28 ........................................................................................................................................................ 69
Equation 3.29 ........................................................................................................................................................ 69
Equation 3.30 ........................................................................................................................................................ 70
Equation 3.31 ........................................................................................................................................................ 70
Equation 3.32 ........................................................................................................................................................ 70
Equation 3.33 ........................................................................................................................................................ 70
Equation 3.34 ........................................................................................................................................................ 70
Equation 3.35 ........................................................................................................................................................ 70
Equation 3.36 ........................................................................................................................................................ 70
Equation 3.37 ........................................................................................................................................................ 71
Equation 3.38 ........................................................................................................................................................ 71
Equation 3.39 ........................................................................................................................................................ 73
Equation 3.40 ........................................................................................................................................................ 73
Equation 3.41 ........................................................................................................................................................ 73
Equation 3.42 ........................................................................................................................................................ 74
Equation 3.43 ........................................................................................................................................................ 74
13
TB 766 - Network modelling for harmonic studies
14
TB 766 - Network modelling for harmonic studies
HV High Voltage
HVDC High Voltage Direct Current
IGBT Insulated-Gate Bipolar Transistor
IGCT Integrated Gate Commutated Thyristor
LCC Line Commutated Converter
LCI Line Commutated Inverter
LV Low Voltage
MMCC Modular Multi-level Cascaded Converter
MoM Method of Moments
MSC Mechanically Switched Capacitor
MSCDN Mechanically Switched Capacitor with Damping Network
MV Medium Voltage
OEM Original Equipment Manufacturer
OHL Overhead Line
PCC Point of Common Coupling
PoC Point of Connection
PE Power Electronics
PFC Power Factor Correction
PLL Phase-Locked Loop
PQ Power Quality
PV PhotoVoltaic
PWM Pulse Width Modulation
RES Renewable Energy Sources
SAF Submerged Arc Furnace
SC Single Core
SCR Silicon Controlled Rectifier (Thyristor)
SHE Selective Harmonic Elimination
SM Sub-Module
SRF-PLL Synchronous Rotating Frame PLL
SSC Static Shunt Capacitor
SSSC Solid State Series Compensator
STATCOM Static Synchronous Compensator
SVC Static Var Compensator
TCR Thyristor Controlled Reactor
THD Total Harmonic Distortion
TO (TSO) Transmission System Operator
TSC Thyristor Switched Capacitor
TSR Thyristor Switched Reactor
UGC UnderGround Cable
UHV Ultra High Voltage
15
TB 766 - Network modelling for harmonic studies
16
TB 766 - Network modelling for harmonic studies
1. Introduction
1.1 Background
Harmonics have been part of the electric power network since their introduction into the system by
nonlinear devices. They occur as emissions, usually as an undesirable side-effect of the main function
of equipment, at multiple or sub-multiple integers of the system frequency. It is also possible to define
inter-harmonics as those frequencies which appear at non-integer multiples of the system frequency.
The issues caused by harmonics can be categorised under two distinct but heavily-interrelated areas:
steady-state distortion and transient effects. The first area relates to steady-state harmonic voltage
distortion, where system resonance or resonances are excited via the introduction of harmonic injections
by some active devices. Typical examples of active devices are power electronics being utilised in
converter or inverter circuits. These devices are normally under continuous operation and hence are a
source of either harmonic current or voltage which creates a steady-state distortion on the system. The
second area relates to the excitation of system resonances by harmonic sources that are transient in
nature. The most typical case is the energization of a transformer. When a transformer is energized by
uncontrolled or random closing of the energizing device, the magnetic characteristic of the transformer
core results in it being driven into partial or full saturation, and hence a current rich in harmonics is
drawn. This current could excite a system resonance thereby creating a temporary overvoltage. The
phenomenon is short in duration and in most cases the level of overvoltage reached is not excessively
high. This Technical Brochure (TB) concentrates on aspects related to steady-state harmonic distortion.
The adverse effects of harmonics have been discussed in many publications (e.g. [3], [4]) and will
therefore not be treated in detail in this introduction. However, it suffices to say that harmonics could
have detrimental effects on electrical equipment via the introduction of additional heating or the creation
of a complex environment for protection or measuring devices to function properly. Therefore,
standardisation exists in order to control and limit the amount of harmonic distortion present in the
system. This is done using limits for current or voltage distortion. These limits are normally coordinated
across different voltage levels to have a graded approach and are usually two-tier (i.e. planning and
compatibility limits) depending on which standard is utilised. Planning limits are usually classified as the
internal objectives for utilities, i.e. to plan the system such that these levels are not exceeded.
Compatibility limits on the other hand are the absolute levels that should not be breached as they define
the safe operating environment of all connected equipment.
To keep the harmonic distortion levels within planning limits, utilities need to perform extensive studies
or assessments at the planning stage of any changes being introduced to the electric power network.
The changes could be due to the expansion of the system itself, such as the installation of a new cable
route, or the integration of a new non-linear installation.
Integration of large renewable energy sources, and in particular those that involve offshore connection
requiring the use of long HVAC cables, have a twofold effect. Not only do the wind turbine generators
introduce harmonic current emissions into the system, but the integration of long HVAC cables moves
system resonance frequencies further down in the frequency spectrum and hence has an amplifying
effect on the low-order harmonics that are already present on the system. The introduction of FACTS
devices and HVDC converter technology, be it voltage or current source, also brings further harmonic
sources onto the system and hence the requirement to control them. Harmonic sources can excite local
resonances but also remote resonances in the power system, which are not detectable locally at the
injection point. Therefore, the need for performing harmonic assessments is becoming even more
pronounced as the connection of nonlinear devices and/or loads increases, driven by integration of
renewable energy sources and connection of new HVDC converters.
Adequate harmonic performance specification is therefore crucial as it impacts the subsequent design
of harmonic filters with associated costs and risks. In order to draw up proper specifications or limitations
to coordinate the emission of harmonics onto the system, the vast majority of cases require proper
modelling of the power system for this purpose. The assessment can take different forms according to
the local or international standard requirements, however in most cases network changes or connections
above 33kV will require some form of detailed system modelling.
All utilities usually have a model of their network that will replicate ordinary power flows and be able to
facilitate short-circuit calculations. In addition, they also have a dynamic model of their system in order
to perform stability calculations. However, many of them do not have a frequency-dependent system
model of their network for the purposes of harmonic distortion analysis. Traditionally, most utilities also
have some form of knowledge of the transient nature of harmonics acquired during insulation
17
TB 766 - Network modelling for harmonic studies
coordination studies. However, this type of study approach is limited by the extent of the model, and the
knowledge and applicability has usually been confined to a select few. This used to be an acceptable
situation in the past as the need to perform this type of harmonic distortion assessment was seldom.
For example, there was previously no need to quickly perform an assessment and specify emission
limits to the connectee in a commercial environment when different projects were competing for a
connection to the same system.
In general, there is some lack of knowledge when it comes to performing meaningful harmonic distortion
assessments in modern power systems. Most of the issues stem from the lack of practical information
on modelling electrical plant equipment for such studies. Equally, the parameters that need to be
considered and the extent of this consideration have not been explained for the purposes required by
the practising power system engineer. Availability of information and guidance for such modelling
requirements are either scarce or in scattered form, mostly delegated to appendices of various
documents (examples are the short articles in Electra 164 [1] and 167 [2], published in 1996). The source
of many of these models has been lost in history, and reference can only be made to publications, some
of them by CIGRE, which contain formulae for the representation of network components but with little
or no proof of derivation. Both Electra articles are authoritative in nature but suffer from the issues
outlined above and are also becoming outdated and inaccessible. Previous CIGRE publications have
tried to bridge the gap to a certain extent but the need for such studies has increased tremendously in
most parts of the world, and hence the need for up-to-date information on the topic of modelling.
The main objective of this Technical Brochure is to provide comprehensive guidelines for practising
power system engineers when they need to perform harmonic distortion assessments. The focus is on
practical aspects of modelling for direct application in the planning process of connecting a new
customer to the transmission or distribution system, or when introducing a change to the system as part
of asset replacement or system expansion. As such, the guidelines are concentrated on frequency-
domain modelling for steady-state AC harmonic analysis in power systems, typically in the range from
power frequency up to the 50th harmonic (2.5 kHz in 50 Hz systems or 3 kHz in 60 Hz systems),
consistent with typical power quality assessments. The approach and modelling guidelines provided are
reasonably valid up to the 100th harmonic if specialized studies are required. These guidelines will be
valuable in the definition of harmonic performance specifications for new HVDC converters, FACTS
devices or other non-linear installations. They will also assist connectees when modelling their
installation to assess or demonstrate compliance with the emission limits provided by the System
Operator and to investigate and specify mitigation measures such as harmonic filters. Furthermore, this
document can also be used post-commissioning for any incident investigation or to assist resolution of
customer complaints via modelling and analysis.
1.2 Scope
The scope of this Technical Brochure, as defined in the Terms of Reference for the Working Group
(JWG C4/B4.38) concerning system harmonic modelling, is to:
1. Collate and provide all available information in the literature on modelling individual electrical plant.
2. Evaluate and suggest best practice in the use of available models to represent modern equipment.
3. Identify any shortfalls with available models and the possible need for further development in this
area.
4. Provide clear and concise guidelines on modelling existing nonlinear devices (HVDC converter
stations, wind farm generation, etc.) within the system of interest.
5. Provide guidelines on the general approach to such studies and the availability and choice of tools.
Identify any shortfalls with the available analysis tools and suggest possible developments.
The focus of this Technical Brochure is on frequency-domain modelling for steady-state AC harmonic
analysis in electric power networks, typically in the range up to the 50th harmonic (2.5 kHz in 50 Hz
systems or 3 kHz in 60 Hz systems), consistent with typical power quality assessments. The approach
and modelling guidelines provided are reasonably valid up to the 100th harmonic if specialized studies
are required. The document is not intended to cover guidelines for transient issues under harmonic
resonance although most of the modelling guidelines given will equally apply to that area also. As such,
modelling for transient harmonic performance during transformer energization or geomagnetic
disturbances (GMD) is outside the scope of this document. Similarly, the topic of harmonics on the DC
side of converters is outside the scope of this Technical Brochure.
The control systems of power electronic converters directly connected to HV and EHV grids can interact
with system resonances leading to high magnitudes of harmonics in the grid. This phenomenon is
normally referred to as harmonic stability and the root cause is the interaction of a converter controller
18
TB 766 - Network modelling for harmonic studies
with a grid resonance. Converter controller harmonic stability requires different modelling, analysis and
mitigation methods from those described in this document. This topic is outside the scope of this
Technical Brochure. The reader is referred to the outcome of the recent CIGRE Working Group B4.67
(TB754) [5] for information on this phenomenon. Furthermore, the recently established CIGRE Working
Group C4.49 (Multi-frequency stability of converter-based modern power systems) is currently
addressing this topic in detail.
1.3 Structure
This Technical Brochure contains a further five chapters and several appendices. The appendices
include a benchmark system and further detailed analyses and examples to complement the guidelines
and recommendations provided in each of the chapters.
Chapter 2 – Study Domain and Modelling Approaches – provides an overview of analysis techniques
and solution methods for the study of harmonic distortion in power systems. This chapter examines the
strengths and limitations of each technique, including balanced and unbalanced solution methods, and
provides recommendations on applicability based on the purpose of the analysis. State-of-the-art and
future trends in solution techniques are also discussed.
Chapter 3 – Classical Network Element Models – reviews and discusses modelling options and
approaches for the accurate representation of most relevant passive network components in harmonics
studies. These include overhead lines, insulated cables, transformers, loads, synchronous generators
and shunt/series compensation devices. The relevant input data and level of detail required to represent
each component are presented and discussed. Different modelling options are assessed in a benchmark
model as well as in real system models to illustrate the effects and consequences of each type of model
in the context of system-wide studies. When available, models are validated against measurements and
recommendations are provided.
Chapter 4 – Power Electronics Based Network Element Models – reviews and discusses the
representation of non-linear devices connected to power systems acting as sources of harmonic
distortion. This chapter covers converter-based generation, HVDC converters, FACTS devices, traction
systems, electric arc furnaces and variable speed drives. The chapter reviews the mechanisms of
harmonic distortion generation for each device and provides recommendations on the modelling
structure and input data requirements. The harmonic performance of these devices is generally complex
and dependent on many factors such as converter topology, control strategy, operating point, etc. As
such, manufacturer-specific models need to be adopted. Recommendations on the structure of the
models and the features that need to be captured are provided. Emphasis is placed on the need to
capture not only the harmonic current/voltage emissions but also the harmonic impedance of the
devices.
Chapter 5 – General Considerations for Harmonic Studies – discusses practical aspects, other than the
modelling of each component, that need to be considered when performing harmonic studies related to
the connection of non-linear devices to the power grid. The chapter presents the most common types
of harmonic studies, considerations for the representation of the power system including the extent of
the model and the scenarios/contingencies to analyse, guidelines and considerations in the production
of harmonic impedance loci and envelopes, limitations of power frequency short-circuit equivalents,
practical aspects to consider in the representation of customers installations, aggregation of harmonic
sources and issues related to measurement and estimation of background distortion. The discussions
in this chapter aim to address perspectives as seen from (i) the system owner/operator; and (ii) the new
connectee, with the overall objective of minimising risk of equipment failure due to excessive harmonic
distortion as well as avoiding unnecessary investment in mitigation. While finding the right balance is
never an easy task, it is hoped that the considerations presented in this chapter will aid the engineer in
making informed decisions.
Chapter 6 – Conclusions - summarises the guidelines and recommendations of this work. Finally, areas
identified as requiring further research are highlighted. A list of technical references is included in the
bibliography.
19
TB 766 - Network modelling for harmonic studies
20
TB 766 - Network modelling for harmonic studies
Frequency-Domain
Methods
Network Calculation of
Impedance Harmonic
Calculation Voltages and
Currents
Frequency
Scan Harmonic Harmonic
Penetration Load Flow
Balanced Unbalanced
Direct Iterative Balanced Unbalanced
Note: Balanced implies that the three phases are equal, or in sequence component terms, that only the positive sequence
exists. Unbalanced implies unequal phase voltages or phase currents.
21
TB 766 - Network modelling for harmonic studies
22
TB 766 - Network modelling for harmonic studies
Harmonic penetration is highly applicable to large system studies and is typically utilised by system
operators for planning purposes. This method provides a good overview of the harmonic level and
general system harmonic performance under steady-state operation.
Harmonic penetration solution methods provide a relatively simple means of analysis in which it is
assumed that there is no harmonic interaction between the network and non-linear devices. This means
that phenomena such as cross-coupling between the AC and DC sides of the converter and the
subsequent coupling between frequencies introduced by converters are ignored. It also means that the
harmonic injection as well as the converter harmonic impedance is assumed to be independent of (i)
any change in the operating point of the converter; or (ii) changes that may occur on the power system
such as the background distortion level.
There are cases in which the dependence of the converter on the operating point needs to be considered
in the frequency domain for both frequency scan and penetration studies. This can be achieved via the
use of several look-up tables (for instance those provided by the converter vendor) that change in
discrete steps as the operating point of the converter changes. This is a crude approach and quickly
results in the necessity to handle many look-up tables if the harmonic impedance and/or emission of the
converter is highly dependent on operating point. In practice, the highest emission level is normally
chosen while the converter and grid impedance are varied over all credible operating points, which leads
to conservative results.
Non-linearities in network components or power electronic devices are usually ignored in the frequency
domain. For studies where these are relevant, analysis should be performed in either the time or
harmonic domain.
23
TB 766 - Network modelling for harmonic studies
Multi-phase harmonic analysis is preferred or required under certain network conditions or topological
arrangements. In unbalanced (and balanced) systems for example, non-zero-sequence triplen
harmonics can penetrate regardless of transformer connection [3]. For the analysis of large networks,
the use of a multi-phase formulation may involve long computation times and large memory
requirements [16].
24
TB 766 - Network modelling for harmonic studies
Figure 2-2 Phase and Sequence Voltage in Inter-sequence Coupled Cable Radial Model (Receiving
End)
25
TB 766 - Network modelling for harmonic studies
Harmonic load flow solutions can be formulated in the frequency domain, time domain, or a hybrid
combination of both, and can be applied on a single-phase or multi-phase basis [8]. An overview of
various methods including detailed mathematical descriptions is provided in [4].
26
TB 766 - Network modelling for harmonic studies
2.6 Recommendations
The frequency domain solution methods provide adequate results for many applications, along with
straight-forward modelling of harmonic injections and frequency dependence, making this study domain
accessible and attractive to many engineers. It is generally numerically-robust and the computational
overheads of executing analyses in this domain are usually minimal and many study cases may be
considered with ease. For applications such as the calculation of AC system wide harmonics, evaluation
of AC filter ratings, impedance scans to evaluate potential resonance issues, modelling of network
harmonic impedance envelopes, DC-side harmonic profiles and filter design (amongst others), the
frequency domain is adequate and highly-recommended. However, if the nonlinear region of device
operation and any associated interaction should be modelled, the frequency domain may not provide
the engineer with sufficient accuracy.
The time domain provides a highly-accurate and powerful means to model the nonlinear time-varying
behaviour of devices. It can assist in identifying harmonic instability and other nonlinear interaction and
in performing analyses of AC/DC harmonic interaction for HVDC schemes (for control tuning and DC-
side harmonic performance validation). Significant effort and experience may be required on the part of
the engineer to incorporate all nonlinear models (e.g. transformer saturation, converter controllers
operating in their non-linear region, etc.) and to accurately represent frequency-dependent parameters.
Time-domain studies require validation that the system has reached steady-state before the parameters
of interest can be extracted (e.g. voltages and currents). For accurate representation of the low order
harmonic emission of a voltage-source converter, the dead-time of the switching elements must be
considered in the model. The computational overheads of time-domain simulations are larger than of
those performed in the frequency domain. These can be improved with automated execution,
exploitation of parallelised calculations and dedicated hardware. It should be mentioned that some
27
TB 766 - Network modelling for harmonic studies
commercial programs offer steady-state initialization routines that can significantly reduce simulation
time. However, such routines are however often not applicable for custom models (for instance OEM
HVDC models).
The combination of the advantages of both the frequency- and time domains was the driving force
behind the development of the hybrid domain. At the time of writing, the use of this domain in commercial
software packages generally requires the manual interfacing of frequency- and time-domain
functionality, requiring experience on the part of the engineer. It may present difficulties in cases where
models are supplied by different vendors, due to complex model interfaces.
The advantages, disadvantages and possible applications of the three domains are detailed in Table
2-1.
28
TB 766 - Network modelling for harmonic studies
29
TB 766 - Network modelling for harmonic studies
30
TB 766 - Network modelling for harmonic studies
31
TB 766 - Network modelling for harmonic studies
each phase, which can result in a large unbalance in the voltages and currents at certain harmonic
frequencies (see Section 2.2.2.6 for an illustrative example). When telephone interference is of
concern, zero sequence parameters are important. Considering the standard computing capabilities
available nowadays, the use of multi-phase models is normally the default option. Circuit
transpositions should be represented in detail, as they can influence unbalance at certain harmonic
frequencies. Likewise, any significant inter-circuit coupling should be captured, for instance parallel
circuits sharing right of way or close to metallic telephone lines. The size of the [Z] and [Y] matrices
will increase from 3x31 (single circuit) to 6x6 (double circuit) or higher for complex rights of way.
2. Capture the frequency dependency of the line parameters: The main effects contributing to
frequency-dependency are the conductor skin effect and the earth-return paths for zero sequence
currents.
3. Choice of nominal-PI (lumped parameters) or equivalent-PI (distributed parameters/long-line
effects). As discussed above, lumped parameter representations are only adequate for short lines
and low frequencies. However, when the line length approaches the wavelength of the frequency of
interest, the errors introduced are significant and a distributed parameter representation is required
(highlighted in section 3.1.5.1). Cascading nominal-PI sections is an alternative technique to
represent long lines. The more sections used, the closer the model approaches the distributed
nature of the parameters. In theory, increasing the number of cascaded sections to infinity will turn
the lumped parameter model into the distributed parameter model (demonstrated in section 3.1.5.2).
On the other hand, significant errors can be introduced if the sections are too large in relation to the
circuit length and the frequency range of interest. These errors increase with frequency. A practical
balance must be reached between accuracy and computational burden caused by the increased
number of intermediate nodes along the line. Section 3.1.5.2 discusses this topic in detail. Note that
in some cases the use of the cascading technique might be necessary; for example, when a
harmonic voltage/current profile along the line is needed.
The process of developing an OHL model for harmonic propagation studies can be divided into two
steps:
Step 1: Calculate the lumped electrical parameters from the circuit geometry and conductor physical
characteristics, capturing its frequency dependency. This requires calculation of the shunt
impedance and admittance matrices for each frequency within a selected range.
Step 2: Introduction of long-line effects.
The above functionality is typically built into commercial harmonic analysis software tools. The user only
needs to input the geometrical layout and conductor characteristics and specify the type of model
required for the study: lumped (nominal-PI) or distributed (equivalent-PI).
The distributed parameter (equivalent-PI) model is shown in Figure 3-2 and the reader is referred to [3]
for definitions of the parameters.
1
Under the assumption that ground wire reduction does not introduce significant error for the frequency range
of interest and type of study being undertaken [28]
32
TB 766 - Network modelling for harmonic studies
Assuming homogeneous non-ferrous conductors of tubular shape (as an approximation for ACSR
conductors), the internal impedance can be formulated using equations based on Bessel functions [3].
Some utilities instead model the skin effect using simpler correction factors. An approach suggested in
[2] and [27] is shown below.
where µr is the relative permeability of the cylindrical conductor, 𝑓 is the frequency in Hz, and 𝑅𝑑𝑐 is the
conductor DC resistance, in Ω/km. Correction factors developed in France and the UK are presented in
Table 3-1 ([3], [23]).
33
TB 766 - Network modelling for harmonic studies
A B C
Circuit 1 B C A
C A B
B C A
C A B
A B C
Circuit 2
34
TB 766 - Network modelling for harmonic studies
and represents an imaginary mirroring plane below the earth’s surface. Clearly there is a circular
relationship: the depth of the mirror plane varies with the resistivity, but the resistivity varies with depth.
Reasonably accurate solutions to this problem have been developed, normally simplifying the earth
strata to a more manageable equivalent of three different layers.
Such detailed calculation is unrealistic during a typical study involving many transmission lines of
considerable length. While the usual practice is to take a typical average earth resistivity for the area,
the engineer should be aware of the substantial errors which may be introduced; particularly in
calculations involving zero-sequence parameters or mutual impedance.
Illustration 1
Using Equation 3.2 with homogenous earth resistivity of 200 Ohm-m and frequency 550 Hz, the
depth of the mirror plane would be 214 m. However, for 1000 Ohm-m the depth would be 480 m.
Applying the Dubanton equations [32], the impact of these two resistivities on the mutual impedance
at 550 Hz between two conductors at say 50 m separation and 18 m height would be approximately
83%, while the impact on the self-impedance of one conductor would be around 13%.
Illustration 2
The result of an actual field measurement of earth resistivity is shown in Table 3-2 as an illustration
of the wide variations of resistivity with depth which occur in reality. Simply using an average value
of resistivity for line calculations is a very substantial approximation.
Table 3-2 Variation of Earth Resistivity with Depth
35
TB 766 - Network modelling for harmonic studies
The following additional observations can be made about the frequency response observed with the
accurate distributed parameter model:
The circuit harmonic impedance exhibits an infinite series of resonant points (series and parallel). It
can be demonstrated that these points are uniformly distributed over the frequency spectrum at
intervals of one quarter of its wavelength at fundamental frequency (λ 50/4) [3]. In this example, the
circuit wavelength is λ 50 = 5850 km and the observed intervals between “peaks” and “valleys” are
exactly as predicted: h=29.3 for 50 km length, h=14.8 for 100 km length, etc.
The frequency of the first resonant point drops as the circuit length increases. This is because the
parallel resonances occur at the half-wavelength frequencies of the line, considering on its length.
The amplitude of the resonant peaks drops as the frequency increases. This upper asymptotic
behaviour of the decreasing amplitudes is dependent upon the characteristic impedance of the line,
and the hyperbolic cotangent of the attenuation constant and the line length [3].
36
TB 766 - Network modelling for harmonic studies
Figure 3-5 Comparison of Lumped and Distributed Parameter Line Model for Various Line Lengths
(Blue=Lumped Parameter Model; Red=Distributed Parameter Model)
37
TB 766 - Network modelling for harmonic studies
Key Point:
A frequency-dependent, distributed parameter model capturing long line effects should be default
option for modelling OHL circuits in harmonics studies, for all but the very shortest lines.
Figure 3-6 Distributed vs. Cascaded Lumped Parameter Model Representation (Length = 250km)
38
TB 766 - Network modelling for harmonic studies
Figure 3-7 Skin Effect Modelling (Length = 250km) (Blue: Consideration of Skin Effect; Red=Neglection of
Skin Effect)
The following observations can be made:
The skin effect is only noticeable at the resonant points, providing increased damping. In this
example, the impedance of the first resonant peak calculated without skin effect is 45% higher than
with skin effect.
There is an attenuation of amplitude and phase angle when the skin effect is included. This means
that if the skin effect is neglected, the error thereby introduced increases with frequency.
Skin effect also results in a slight upward shift of resonant frequencies. This effect is only noticeable
at high frequencies (see the third resonant peak in Figure 3-7) and it is caused by an effective
reduction in the internal inductance of the phase conductors.
Key Point:
Neglecting skin effect leads to underestimation of circuit damping at resonant frequencies.
2
Note that this section refers to positive sequence only for simplicity. The reader is advised that conclusions
related to positive sequence can be extended to negative sequence for the purposes of OHL modelling.
39
TB 766 - Network modelling for harmonic studies
Figure 3-8 Effect of Earth Resistivity on Positive Sequence Impedance (Length = 250km)
Figure 3-9 Effect of Earth Resistivity on Zero Sequence Impedance (Length = 250km)
40
TB 766 - Network modelling for harmonic studies
Key Point:
Accurate estimation of earth resistivity is not necessary if only positive-sequence magnitudes are of
interest. If unbalanced conditions need to be assessed, an accurate estimation of earth resistivity is
needed to derive a correct zero-sequence model. For long circuits crossing areas with very different
earth conditions, it may be necessary to split the model into sections calculated with representative
earth resistivities.
However, this value is dependent on several physical factors such as terrain, river/motorways crossings,
span length, conductor tension, ambient temperature, etc. and electrical factors such as circuit loading.
Typically, an average figure is assumed to account for these uncertainties in the practical development
of a model. This example illustrates the sensitivity of the harmonic impedance to variations in conductor
height of ±10% of the assumed average value (as per Equation 3.3). A 250 km long circuit was selected
for this example. The OHL circuit was represented using the accurate distributed parameter line model.
Three cases were compared:
(i) Low average conductor height = 0.9 x 16.4 = 14.76 m
(ii) Medium average conductor height = 16.4 m
(iii) High average conductor height = 1.1 x 16.4 = 18.04 m
The results of the frequency scan analysis are presented in Figure 3-10.
41
TB 766 - Network modelling for harmonic studies
Figure 3-10 Effect of Average Conductor Height on the Positive Sequence Impedance (Length = 250km)
3.1.7 Summary
Various aspects of harmonic modelling of OHLs were discussed in this section and supporting examples
provided clear illustrations of the most relevant aspects. A summary of the analyses presented is given
in Table 3-3.
A comparison of line models as a function of the circuit length was presented, highlighting the
importance of using frequency-dependent, distributed parameter models when performing harmonic
studies. Modelling the frequency dependence of the series resistance (due to the skin effect) was shown
to have a significant impact on simulation results. Cascading PI sections was demonstrated to improve
the accuracy of the lumped parameter model up to a certain frequency, depending on the circuit length
and the number of sections modelled.
Values for the earth resistivity should be as accurate as possible to avoid significant approximation
errors, which significantly affect the zero-sequence harmonic impedance. This is relevant if unbalanced
conditions need to be assessed. The impact of small variations in the average conductor height above
ground was shown to be minor.
42
TB 766 - Network modelling for harmonic studies
43
TB 766 - Network modelling for harmonic studies
3.2 Cables
The network harmonic impedance is highly dependent upon the impedances of OHLs and cables.
Cables are more likely than OHLs to cause resonance at lower frequencies, which are of interest as
damping tends to be lowest at the lower end of the frequency spectrum. Cables typically have a lower
series impedance and higher shunt capacitance than OHLs. Both should be accurately modelled in
harmonic studies [34], [35]. The accurate modelling of submarine cable losses, for example, is difficult
and this impacts upon the damping of harmonics [36]. Achieving the required accuracy can prove difficult
due to studies being carried out during the planning stage of a project, at which time manufacturer-
provided design parameters may not be available and studies are instead based on generic data.
Deviations between data sheet values and physical values often exist for installed cables. The laying
and bonding configuration of cables play a key role in the local harmonic behaviour of the system and,
for older systems, these details might not be known. Other aspects influencing accuracy can be
simplifications or approximations made for reasons of computational efficiency in commercial power
system simulation packages. Simplistic formulae to model skin and proximity effects suffice at power
frequency but may lead to inaccuracies in terms of losses and damping at harmonic frequencies.
44
TB 766 - Network modelling for harmonic studies
Key Point:
Cable layout significantly affects the frequency, magnitude and number of resonances in the
positive-sequence harmonic impedance.
3.2.2 Models
Both cables and OHLs can be represented by a pi-circuit using either lumped or distributed parameters
as described in 3.1.5.1. The primary difference between models used for cables and OHLs is the
calculation of the impedances and admittances. This is due to:
Inherent design differences (a cable has a metal sheath, insulating layers, etc.);
An OHL is installed above ground, whereas cables are usually buried which affects the earth
impedance.
For a thorough discussion of the differences between the electrical characteristics of cables and OHLs,
the reader is referred to [34].
3.2.2.1 PI Models
The modelling of single-core (SC) cables has been well addressed in the literature [33], [38]. Models of
cables for several types of studies are discussed in [35], and recommendations provided. All models
are based on series impedances and shunt admittances, which are derived from the cable system
geometrical and physical characteristics. Like OHLs, cables can be modelled using either a nominal-PI
or an equivalent-PI representation. These are illustrated in Figure 3-1 and Figure 3-2, respectively.
A comparison of the cascaded nominal-PI and the equivalent-PI model is shown in Figure 3-11. As
shown, the number of nominal-PI sections that must be cascaded to maintain a certain accuracy
increases with frequency [3].
45
TB 766 - Network modelling for harmonic studies
A comparison of nominal- and equivalent-PI models for harmonic studies was made in [40]. For long
cables connected to a strong grid, the nominal-PI model was found to introduce significant error.
Conversely, the nominal-PI and equivalent-PI models yielded comparable results for the computed
resonance frequencies for short cables connected to a weak grid. Small errors in resonance frequencies
were found to lead to large errors in harmonic levels. With the computational power readily available at
the time of writing, the benefit of the nominal or cascaded nominal-PI section is limited.
Key Point:
Nominal-PI model: Should only be used to get an approximate idea of the frequency range in which
resonances can be expected. Unless used in cascade (rarely done in practice), this model can only
represent one resonant frequency. The nominal-PI model should be used with caution when
modelling long cables connected to a strong grid.
Equivalent-PI model: Should always be used for improved accuracy in harmonic studies.
46
TB 766 - Network modelling for harmonic studies
Figure 3-12 Cable Internal Impedance: Approximate vs. Analytical Consideration of Skin Effect [39]
Key Point:
The skin effect should be considered when calculating the harmonic impedance of either SC or
multi-core cables. Analytical methods which employ Bessel functions provide the highest
accuracy whereas approximate, truncated formulae can be inaccurate above a certain frequency.
Skin effect correction factors can be inaccurate at harmonic frequencies and may therefore lead to
incorrect harmonic losses and damping.
47
TB 766 - Network modelling for harmonic studies
This method can consider core-core, core-armour proximity effects and account for
individual strands of stranded armours. However, in order that the model be accurate,
detailed knowledge of the physical and electrical parameters of the cable is required [36].
(v) The subdivision of metallic components into sub-conductors. This shows good agreement
with FEM methods and is less computationally complex. The interested reader is referred
to [48] for further details.
A comparison of some of these methods is illustrated in Figure 3-13.
16,00
MoM-SO with S & P Effects (Stranded Armour)
MoM-SO with Skin and Proximity Effect
14,00 MoM-SO with Skin Effect
IEC 60287-1-1:2006
EDF Approximation
12,00
NGC Approximation
10,00
R1 [pu]
8,00
6,00
4,00
2,00
0,00
f [Hz]
A recent study [47] demonstrates that the inclusion of accurate modelling of frequency-dependent skin
and proximity effects in cables can be critically important and may yield large reductions in resonant
peaks in some cases. Inaccurate modelling of the cable can lead to incorrect evaluation of harmonic
damping and stability margins. For studies such as Power Park Module (PPM) design, this can result
in excessive harmonic levels in the PPM, or a costly over-design with unnecessarily reduced harmonic
levels. Similarly, the stability margins can become too low, thereby compromising PPM operation. The
positive sequence resistance calculated using various levels of accuracy is depicted in Figure 3-13 and
a frequency scan showing the positive sequence magnitude (taken from [47]) is provided in Figure 3-14.
It should be noted that in [47], to illustrate the effect on damping, the positive sequence inductance is
equal for each of the cases shown, and only the frequency-dependent, positive sequence resistance
characteristic is changed. This ensures that the resonances are aligned and only the effect of damping
can be seen.
48
TB 766 - Network modelling for harmonic studies
10000
Ideal (No Harmonic Losses Modelling)
MoM-SO with Skin Effect
MoM-SO with Skin and Proximity Effects
MoM-SO with Skin and Proximity Effects (Stranded Armour)
1000
|Z1| [Ω]
100
10
1
100 600 1100 1600 2100
f [Hz]
Figure 3-14 Consideration of Skin and Proximity Effects in MoM-SO: Effect on Positive Sequence
Impedance Magnitude [47]
Key Point:
The proximity effect should ideally be considered when calculating the harmonic impedance of either
multi-core cables or cables placed with cores in close proximity. Engineers should be aware that
proximity effect correction factors can be inaccurate at harmonic frequencies. Advanced methods
such as MoM-SO considering both skin and proximity effects (although currently only of limited
availability for use with commercial software) are recommended for highest accuracy when
calculating the harmonic impedance.
49
TB 766 - Network modelling for harmonic studies
3.2.4 Summary
A sensitivity analysis highlighted the most important parameters to capture accurately when modelling
cables for harmonic studies. The results are summarised in Table 3-4.
The impact of selection of model was discussed and the limitations of the nominal-PI model (including
cascaded nominal-PI sections) stated.
Formulae for the calculation of the cable impedance and admittance were referenced and discussed.
Phenomena such as the skin effect and proximity effects can be considered via various means from
simple correction factors to methods such as FEM and MoM-SO. When selecting the accuracy with
which to model these effects, the engineer should be aware of the possibility of incorrect estimation of
harmonic damping. To avoid this, the most accurate modelling available should be used.
50
TB 766 - Network modelling for harmonic studies
(A)
HV LV
(B)
Figure 3-15 Transformer Model (A) Physical Representation. (B) Single-Phase Equivalent Model
Transformer stray capacitances have no material impact in the range of frequencies of interest for typical
harmonics studies, and therefore do not need to be included in the model. The frequency threshold
above which transformer capacitances start to have some effect depends on the type and size of the
transformer unit. As guidance, a value of 4 kHz can be used to start considering the capacitances for
most transformers [27].
The magnetising branch (Lm and RFE) can be ignored under normal conditions, as it is generally assumed
that the transformer core operates in its linear region on a continuous basis. However, if transformer
saturation becomes a primary concern (e.g. operation at elevated voltage levels for extended periods
of time), this effect can be captured by including a harmonic current source in place of Lm and RFE. The
value of the current source should be determined from the flux-current curve and the operating voltage
[27].
Based on the above considerations, for typical harmonic studies within the scope of this Technical
Brochure, the transformer model shown in Figure 3-15 can be reduced to an impedance connected in
series between the two transformer terminals.
The position of the tap changer affects the leakage reactance of the transformer and the transfer
impedance from either side of the transformer (which follows the square of the turns ratio). The variation
in leakage reactance can produce shifts in resonant frequencies while the transfer impedance can affect
the level of damping provided between the two voltage levels. It is therefore recommended to capture
the effect of tap changers (range and step size) in transformer models for harmonic studies.
Transformer vector groups can introduce phase shifts to harmonic voltages and currents depending on
the harmonic order, the sequence and the transformer connections [27]. Therefore, explicit
representation of the winding connections is necessary for most harmonic studies. As an example, the
51
TB 766 - Network modelling for harmonic studies
parallel connection of wye and delta transformer windings is used to combine two 6-pulse rectifiers into
a 12-pulse one. In this application, the 30° phase shift between a star-connected winding and a delta-
connected winding introduces cancellation of the 6n harmonic orders.
Transformer winding connections also affect the flow of zero-sequence harmonic currents. In general,
triplen harmonic currents tend to be trapped in delta winding connections (assuming perfectly balanced
conditions) while they add up in grounded transformer neutrals. This may result in excessive currents in
the transformer neutral leading to operation of protection relays. Explicit representation of winding and
neutral connections is recommended.
One of the key aspects of transformer modelling, and normally neglected, is the correct representation
of losses. While power transformers are specified to minimise resistive losses at fundamental frequency,
their presence at higher frequencies provides beneficial damping of harmonic resonances. Correct
estimation of this damping can reduce, or even negate, the need for harmonic mitigation measures. This
needs to be accurately captured in the series impedance components of the transformer model.
The following sub-sections concentrate on the representation of the frequency dependent behaviour of
the series impedance components of the transformer model. An overview of power transformer models
that are frequently used in industry is provided first. Then, the performance of the transformer models is
assessed and compared in a larger model representing a real power system and recommendations are
provided. Finally, comparisons of simulation results using the transformer models against real
measurements provide the basis for further recommendations.
It should be noted that the models presented in this section are based on 2-winding transformers, and
that significantly higher harmonic losses should be expected in 3-winding transformers where a tertiary
is involved. Models used for 2-winding transformers may be extended for application to 3-winding
transformers, however this should be done with caution. Furthermore, the models presented here
assume positive sequence impedances. A similar approach could be adopted for representation of zero-
sequence impedances, provided that winding and neutral connections are correctly represented in the
zero-sequence equivalent network.
The input data requirements for this model are the transformer leakage reactance at fundamental
frequency (X1 ) in Ohm and the rated nominal power of the transformer (Sr ) in MVA.
The model assumes that the leakage inductance 𝐿σ is constant over the range of frequencies of interest,
therefore Xh increases linearly with frequency as per Equation 3.4 and Equation 3.5.
52
TB 766 - Network modelling for harmonic studies
𝑋ℎ = ℎ𝑋1
Equation 3.5
The resistances R s and R p are constant at all frequencies and their values (in Ohm) are given by
Equation 3.6 and Equation 3.7, based on the rated nominal power of the transformer (in MVA), as per
Equation 3.8 (if better knowledge is not available).
𝑋1
𝑅𝑆 = Equation 3.6
𝑡𝑎𝑛( 𝜓1 )
𝑅𝑝 = 10 ⋅ 𝑋1 ⋅ 𝑡𝑎𝑛( 𝜓1 )
Equation 3.7
0.693+0.796 𝑙𝑛 𝑆𝑟 −0.0421(𝑙𝑛 𝑆𝑟 )2
𝑡𝑎𝑛 𝜓1 = 𝑒
Equation 3.8
The equivalent impedance of the transformer Zh is given by Equation 3.9 below, which can be easily
implemented in a software tool as a series combination of R(h) and X(h) with frequency dependant
correction factors.
𝑗𝑋ℎ 𝑅𝑝 ℎ2 𝑋12 𝑅𝑝 ℎ𝑋1 𝑅𝑝2
𝑍ℎ = 𝑅(ℎ) + 𝑗𝑋(ℎ) = 𝑅𝑠 + = (𝑅𝑠 + 2 2 ) + 𝑗 ( ) Equation 3.9
𝑅𝑝 + 𝑗𝑋ℎ 𝑅𝑝 + ℎ2 𝑋1 𝑅𝑝2 + ℎ2 𝑋12
Equation 3.9 shows that the resistive component of Zh is a function of the frequency, demonstrating that
skin effect is captured in this model. Furthermore, even though X(h) presents a complex dependency
with frequency, experience shows that a linear approximation is adequate for the range of frequencies
of interest in harmonic analysis. Therefore, 𝑋(ℎ) ≅ ℎ𝑋1 is usually adopted as a valid approximation.
s h
The input data requirements for this model are the transformer DC resistance R DC (in Ohm) and the
leakage reactance at fundamental frequency X1 (in Ohm).
The model assumes that the leakage inductance 𝐿σ is constant over the range of frequencies of interest,
therefore Xh increases linearly with frequency as per Equation 3.10 and Equation 3.11.
The resistance modelling accounts for skin effect. The frequency dependency of the resistance is
calculated using Equation 3.12. The recommended values for A and B are 0.1 and 1.5, respectively [49].
𝑅S (ℎ) = 𝑅DC (1 + 𝐴ℎ𝐵 ) Equation 3.12
53
TB 766 - Network modelling for harmonic studies
s h
The input data requirements for this model are the short-circuit resistance R t (in Ohm) and the leakage
reactance at fundamental frequency X1 (in Ohm).
The model assumes that the leakage inductance 𝐿σ is constant over the range of frequencies of interest,
therefore Xh increases linearly with frequency as per Equation 3.13 and Equation 3.14.
𝑋1 = 2𝜋𝑓𝑛 𝐿𝜎 Equation 3.13
The resistance modelling accounts for skin effect. The frequency dependency of the resistance is
calculated using Equation 3.15. The recommended values for the coefficients are given in Table 3-5
with the condition that 𝒂𝟎 + 𝒂𝟏 + 𝒂𝟐 = 𝟏 [1].
𝒂𝟎 𝒂𝟏 𝒂𝟐 𝒃
Small system transformer 0.85 – 0.90 0.05 – 0.08 0.05 – 0.08 0.9 – 1.4
Large system transformer 0.75 – 0.80 0.10 – 0.13 0.10 – 0.13 0.9 – 1.4
Usually, it can be assumed that the coefficients for the small system transformer should be used if its
power rating is below 100 MVA. Otherwise the coefficients for a large system transformer should be
used.
s h
The input data requirements for this model are the transformer resistance at fundamental frequency 𝑅
(in Ohm) and the leakage reactance at fundamental frequency X1 (in Ohm).
The leakage inductance 𝐿σ is not assumed to be constant and a typical L-f characteristic is included in
[24] showing a decay in inductance with increasing frequency.
The harmonic impedance is calculated using Equation 3.16 below.
𝑍T (ℎ) = 𝑅 √ℎ + jℎ𝑋1
Equation 3.16
54
TB 766 - Network modelling for harmonic studies
s h
The input data requirements for this model are the transformer resistance at fundamental frequency R fn
(in Ohm) and the leakage inductance at fundamental frequency Lfn (in Henry).
The frequency dependency of the resistance and inductance of the transformer are calculated using
Equation 3.17 and Equation 3.18 below. The recommended values for the coefficients are given in Table
3-6 [50].
BR
f
R s = R fn (1 + AR ( − 1) ) Equation 3.17
fn
f BL
L = L fn A L ( ) Equation 3.18
fn
𝑹(𝒇) 𝑳(𝒇)
Transformer type
𝑨𝐑 𝑩𝐑 𝑨𝐋 𝑩𝐋
20 kV / 0.4 kV, 250 kVA 0.2 1.5 1 -0.03
108 kV / 10.5 kV, 40 MVA 0.2 1.4 1 -0.02
220 kV / 110 kV, 200 MVA 0.2 1.6 1 ≈0
3.3.2.1 Example 1
Single-phase two winding transformer unit with the following technical parameters:
Rating: 100 MVA (single phase)
Primary Voltage: 735 kV
Secondary Voltage: 26 kV
Nominal frequency: 60 Hz
Short Circuit impedance (Zk): 12.92%
X/R at nominal frequency: 68
Frequency response measurements were performed by white noise injection at the 26 kV transformer
terminals.
The response of each transformer model option is compared against the measured data for this
transformer unit in Figure 3-21 (resistance) and Figure 3-22 (reactance). The magnitudes plotted in
these graphs are expressed in per unit, based on the rated MVA of the transformer. The model
parameters selected for each assessment are included in Table 3-7.
Figure 3-21 (a) shows the calculated resistance values using the recommended default model
parameters. It can be observed that Model 1 results in an over-estimation of the transformer resistance
55
TB 766 - Network modelling for harmonic studies
as frequency increases. Similarly, Model 2, Model 4 and Model 5 result in an under-estimation of the
transformer resistance. In this example, Model 3 yields the closest match to the measured values.
Further model tuning was performed in an attempt to find a better fit with the measurements - Figure
3-21 (b). The best overall result was obtained with Model 1 by using the measured X/R ratio (at 60Hz)
instead of the calculated tan 𝛹. Model 2 and Model 5 also gave reasonable results when the parameters
were optimised to fit the measured values.
Figure 3-22 shows the calculated and measured reactance values as a function of frequency. All models
produce very similar results and are in good agreement with the measurements. The reactance can be
considered to vary linearly with frequency for the range of interest in harmonics studies. Above 2 – 3
kHz this behaviour starts to deviate as the effective winding capacitances start having a noticeable
impact on the transformer response.
The measured and calculated values of the time constant L/R(f) are plotted in Figure 3-23 for
completeness.
Table 3-7 Example 1: Transformer Model Parameters
0,50
(a) Default Model Parameters
0,40
R(f) [pu]
0,30
0,20
0,10
0,00 f [Hz]
0 500 1000 1500 2000 2500
0,40
0,35 (b) Modified Model Parameters
0,30
R(f) [pu]
0,25
0,20
0,15
0,10
0,05
0,00 f [Hz]
0 500 1000 1500 2000 2500
Measurement Model 1 (modified) Model 2 (modified) Model 3 Model 5 (modified)
Figure 3-21 Example 1: Calculated and Measured Transformer Resistance as a Function of Frequency
56
TB 766 - Network modelling for harmonic studies
6,0
5,0
4,0
X(f) [pu]
3,0
2,0
1,0
0,0 f [Hz]
0 500 1000 1500 2000 2500
Measurement Model 1 Model 2 Model 3 Model 4 Model 5
Figure 3-22 Example 1: Calculated and Measured Transformer Reactance as a Function of Frequency
1000
100
L/R(f) [ms]
10
0,1 f [Hz]
60 600 6000
Measurement Model 1 (modified) Model 2 (modified) Model 3 Model 5 (modified)
3.3.2.2 Example 2
Three-phase wind farm transformer with the following parameters:
Rating: 90 MVA
Primary Voltage: 132 kV
Secondary Voltage: 33 kV
Nominal frequency: 50 Hz
Short circuit impedance (Zk): 11%
X/R at nominal frequency: 51.7
The frequency response measurement method is detailed in [51]. Figure 3-24 and Figure 3-25 below
show the variation with frequency of the resistance and inductance for this transformer, referred to the
datasheet values (at nominal frequency). A fast increase in the resistance value can be observed within
the range of frequencies of interest for harmonic analysis. However, the transformer inductance changes
very slowly with frequency. In practical terms, a representation based on a constant inductance appears
reasonable for this transformer. Therefore, this section will concentrate on assessing the representation
of the resistive component only.
57
TB 766 - Network modelling for harmonic studies
The response of each model option is compared to the measured resistance data for this transformer in
Figure 3-26. The resistance values plotted in this figure are expressed in per unit, based on the rated
58
TB 766 - Network modelling for harmonic studies
MVA of the transformer. The model parameters selected for each assessment are provided in Table
3-8.
Figure 3-26 (a) shows the calculated resistance values using the recommended default model
parameters. It can be observed that Model 1 and Model 3 result in an over-estimation of the transformer
resistance as frequency increases. Similarly, Model 2, Model 4 and Model 5 result in an under-estimation
of the transformer resistance. These responses match the observations made in Example 1, except for
Model 3 which gives a reasonable fit in Example 1 but overestimates the resistance in Example 2.
Further model tuning was performed to find a better fit with the measurements and the results are shown
in Figure 3-26 (b). The improved parameters gave very close results to the measured values for Model
1, Model 2, Model 3 and Model 5. Essentially, most modelling input data can be tuned to reproduce
measurements. It should be noted that, unlike Example 1, the use of the power frequency X/R ratio with
Model 1 did not produce a good match in this transformer and a higher value had to be selected.
The measured and calculated values of the time constant L/R(f) are plotted in Figure 3-27 for
completeness. This graph is almost identical to the one presented for Example 1 (Figure 3-23) and can
be used as reference for this type of transformer.
Table 3-8 Example 2: Transformer Model Parameters
Default Parameters Modified Parameters
Model 1: Electra-167 [2] tan 𝛹 = 30.64 tan 𝛹 = 65
Model 2: IEEE-399 [49] A = 0.1, B = 1.5 A = 1, B = 1.95
a0 = 0.87, a1 = 0.05,
Model 3: Electra-164 [1] a0 = 0.775, a1 = 0.115, a2 = 0.11, b= 1.15
a2 = 0.08, b= 1.2
Model 4: Arrillaga [24] N/A N/A
Model 5: Funk [50] AR = 0.2, BR = 1.6 AR = 0.25, BR = 1.72
1,00
(a) Default Model Parameters
0,80
R(f) [pu]
0,60
0,40
0,20
0,00 f [Hz]
0 500 1000 1500 2000 2500
Measurement Model 1 Model 2 Model 3 Model 4 Model 5
0,50
(b) Modified Model Parameters
0,40
R(f) [pu]
0,30
0,20
0,10
0,00 f [Hz]
0 500 1000 1500 2000 2500
Measurement Model 1 (modified) Model 2 (modified) Model 3 (modified) Model 5 (modified)
Figure 3-26 Example 2: Calculated and Measured Transformer Resistance as a function of Frequency
59
TB 766 - Network modelling for harmonic studies
1000
100
L/R(f) [ms]
10
0,1 f [Hz]
60 600 6000
Measurement Model 1 (modified) Model 2 (modified) Model 3 (modified) Model 5 (modified)
Key Points:
For the transformer assessed in Example 2:
None of the transformer model options gave a reasonable match to the measured transformer
resistance. Model 1 and Model 3 over-estimated the resistance values while Model 2, Model 4 and
Model 5 under-estimated the resistance.
The response of all transformer model options is very sensitive to the selected parameters. In all
cases it is possible to reproduce measurements by back-calculating the parameters.
The transformer inductance can be assumed to be constant for the range of frequencies of interest.
Under this assumption, the transformer reactance increases linearly with frequency.
3.3.2.3 Example 3
Wind turbine generator transformer with the following technical parameters:
Rating: 8.2 MVA
Primary Voltage: 33 kV (delta connection)
Secondary Voltage: 0.69 kV
Nominal frequency: 50 Hz
Resistance at 50 Hz: 3.32 Ω (delta connection)
Reactance at 50 Hz: 31.07 Ω (delta connection)
X/R at nominal frequency: 9.35
The response of each model option is compared against the measured resistance data for this
transformer in Figure 3-28. The resistance values plotted in this figure are expressed in per unit, based
on the rated MVA of the transformer.
Figure 3-28 shows the calculated resistance values using the recommended default model parameters.
It can be observed that Model 1 and Model 3 result in an over-estimation of the transformer resistance
as frequency increases. Similarly, Model 4 underestimates the transformer resistance. For this
transformer, Model 2 and Model 5 gave reasonable matches to the measurements.
The measured and calculated values of the time constant L/R(f) are plotted in Figure 3-29 for
completeness.
60
TB 766 - Network modelling for harmonic studies
1,50
R(f) [pu]
1,00
0,50
0,00 f [Hz]
0 500 1000 1500 2000 2500
Measurement Model 1 Model 2 Model 3 Model 4 Model 5
Figure 3-28 Example 3: Calculated and Measured Transformer Resistance as a function of Frequency
1000
100
L/R(f) [ms]
10
0,1 f [Hz]
60 600 6000
Measurement Model 2 Model 5
Key Points:
For the small transformer assessed in Example 3:
Model 2 (IEEE-399) and Model 5 (Funk) gave the best matches to the available measurements.
61
TB 766 - Network modelling for harmonic studies
transformers have a material impact on the harmonic impedance in the frequency range from 2 kHz
to 2.4 kHz. Within this range, Model 1 (Electra 167) provides the highest damping with a 47%
reduction in the maximum harmonic impedance at resonance. Model 5 (Funk) provides a 25%
reduction in impedance with respect to base case and Model 4 (Arrillaga) provides the lowest
damping with just a 2% reduction in harmonic impedance.
Node #3 is a transmission collector station for a cluster of wind farms. The station has 2 x 250MVA
220/110kV transformers and several 110kV insulated cable connections. These power
transformers have a material impact on the harmonic impedance in the frequency range from 250
Hz to 450 Hz. In this case, the impedance reduction observed with each transformer model is 24%,
13% and 8% for Model 1, Model 5 and Model 4, respectively.
Node #4 has been fictitiously created to introduce a parallel resonance at 1220 Hz between a
250MVA 220/110kV transformer and a 110kV radial cable connection. In this case, the impedance
reduction observed with each transformer model is 84%, 71% and 23% for Model 1, Model 5 and
Model 4, respectively.
Key Points:
The choice of transformer model is of key importance for nodes where the harmonic impedance is
dominated or influenced by power transformers. It has been shown that different transformer
representations result in large deviations in harmonic impedance at major resonance points.
For nodes that are electrically distant from power transformers, the selection of a transformer model
(or constant resistant) does not have material impact on the calculation results.
All transformer models considered in this document introduce some level of frequency dependent
resistive losses. The effect of these losses is only material at parallel resonant points in the power
system, where the harmonic resistance reaches a peak. All models exhibit similar behaviour outside
the resonant frequencies.
The quantitative effect on damping introduced by each model depends on the location of resonant
frequencies for the node of interest. All models provide increasing damping with frequency;
therefore, higher differences in performance are observed when the main resonances appear at the
higher frequency range.
All modelling options described in this document assume that the transformer leakage inductance
is constant for the range of frequencies of interest. Therefore, the reactance increases linearly with
frequency in all model options. This means that the selection of a transformer model does not affect
the location of the resonant frequencies, just the amplitude of the harmonic impedance. This can be
used as a sensitivity test to check whether the harmonic impedance of a node is dominated by a
transformer or by other network components.
62
TB 766 - Network modelling for harmonic studies
2500 2500
2000 2000
1500 1500
1000 1000
500
500
0
0
0 500 1000 1500 2000 2500
2000 2050 2100 2150 2200 2250 2300 2350 2400
f [Hz]
Constant Resistance Model 1 Model 4 Model 5 Constant Resistance Model 1 Model 4 Model 5 f [Hz]
150 150
100 100
50 50
0 0
0 500 1000 1500 2000 2500 250 300 350 400 450
f [Hz] f [Hz]
Constant Resistance Model 1 Model 4 Model 5 Constant Resistance Model 1 Model 4 Model 5
15000 15000
10000 10000
5000 5000
0 0
0 500 1000 1500 2000 2500 1150 1170 1190 1210 1230 1250 1270 1290 f [Hz]
f [Hz]
Constant Resistance Model 1 Model 4 Model 5 Constant Resistance Model 1 Model 4 Model 5
Figure 3-30 Effect of Detailed Power Transformer Models on Harmonic Impedance in a real Transmission
System
3.3.4 Summary
In general, a transformer can be represented by a series combination of resistance and inductive
reactance. Both components are frequency dependant and this section deals with different approaches
to capture this feature.
Five models were described in this section and case studies presented highlighting various important
aspects of power transformer modelling for harmonics studies. The reader is reminded that, for
simplicity, the models presented here are based on 2-winding transformers and positive-sequence
parameters. The models, however, can be extended to represent higher number of windings and zero-
63
TB 766 - Network modelling for harmonic studies
sequence behaviour in line with common power frequency modelling approaches, if data describing
their frequency behaviour is available. .
The data required for each model can be easily obtained from the transformer data sheet, or preferably,
from the Factory Acceptance Test (FAT) report. Some of the models require additional factors for which
default values have been provided. However, best calculation results are achieved if the manufacturer
of a transformer provides a frequency-dependent model or if an appropriate model is developed by the
operator based on their own measurement data.
Frequency dependency of the transformer resistance is an area where caution is advised as certain
models used may introduce optimistic damping.
All model options capture the frequency dependency of the resistive losses with various levels of
complexity. This effect can be implemented in most commercial software tools by a correction function
applied to the 50/60Hz transformer resistance. The transformer reactance can be assumed to increase
linearly with frequency for the typical range of interest in harmonics studies.
In general, it is important to consider the position of the tap changer as it affects the leakage reactance
of the transformer and the transfer impedance from either side of the transformer (which follows the
square of the turns ratio). Transformer stray capacitances do not have material impact in the range of
frequencies of interest for most harmonics studies, and therefore do not need to be included in the
model. As a general rule, transformer capacitances can be ignored for frequencies up to 4 kHz in most
cases. The transformer winding connections should be represented to take account of the phase-shifting
effect on harmonic currents. The magnetising branch can be ignored under normal conditions, but it
should be represented as a harmonic source if transformer saturation becomes a primary concern.
The five transformer model options presented in this document were assessed against measurements
from three power transformers. This assessment highlighted significant differences between all the
options and their high sensitivity to the selected model parameters. It was observed that, in general,
Model 1 (Electra No. 167 [2]) tend to overestimate resistive losses if the default expression for tan 𝛹
was used, however reasonable results were observed when the 50/60Hz datasheet X/R value was used
instead. Model 4 (Arrillaga [24]) always gave the lowest resistive losses, significantly below the
measured values.
A comparison of the models via frequency scans using a representation of the Irish transmission grid
suggests that the selection of a transformer model only has a noticeable effect at the parallel resonant
points and at nodes whose harmonic impedance is dominated or influenced by power transformers.
If high accuracy is needed in the studies, it is recommended to obtain frequency dependent
characteristics for R and X from the transformer manufacturer. In the absence of measurements to
validate models, the analysis presented in this section suggest the use of Model 5 (Funk [50]) with
default parameters or Model 1 (Electra No. 167 [2]) with 50/60Hz datasheet X/R value as options with
reasonable results.
Alternatively, measurements from a range of transformers presented in [51] and [52] may be used. In
this recent work ([51]) the R(ω)/L ratio of different transformers types and sizes is compared (see Figure
3-31). The measurements were obtained using frequency response analysis (FRA) devices. The reader
is advised that the accuracy of the FRA measurements can be uncertain for smaller transformers within
the low frequency range, i.e. close to the fundamental component. Therefore, it is recommended to
confirm with the supplier whether the measurements are reliable and/or use a hybrid approach including
fundamental frequency short-circuit test values to define the transformer frequency-dependent
characteristic behaviour within the low frequency range [53].
64
TB 766 - Network modelling for harmonic studies
Figure 3-31 Comparison of the 𝑳/𝑹(𝝎) Ratio of Different Transformer Types [51]
65
TB 766 - Network modelling for harmonic studies
3.4 Loads
Loads have a considerable influence on the harmonic characteristics of the network in terms of location
of resonances and level of damping, and their appropriate modelling is a very important consideration.
This is becoming more important due to the increasing number of HVDC interconnectors, offshore and
onshore wind power plants, PV-based generation and large-scale integration of power electronic based
distributed generation. This section provides an overview of load models that are frequently used in
industry and reviews and discusses reported new harmonic load models as well as novel methods to
tackle the changing load composition due to the increase in power electronic type devices. Comparisons
of simulation results using the most common load models provide the basis upon which
recommendations are provided.
Load behaviour is usually classified as either linear or nonlinear. Linear loads draw current in a
sinusoidal manner and can be grouped into static loads (lighting, resistive heating, etc.) and rotating
loads (asynchronous induction and synchronous motors). While static loads are generally well
represented by their active and reactive power demand at fundamental frequency, the same does not
apply to rotating loads which require more sophisticated models.
Nonlinear loads include power electronic-type loads such as PVs, variable speed drives, battery storage
systems, electric plug-in vehicles, data centres, etc. These loads draw current in a non-sinusoidal
manner. Accurate harmonic modelling of nonlinear loads often presents difficulties due to them being
sources of harmonics, compounded by the fact that their variable R, L, C configuration and nonlinear
characteristics cannot be represented using the linear harmonic equivalent model [3]. Consideration of
nonlinear loads in harmonic studies may need to include (i) the type and topology of the nonlinear loads;
(ii) the interaction with the AC system impedance; and (iii) the variation in the harmonic spectra [16]. In
addition, power electronics-based loads exhibit a more complex damping characteristic than that
exhibited by linear loads, and the effect of controllers on load behaviour should be considered.
The general load composition and its harmonic behaviour is normally one of the biggest unknowns when
performing harmonic assessments in power systems. On the one hand the loads are highly variable in
time and on the other hand, the active power absorbed by rotating machines does not exactly
correspond to a damping value [26]. A common approach is to aggregate the effect of the general load
and distribution networks into a single reduced model. The aggregate harmonic load model refers to an
equivalent representation of a group of loads at a particular voltage level (i.e. LV or MV) and will include
elements of a transmission and distribution network. This means that network elements such as grid
transformers at Grid Supply Points (GSPs), OHLs, cables and shunt compensation devices would also
be part of the model. Therefore, the loads can affect the harmonic response in the network through
damping and the introduction of both parallel and series resonances. If the load is not accurately
represented in harmonic studies, results will likely be unrealistic.
The derivation of the aggregate harmonic load model can be either component-based (bottom-up or
knowledge-based) or measurement-based (top-down or behaviour-based), as described in detail in [54].
The important difference is that a component-based approach can be applied to load consisting of
multiple components, while a measurement-based approach can be applied to load consisting of single
and multiple components.
66
TB 766 - Network modelling for harmonic studies
𝑈2
𝑅= Equation 3.19
𝑃1
This method is recommended only when the motor part is very small; i.e. for commercial
and domestic loads in which the motor component is so partitioned that the resistive effect
is dominant [8].
R jhX
Figure 3-33 Method 2 (CIGRE WG 36.05) Load Model (Static and Rotating)
Where P1 is the total MW demand, K is the motor fraction of the total MW load, K E is the
electronic controlled load fraction of the total MW load and K1 represents the severity of the
motor starting condition (to estimate the equivalent negative sequence inductance).
According to [8], the following values can be assumed when no better information is
available:
K can assume a value of 0.8 for industrial loads and 0.15 for commercial and domestic
loads.
K1 assumes normally a value between 4 and 7.
KE can assume values around 0.
67
TB 766 - Network modelling for harmonic studies
An additional parallel resistor representing the motor damping can be included as R 1 = L/K2
where K2 is a fraction of the negative sequence or locked-rotor inductances. Typical values
of K2 are around 0.2.
It should be noted that more research is required to evaluate the applicability of these
assumptions to modern load compositions.
Method 3 (CIGRE WG 36.05):
This method is based on experimental results at MV distribution networks in France by
EDF. The loads are represented by a reactance, Xs, in series with resistance R (static
branch); this assembly being connected in parallel with a reactance X p (rotating branch),
as shown in Figure 3-34. This model is applicable over a frequency range approximately
between the 5th and the 20th harmonic orders [26].
jXS
jXP
The model parameters are calculated from both the active and reactive power components
of the load at power frequency and empirical constants, as follows:
𝑈2
𝑅= Equation 3.22
(1 − 𝐾) ∙ 𝑃1
𝑋𝑠 = 0.073 ℎ𝑅 Equation 3.23
ℎ𝑅
𝑋𝑝 = Equation 3.24
𝐾 ∙ 6.7 tan ∅1 − 0.74
Where U is the nominal voltage of the network, P1 is the active power of the load (at the
fundamental frequency), K is the motor fraction of the total MW load, 𝑡𝑎𝑛 ∅1 = 𝑄/𝑃 and Q
is reactive power of the load (at the fundamental frequency). Compensation capacitors and
cables are not captured in this model and must be represented externally.
68
TB 766 - Network modelling for harmonic studies
R jhX
𝑈2
𝑅= Equation 3.25
𝑃
𝑈2
𝑋= Equation 3.26
𝑄
jhX
69
TB 766 - Network modelling for harmonic studies
Figure 3-37 Method 1 (CIGRE JTF 36.05.02/14.03.03) “Passive” Load Model (Domestic)
The equivalent resistance is estimated from the active power at fundamental frequency as
per Equation 3.31, where U is the nominal voltage of the network and P is the active power
of the load (at the fundamental frequency). The factor √ℎ captures the frequency
dependence of the resistive damping. The equivalent reactance is calculated using
Equation 3.32 and represents the relatively small motor content assumed in domestic load.
In some regions, particularly where residential air conditioning is prevalent, motors can
comprise a major if not dominant portion of domestic loads.
jXmh
Rmh
70
TB 766 - Network modelling for harmonic studies
R jhX
𝑈2
𝑅= Equation 3.37
(0.1ℎ + 0.9)𝑃
𝑈2
𝑋= Equation 3.38
(0.1ℎ + 0.9)𝑄
Where P and Q are the fundamental frequency active and reactive power measured at
voltage level U.
71
TB 766 - Network modelling for harmonic studies
Model Parameters
Model 1 (IEEE; “RL Series”). Series
jhX
𝑉2
𝑅=𝑃∙
𝑃2 + 𝑄2
𝑉2
𝑋=𝑄∙
R 𝑃2 + 𝑄2
𝑉2
𝑅=
𝑃
R jhX 𝑉2
𝑋=
𝑄
𝑉2
𝑅(ℎ) =
𝑚(ℎ) ∙ 𝑃
𝑉2
R(h ) jhX ( h ) 𝑋(ℎ) =
𝑚(ℎ) ∙ 𝑄
𝑚(ℎ) = 0.1ℎ + 0.9
72
TB 766 - Network modelling for harmonic studies
𝑉2
𝑅2 =
(1 − 𝐾) ∙ 𝑃
jhX 2 𝑋2 = 0.073 ∙ 𝑅2
𝑉2
jhX 1 𝑋1 =
𝐾 ∙ 𝑃 ∙ (6.7 ∙ tan 𝜙 − 0.74)
R2 𝑄
tan 𝜙 =
𝑃
𝑋1 and 𝑅2 as in Model 4.
jhX 2 jhX 1
𝑋2 = 0.1 ∙ 𝑅2
𝑋1
𝑅1 =
𝐾3
R2 R1
𝐾3 is the effective quality factor of the motor circuit (≈ 8)
jXS jXa
RS Ra
Static Rotating
part part
73
TB 766 - Network modelling for harmonic studies
𝑈2
𝑋𝑎 = 𝑋 ℎ [2(ℎ 𝑓0 )𝛽 ] Equation 3.42
𝑃 𝑝 𝑙𝑟
Where P is the active power absorbed by entire load area, p is the percentage of rotating loads, Xlr is
the locked rotor mean equivalent reactance, Rm is the equivalent series resistance, f0 is the fundamental
frequency, h is the harmonic order, while β and k are the parameters of the proposed model.
Typical parameters for use in this model are provided in [56].
A refinement to the model in [56] was presented in [57], also validated through experimental tests in
Italy. The equivalent circuit was identical to the one proposed in [56]. All equations for determination of
parameters were the same, except for the modification in the expression for Xs as shown below:
𝑋𝑠 = 𝛼 ℎ𝑅𝑠 Equation 3.43
The parameter α is an estimate of an average power factor of the static load (i.e. known Q/P). The
average values for k were found to be between 0.25 and 0.75. The field tests determined that the
parameter β is between -0.12 and -0.17. A value in the range of 0.03-0.04 pu (20% of Xlr) is
recommended for Rm.
BHV
XT
0,415 kV
RLV BLV
Figure 3-42 Downstream Harmonic Component Load Model Representation at Supply Bus GSP (Typically
132kV) as used by National Grid TSO
The selection of the model parameters depends on the maximum permitted MW load seen at the supply
bus, the topology and the load composition. The load composition will depend on whether the load is
domestic, agricultural, commercial, commercial/domestic, industrial, lighting or traction, etc.
74
TB 766 - Network modelling for harmonic studies
For National Grid, RLV is based on the proportion of load energy sales, where the various categories are
compiled according to domestic, farm, commercial, industrial and rail for the resistive and motor load
breakdowns with its assumed power factor correction (PFC). Table 3-9 gives an example of typical
proportions based on the maximum Average Cold Spell (ACS) demand (MW) to calculate RLV damping
(MW) and PFC (Mvar). It is important to note that these factors are pessimistic and are representative
of the typical load profile in the entire National Grid system.
Table 3-9 Typical values of RLV and PFC *courtesy National Grid
The actual Mvar values for BHV and BLV are based on a nominal system feeding a 10 MW load from a
typical 132kV supply distribution system based on work carried out by the UK Electricity Council and
published in CIRED 1983. This has been used to derive the typical equivalent transformer reactance XT
and BLV as given in Table 3-10.
Table 3-10 Typical BLV (Power Factor Correction (PFC) + Cable Capacitance) and Equivalent XT
Transformer Reactance *courtesy National Grid
Approximate
Permitted Supply Bus Step-Down XT to 0.415kV BLV (Total 0.415kV
Max. MW Voltage (kV) Transformers (% on 100MVA) Capacitance) Mvar
Load
400 132 16 x 60 MVA 3.7 PFC + 26
350 132 14 x 60 MVA 4.1 PFC + 23
300 132 12 x 60 MVA 4.9 PFC + 20
250 132 10 x 60 MVA 5.7 PFC + 17
200 132 8 x 60 MVA 7.3 PFC + 13
150 132 6 x 60 MVA 9.5 PFC + 11
100 132 4 x 60 MVA 14.7 PFC + 7
50 132 2 x 60 MVA 26.3 PFC + 4.0
180 33 18 x 23 MVA 5.1 PFC + 3.6
150 33 15 x 23 MVA 6.1 PFC + 3.0
120 33 12 x 23 MVA 7.7 PFC + 2.4
90 33 9 x 23 MVA 10.2 PFC + 2.0
60 33 6 x 23 MVA 15.3 PFC + 1.0
60 11 3.7 PFC
50 11 4.4 PFC
40 11 5.5 PFC
30 11 7.3 PFC
20 11 60 x 500 KVA plus 11.0 PFC
200 x 50 KVA
The equivalent BHV is calculated based on the composition of cables in this typical distribution system
considering the cable capacitance at 33kV and 11kV. This is summarised in Table 3-11.
75
TB 766 - Network modelling for harmonic studies
BHV
''
Z shc XT
0,415 kV
M
RLV BLV
Figure 3-43 Downstream Harmonic Component Load Model Representation at Supply Bus GSP as used
by Scottish Power TSO
To represent the network supplied from the GSP, Scottish Power has adopted a different approach than
that used by National Grid for deriving the equivalent BHV, BLV and XT and additional Zsc” values:
1. BHV is calculated by assuming a circuit capacitance of 0.1 µF/km at each GSP using an estimated
total length of circuits for the distribution system being fed from that GSP;
2. BLV is calculated by setting it to 1% of the actual P MW load in Mvar;
3. XT is constant (not scaled) and is assumed to be at 13.4% on 12 x n MVA, where n is the number
of primary transformers at each GSP;
4. Zsc” represents the non-static part of the load (motor) and is scalable at 1MVA per MVA fault infeed
to give a more pessimistic result for low demand conditions.
76
TB 766 - Network modelling for harmonic studies
Three identical feeders were used, each having an average load composition. Only feeder 235 was
modelled in detail and its composition is given in Table 3-13. Each feeder has approximately 1.8 Mvar
of PFCs. The assumed load compositions are provided in
Table 3-14.
Feeder 235
Residential 31.0 %
Commercial 3.5 %
Industrial 62.0 %
Others 3.5 %
Total 11.6 MVA
77
TB 766 - Network modelling for harmonic studies
The industrial loads were comprised of 65% small motors and 35% large motors (constant MVA).
Feeders with loads were modelled in detail and due to non-linear elements (e.g. power electronic
devices), the frequency responses were obtained by time-domain simulations carried out in batch mode.
78
TB 766 - Network modelling for harmonic studies
79
TB 766 - Network modelling for harmonic studies
From the results, Method 6 (IEEE; “2 RL//”) seems to be the most accurate representation when
compared to the detailed feeder model. Method 1 (IEEE; “RL series”) and Method 2 (IEEE; “RL//”) do
not fit the impedance angle below 260 Hz.
The three load models (Method 1 (IEEE; “RL series”), Method 2 (IEEE; “RL//”) and Method 6 (IEEE; “2
RL//”)) from Figure 3-40 were then compared against the detailed feeder model for the case of the feeder
with light load. The frequency scan of the feeder with light load is shown in Figure 3-47.
80
TB 766 - Network modelling for harmonic studies
From the results, Method 6 (IEEE; “2 RL//”) seems to be the most accurate representation when
compared to the detailed feeder model. Method 1 (IEEE; “RL series”) and Method 2 (IEEE; “RL//”) do
not fit the impedance angle below 250 Hz.
81
TB 766 - Network modelling for harmonic studies
Figure 3-48 Frequency Response as seen from the Wind Farm (120kV)
Key Points:
There is an important damping effect at resonance even under light load;
It is important to represent the load feeders to obtain an accurate resonance frequency.
82
TB 766 - Network modelling for harmonic studies
Figure 3-49 Model Comparison as seen from the Wind Farm (120kV)
Key Points:
With feeder representation, all models reproduce the correct resonance frequency.
The damping effect of the load is not well represented by any of the 3 models.
83
TB 766 - Network modelling for harmonic studies
3.4.2.1.6 Results
This section presented an example in which Method 1 (IEEE; “RL series”), Method 2 (IEEE; “RL//”) and
Method 6 (IEEE; “2 RL//”) were compared. The key points observed from the study results are:
Key Points:
Method 6 (IEEE; “2 RL//”), without being perfect, was found to most accurately model the load
behaviour with respect to the frequency.
Detailed feeder and load modelling are the only way to precisely study the damping effects of the
load.
Consideration of the non-linearities introduced by components such as diodes in the power supplies of
televisions, computers and compact fluorescent lighting, may sometimes require that load modelling be
performed in the time-domain in batch mode.
84
TB 766 - Network modelling for harmonic studies
Figure 3-51 Extended CIGRE 14-Bus System for Comparison of Harmonic Load Models
85
TB 766 - Network modelling for harmonic studies
This section presents the frequency impedance characteristics of the harmonic load models at the
busbars where the loads are connected (at the 33kV distribution level) eventually stepping up two
voltage levels through the grid transformers (110kV and 220kV) to the transmission system. A
combination of OHL and cable circuits were extended from the 110kV bus and modelled to represent a
simple connected 33kV distribution network as detailed in Table 3-15. Three loads at 33kV were
modelled and connected at “Bus3 33kV”, “Bus4 33kV” and “Bus5 33kV”, respectively. The profiles of
the loads representing winter, autumn and summer are provided in Table 3-16.
3.4.3.1 Comparison of Results Two Voltage Levels above Load Placement (at
“SB4 220kV”)
The results for the different load models show no observable difference for the harmonic frequencies
seen at 220kV for all demand levels (frequency scans of impedance are shown in Figure 3-52 to Figure
3-54). Only Model 1 using the series R and L was observed to show reduced damping when demand
levels decrease at the 11th harmonic order (Figure 3-54 shows a higher peak than Figure 3-53).
86
TB 766 - Network modelling for harmonic studies
3.4.3.2 Comparison of Results One Voltage Level above Load Placement (at
“Bus1 110kV”)
At “Bus1 110kV”, the effects of the different load models at harmonic frequencies for all demand levels
(frequency scans of impedance are shown in Figure 3-55 to Figure 3-57) can be seen.
Model 1 using the series R and L showed significant differences as the demand levels reduced, with a
much higher parallel resonance around the 35th harmonic order for the autumn load, shifting to around
the 32nd for the summer load. This shift is expected due to the increased inductive component of the
load, and because this model is the only series-connected model (the others are parallel-connected).
This affected the resonance noticeably at the higher harmonic orders (above the 22nd in our test CIGRE
network model) when the demand levels were lower (Figure 3-56 and Figure 3-57).
During the winter load, as shown in Figure 3-55, the UK utility load model (NG Model) is largely in
agreement with Models 2 - 6 but has introduced an additional peak resonance near the 17th harmonic
order. This is due to the capacitance representing the cables and power factor correction in the
distribution network.
87
TB 766 - Network modelling for harmonic studies
88
TB 766 - Network modelling for harmonic studies
89
TB 766 - Network modelling for harmonic studies
90
TB 766 - Network modelling for harmonic studies
91
TB 766 - Network modelling for harmonic studies
92
TB 766 - Network modelling for harmonic studies
93
TB 766 - Network modelling for harmonic studies
3.4.3.7 Results
From the results comparisons, it was seen that different load models used will have a significant
influence on harmonic characteristics depending on (i) where the models are connected (e.g. connected
at MV and influence at EHV) and (ii) which models are used. The key points observed from the study
results are:
Key Points:
The CIGRE and IEEE load models need to include other distribution components in the transmission
network when used to represent the demands on a distribution system. If specific knowledge of load
composition at the connected node is known, the appropriate CIGRE and IEEE models can be used.
Appropriate load models should be used as these can affect what is seen at EHV and HV levels
even when connected at MV (e.g. at 33kV in the 14-bus IEEE model). The differences between the
load models seen at EHV and HV buses diminish as the number of transformation levels increases.
Different load models are required to represent the various categories of loads in the network
(distribution loads, industrial loads, power station loads etc.).
The utility aggregated load model used for the UK represents the typical load composition of a
distribution load in an urban area. The various parameters for the load model are therefore not
representative for different regions and countries.
3.4.4 Summary
Providing the engineer with clear guidelines to develop equivalent harmonic load models is a non-trivial
task.
Considering load modelling from a system point of view, lumping load at a certain voltage level (probably
in the range of MV), the following should be considered:
i. According to the different operating conditions of the network (light/peak, summer/winter), the
equivalent lumped load could change not only in terms of rated power but also in terms of load
type (for example modelling operating conditions with different percentages of rotating loads
with respect to static loads);
ii. The equivalent lumped load, seen from a network bus, could consist of different contributions
to be used to represent the main characteristics of the distribution network. For example, a
lumped load could consist of motors, static loads and shunt capacitor banks of the distribution
network (or cables). These contributions can have different weightings according to the different
operating conditions of the network.
The equivalent load is a particular component of the network model, and it can assume different
operating states to model the different operating conditions of the distribution network. This factor can
94
TB 766 - Network modelling for harmonic studies
result in many cases for harmonic studies (to be combined, for example, with different contingencies).
Guidelines for equivalent load representation should also consider consequences regarding the number
of cases that will then need to be studied. This number will be largely determined by a combination of
load conditions, contingencies and other possible factors.
Loads usually have an important damping effect near harmonic “emitting” installations such as SVCs,
HVDC and wind farms. Analyses presented in this section highlighted the difficulties in obtaining an
accurate assessment of the damping using various frequency-domain models. For a highly accurate
assessment of the damping effect of loads, detailed time-domain modelling may occasionally be
preferable. Neglecting loads may result in a pessimistic evaluation of harmonic distortion and
subsequent over-filtering.
The models described in this section are linear load models. A significant increase in power electronic
type loads is expected (in the form of embedded PVs, EVs and standalone PVs, wind farms, charging
sites for EVs at distribution levels, data centres, etc.), meaning that it will become increasingly important
to consider the non-linear effects of these types of loads in the models. This is an area where further
research work is required to develop enhanced load models that accurately capture the behaviour of
these new types of devices distributed throughout the power system.
Detailed modelling of loads is a very difficult task primarily because of the lack of information about the
load composition, type of feeders (overhead or underground) and locations of distributed power factor
correction capacitors. Correct consideration of the impacts of loads in harmonic studies is a large task
and requires time-consuming efforts in gathering information and performing component modelling.
Depending on the specific installation, it is up to the engineer to judge whether such efforts are
economically justified.
As the rotating magnetic field created by stator harmonics rotates at a speed significantly higher than
that of the rotor, the harmonic impedance of the generator approaches its negative sequence impedance
[8]. The IEEE Task Force on Harmonic Modelling and Simulation suggests that the synchronous
machine harmonic reactance be taken to be either the negative sequence impedance or the average of
95
TB 766 - Network modelling for harmonic studies
the direct and quadrature sub-transient reactance. The generator reactance is commonly represented
by [8]:
(𝑋𝑑" + 𝑋𝑞" )
𝑋𝑔𝑒𝑛ℎ = ℎ ⋅ 𝑋2 = ℎ ⋅ Equation 3.44
2
Various expressions are available to describe 𝑅𝑔𝑒𝑛ℎ which is highly frequency-dependent. Figure 3-71
illustrates the variation of 𝑅𝑔𝑒𝑛ℎ and 𝑋𝑔𝑒𝑛ℎ as a function of harmonic order, derived by the manufacturer
of a 370 MVA salient pole generator owned by Hydro-Quebec (HQ). The sub-transient reactance of this
machine is 0.24 pu at 13.8 kV.
Figure 3-71 Sample 370 MVA Synchronous Generator Harmonic Equivalent Resistance, Rgenh, and
Reactance, Xgenh
The following subsections describe three common approaches used to represent the harmonic
impedance of synchronous generators:
96
TB 766 - Network modelling for harmonic studies
Where R2 and X2 are the negative sequence resistance and reactance, respectively, measured at power
frequency.
Measurement data shows that the resistance, R h , is highly frequency dependent. An α value of 0.5 is
theoretically correct, however a value of 1.0 was verified through limited tests [59].
3.5.1.3 Model 3: HQ
The Hydro Quebec (HQ) model is an in-house model developed to fit manufacturer data shown in Figure
3-71. It models an equivalent passive shunt impedance and is applicable in both the frequency domain
and time domain.
97
TB 766 - Network modelling for harmonic studies
𝑋
𝑅1 = 𝑍𝑏𝑎𝑠𝑒 ⋅ [𝛺] Equation 3.51
10
𝑋
𝐿1 = 𝑍𝑏𝑎𝑠𝑒 ⋅ [𝐻] Equation 3.52
2 ⋅ 𝜋 ⋅ 𝑓nom
𝑅2 = 𝑍𝑏𝑎𝑠𝑒 ⋅ 𝑋 ⋅ 100 [𝛺] Equation 3.53
20
𝐿2 = 𝑍𝑏𝑎𝑠𝑒 ⋅ 𝑋 ⋅ [𝐻] Equation 3.54
2 ⋅ 𝜋 ⋅ 𝑓nom
𝑅 = 𝑍𝑏𝑎𝑠𝑒 ⋅ 𝑋 ⋅ 500 [𝛺] Equation 3.55
1000
600
800
400 600
400
200
200
0 H 0 H
0 5 10 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
base case IEEE-ALPHA_0.5 IEEE-ALPHA_1 base case IEEE_ALPHA_0.5 IEEE_ALPHA_1
IEEE-ALPHA_1.5 ELECTRA_167 HQ IEEE_ALPHA_1.5 ELECTRA_167 HQ
Figure 3-75 Effect of Detailed Synchronous Generator Models on Harmonic Impedance in a Real
Transmission System (a) Close to Generation and (b) Far from Generation
To better compare the models, damping factors were calculated as the ratio between the harmonic
impedance obtained with each model to the harmonic impedance obtained in base case, with a constant
generator resistance. A summary of the damping ratio factors provided by each model is presented in
Figure 3-76 for each of the 18 transmission nodes. In this example, the Electra 167 model gave damping
ratio factors close to 1 at all transmission nodes, indicating that this representation is almost equivalent
to neglecting frequency dependency of resistive loses (i.e. base case). On the other hand, the IEEE
model with an exponent of 1.5 gave the lowest damping ratio factors, ranging from 0.58 to 0.94
depending on the transmission node. This indicates that, in the most extreme case (transmission node
#17 – conventional generation plant), the choice of model could result in a reduction of the harmonic
impedance by almost half. Those nodes electrically close to synchronous generation (e.g. nodes from
98
TB 766 - Network modelling for harmonic studies
#12 to #17) are most affected by the choice of model, as expected. Therefore, when performing
harmonic analysis in areas close to synchronous generation, the use of frequency-dependent data from
manufacturers should help obtain the most accurate results.
1
0.95
(with respect to Base Case)
0.9
0.85
Damping Factor
0.8
0.75
0.7
0.65
0.6
0.55
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18
Transmission Node
IEEE_alpha_0.5 IEEE_alpha_1 IEEE_alpha_1.5 Electra_167 HQ
Figure 3-76 Maximum Damping Factor (with Respect to Base Case) Provided by each Synchronous
Generator Model
3.5.3 Summary
Three models were described in this section, and case studies presented highlighting various important
aspects of synchronous generator modelling. The value of the alpha exponent used in the IEEE model
should be selected appropriately. The accurate modelling of the (highly) frequency-dependent
resistance, R, depends on having accurate data for R and on the exponent used. Where possible it is
highly recommended in [59] to obtain this data either via the manufacturer or via measurement.
A comparison of the models via frequency scans using a model of the Irish transmission system suggest
that the selection of a synchronous generator model only has a noticeable effect at the parallel resonant
points and at nodes that are electrically close to synchronous generators. The IEEE model with an alpha
exponent of 1.5 provides the highest level of damping and the Electra 167 model provides the lowest
level of damping. It is expected that this model yields the most pessimistic results as it considers √ℎ.
Aspects such as model validation, sensitivity analysis and model simplification were illustrated in
Appendix F via measurement data and simulation results. The accuracy of the HQ model was confirmed
via measurements; the importance of modelling high frequency losses was shown, and the accuracy of
a simplified model demonstrated.
99
TB 766 - Network modelling for harmonic studies
100
TB 766 - Network modelling for harmonic studies
500
Z [] Natural resonance in
network without
400 capacitor banks or with
detuned capacitor
Shifted resonance
banks
due to capacitor
300 bank without
detuning
Capacitor
200 bank
detuned
frequency
100
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20
Harmonic Order
Without Capacitor Banks With 30MVAr Capacitor Bank
With detuned 30MVAr Capacitor Bank With detuned 30MVAr Capacitor Bank - no damping R
The effect of shunt inductive compensation on the harmonic performance of a transmission line has
been illustrated in [3]. It has been shown that, while the shunt compensation improves the voltage profile
at the fundamental frequency, it will also affect the harmonic profile. It is therefore important to represent
these shunt devices in the models used for harmonic propagation studies.
Key Points:
Shunt reactors should be accurately represented in harmonic studies;
Stray capacitances can usually be neglected.
101
TB 766 - Network modelling for harmonic studies
C1 C1 C1
C1
R1 L1 R1
L1 L1 R1 L1
R1 R2 L2 C2 R2 L2 C2
R3 L3 C3
Key Points:
Resistive losses and manufacturing tolerances should be included if data is available.
Representation of the bypass circuit is not normally required, unless tuning of this circuit needs to
be investigated due to high background harmonic.
102
TB 766 - Network modelling for harmonic studies
3.6.7 Summary
The modelling of shunt capacitors, shunt reactors, series capacitors and series reactors as simple
lumped components was discussed in this section. The effect of capacitor detuning and resistive losses
modelling on the harmonic impedance was demonstrated through an example using a shunt capacitor
in a real transmission system model. The effect at low-order harmonic frequencies was highlighted,
showing that modelling as accurately as possible is important. Passive harmonic filters are modelled as
combinations of lumped components, and typical filter configurations were illustrated.
103
TB 766 - Network modelling for harmonic studies
104
TB 766 - Network modelling for harmonic studies
105
TB 766 - Network modelling for harmonic studies
106
TB 766 - Network modelling for harmonic studies
The following subsections describe the most common types of converter-based generators and their
modelling on an individual device level. Aggregation of multiple components in a wind farm / PV farm is
covered in chapter 5, sections 5.6 and 5.7.
DFIGs are widely employed in MW-range wind turbines to facilitate variable speed operation. The DFIG
is essentially a wound rotor induction machine fed by both the stator and rotor. It is comprised of a back-
to-back voltage-source power converter: a rotor-side converter and a grid-side converter. The stator of
the DFIG is directly connected to the grid, and the two converters are included in the rotor circuit [64].
The grid-side converter regulates the DC bus voltage and can absorb and generate reactive power [66].
The rotor-side converter is responsible for controlling the active and reactive power output of the
generator. The DFIG is illustrated in Figure 4-2.
107
TB 766 - Network modelling for harmonic studies
The presence of power converters based on forced-commutated power electronics causes the
production of harmonics and interharmonics. Harmonics are produced in DFIGs via three predominant
mechanisms [66]:
1) Grid-side converter: fast switching from the current-controlled VSC produces high-frequency
harmonics and interharmonics;
2) Rotor-side converter: low- and high-order rotor harmonic components propagate to the grid;
3) DFIG windings: high frequency harmonics in the generated voltage due to the non-linear effect in
the air-gap flux and slip.
108
TB 766 - Network modelling for harmonic studies
Key Points:
An “ideal” current source representation is not adequate to capture the interactions between Type 3
WTGs and the harmonic network impedance.
The use of a Norton/Thévenin equivalent model is recommended (Figure 4-1). This model must
capture the input impedance of the device comprising a “passive” part (determined by the grid-side
filter circuit elements) and an “active” part (determined by the internal control loops).
Harmonic filters associated to the converters need to be captured in the model.
The wind speed (and therefore torque) should be considered as it determines the harmonics
produced by the DFIG [69].
109
TB 766 - Network modelling for harmonic studies
(Norton/Thévenin equivalent circuit) is shown in Figure 4-1. The Norton/Thévenin equivalent circuit is
comprised of an equivalent ideal current/voltage source and equivalent impedance for each
harmonic/frequency order of interest. This provides a common structure for various turbine/converter
vendors, regardless of their specific power converter designs. These two equivalent models are
theoretically identical, and interchangeable in terms of harmonic sources and harmonic impedances.
Regardless of the model used, the harmonic sources and harmonic impedances should be provided at
the frequencies of interest for the intended application.
The harmonic emission profile of a type 4 wind turbine is typically a function of operation points. To
simplify modelling and reduce the amount of input data required, the worst-case operating point at each
harmonic order is often used to yield a conservative result. However, the proposed model structure
allows for the implementation of look-up tables to represent the harmonic behaviour at different
operating points. This approach can be used for probabilistic analysis or sensitivity studies of various
operating points.
The equivalent harmonic impedance is the total converter output impedance which should include the
impact of closed-loop controls. Any other passive components, such as the smoothing reactor, turbine
filters and transformers need to be captured either as part of the equivalent harmonic impedance or
explicitly as an external component.
Unfortunately, it is not easy (especially at lower frequencies) to measure the frequency-dependent
impedance of passive components, which could introduce modelling errors, especially for harmonics
having a low magnitude. In addition, multiple parallel-connected converters are often used in
applications of MW power density, and these converters may introduce unbalance which can affect
internal harmonic current flow between converter modules. Even if coupled sharing reactors are
designed to limit the current imbalance, some asymmetry in harmonic generation between converter
modules can be seen. The manufacturing tolerances of electrical components such as reactors,
capacitors and resistors, must be considered in the uncertainty/tolerance of the modelling.
3
TenneT has included the requirement to provide prevailing phase angle information in the German
implementation of the European HVDC Connection Code (TAR 4131).
110
TB 766 - Network modelling for harmonic studies
Figure 4-4 Example Comparison of WT Harmonic Impedance and Smoothing Reactor of Converter
111
TB 766 - Network modelling for harmonic studies
Ih Zh Filter Ih Ih Filter
A) Complete harmonic model B) Ideal current source C) Current source with passive filters
Figure 4-6 shows a comparison of the three WTG modelling approaches in a frequency scan of the
entire wind farm as seen from the HV side of the wind farm transformer. The internal collector network,
WTG transformers and WTGs are individually represented. It should be noted that in this example the
wind farm is not connected to an external grid in order to highlight the differences observed at the
Connection Point without any masking effects from the external grid harmonic impedance. Approach A
(complete harmonic model) is the most accurate and is therefore defined as the reference case. Figure
4-6 shows that the simulation results produced with the other two modelling approaches fail to reflect
the correct interactions between the WTGs and the system under study (wind farm internal collector
network). In this example, the effect of the converter internal impedance is dominant in the low frequency
range, therefore modelling approaches (B) and (C) are not adequate in that range. It can be concluded
that approaches B (ideal constant current source) and C (constant current source with passive filters)
can yield incorrect results when studying the impact (and interactions) of this wind farm on a power
system.
Key Points:
Converter harmonic impedance and filters should be included when modelling type 3 and type 4
WTGs. Therefore, the use of Thévenin/Norton equivalents are recommended instead of the
constant current source approach.
The user should request detailed WTG models capturing these elements from vendors to improve
accuracy in harmonic studies involving wind farms.
112
TB 766 - Network modelling for harmonic studies
Impedance [Ohm]
100
10
1
0 500 1000 1500 2000 2500 3000
Frequency [Hz]
50
0
0 500 1000 1500 2000 2500 3000
-50
-100
Frequency [Hz]
Figure 4-6 Frequency Scan of a Wind Farm with different WTG Harmonic Models
113
TB 766 - Network modelling for harmonic studies
4.2.4 Summary
Power converters are a source of harmonic distortion. The converter also acts as frequency-dependent
impedance which can impact the resonance frequencies seen in the network. To be able to fully assess
the impact of converter-based generation on the harmonic performance of the grid it is essential that the
models capture both the harmonic emission and the impedance characteristics of the power converter.
It is recommended to model converter-based generation as a Norton or Thévenin equivalent circuit at
the frequencies of interest (typically 2 nd to 50th harmonic order), in accordance with IEC 61400-21-3 [62].
The equivalent circuit consists of an equivalent voltage/current source and a harmonic impedance. The
harmonic impedance must capture the power converter impedance characteristics; it may also include
passive components such as the main reactor impedance, passive filters and turbine transformers,
which will otherwise have to be modelled separately.
114
TB 766 - Network modelling for harmonic studies
The emission and impedance characteristics depend on the operating point of the power converter,
where the worst-case operating point may differ for different harmonic orders. However, for simplicity,
the worst-case for each harmonic order is often used to yield conservative results.
The above detailed harmonic model needs to be provided by the manufacturer (or designated party) to
the user (system operator, wind farm developer, consultant, etc) to facilitate harmonic studies associated
with the connection of the converter-based generation devices to the grid. Some system operators are
already including the provision of these models (and associated documentation) as a pre-connection
requirement within the connection offer process.
The reader is reminded that the above considerations apply to the modelling of converter-based
generation on an individual device level. Aggregation of multiple components in a wind farm / PV farm
is covered in chapter 5, sections 5.6 and 5.7.
The following sections first give a brief description of the main elements and behaviour of an HVDC
converter station, since they are relevant to understanding of its harmonic characteristics. Subsequently,
the important aspects to be considered in modelling are highlighted.
115
TB 766 - Network modelling for harmonic studies
reactive interchange with the AC network within specified limit bands (see Figure 4-8). Hysteresis is
normally introduced in the switching sequence to avoid excessive switching during minor variations in
converter power.
0,8
Q Converter
0,6
Q Filters
Reactive Power (pu)
116
TB 766 - Network modelling for harmonic studies
0,05
I11
0,03
I13
I23
0,02
I25
0,01
I35
I37
0
0,00 0,20 0,40 0,60 0,80 1,00
DC Power (pu)
* pu of fundamental current for nominal DC power
Figure 4-9 Illustrative Variation of Characteristic Harmonics Magnitude vs Load (HVDC LCC)
Key Point:
For initial basic analysis, the converter is often regarded as a harmonic current source, that is, the
harmonic current emission does not vary with the AC side impedance. While this simplification gives
approximately correct results at higher orders, it can be unsatisfactory for low orders, particularly for
the 3rd harmonic. Therefore, a more complex analysis is necessary for such low order harmonics,
as discussed further below under “cross-modulation”.
117
TB 766 - Network modelling for harmonic studies
must be considered together with the effect of network frequency variations for an accurate assessment
of the HVDC LCC installation (see example in next page). Variation of resistance with temperature will
affect the filter damping.
It should be noted that the above considerations are normally only captured in the detailed design and
rating studies performed by the manufacturer or specialist designer of the HVDC LCC installation.
System-wide harmonic studies are not normally concerned with these minor parameter variations and
filters are typically represented with their nominal values. However, the user (system operator or
consultant) may deem prudent to include the filter detuning aspects when performing studies in an area
close to a HVDC LCC installation, especially if resonances are expected near the filter tuned
frequencies.
SUB-BANK SUB-BANK
BRANCH BRANCH
BANK SUB-BANK
Key Point:
Reactive Power compensation devices (especially switchable filters and/or capacitors) need to be
captured in the HVDC LCC converter model as they will interact with the harmonic impedance of
the grid. A range of operating points, and associated status of filters/cap banks, may need to be
analysed to ensure that they do not introduce excessive resonances or harmonic distortion in the
grid.
118
TB 766 - Network modelling for harmonic studies
10000
1000
|Z| Ω)
100
DT (11th, 24th) + TT
10 (13th, 36th, 48th)
1
0 10 20 30 40 50
Harmonic Order
For this particular system, the connection of an additional triple-tuned filter for the 3rd, 5th and 11th is
required in order to meet harmonic performance at maximum DC transfer. Figure 4-12 shows the
impedance of this filter.
10000
1000
|Z| Ω)
1
0 10 20 30 40 50
Harmonic Order
Figure 4-12 Harmonic Filter Impedance - Triple-Tuned Filter
Figure 4-13 shows the resulting impedance when having all the filters required for maximum DC
transfer connected to the system. This plot also shows the impedance change for +3% and -1.5%
filter detuning, which were the maximum values taken into consideration to study the harmonic
performance of this sample HVDC system during the design and rating phases.
119
TB 766 - Network modelling for harmonic studies
4.3.1.4 Cross-Modulation
The assumption that an HVDC LCC converter may be treated as a pure harmonic current source is a
convenient one for practical analysis and is frequently used. For practical design purposes, the errors
introduced by this assumption may be typically around 10% at the 11 th and 13th harmonics and at higher
harmonics much lower still.
The treatment of the converter as a pure current source rests on the assumption that the direct current
is perfectly smooth. In practice however, it always has some “ripple”, that is, a harmonic content. The
mechanism by which the AC and DC side harmonics are related is known as “cross-modulation” and is
discussed in CIGRE Technical Brochures 143 [83] and 553 [23] and other numerous publications.
Analysis of cross-modulation effects is a complex matter, as there exists a circular interaction between
the harmonic voltages, currents and impedances on the AC and DC sides of the converter. In addition,
the transfer function of the converter varies with the operating point and load level. There are methods
in the frequency domain which provide approximate results, and the harmonic domain approach (section
2.5) aims to achieve a more accurate solution. However, to fully account for all the factors involved, and
to take into account the effect of the HVDC LCC converter control, a time domain solution (section 2.3)
using a full representation of the converter and its control system is often the most appropriate approach.
Ignoring cross-modulation effects will have the greatest impact at the low order non-characteristic
harmonic interaction resulting from AC system voltage unbalance, that is, the presence of a small
negative sequence component in the supply voltage. This results in generation of DC side 2 nd harmonic
voltage and consequently current, which in turn is cross-modulated back to the AC side as 3rd harmonic
current of positive sequence (and negative sequence fundamental). The magnitude of this 3 rd harmonic
current can be highly significant, with major implications for whether specific 3 rd harmonic filtering is
required.
When considering cross-modulation, it is no longer valid to assume that the highest voltage distortion
will occur at the highest value of total AC side harmonic impedance (filter plus network). As the cross-
modulation analysis has to consider the complete system of AC and DC side impedances, it is not
possible to make a general statement about what AC side impedance will produce the highest distortion.
This is particularly important for the 3rd harmonic, where the need for filtering may depend on such
assumptions.
120
TB 766 - Network modelling for harmonic studies
The magnitude and variation with load of converter harmonic generation, and the detailed design of the
filters, are matters which cannot be predicted at the planning stage of an HVDC LCC project. They will
only become clear as part of the integrated design of the station, and will depend on such factors as the
choice of converter transformer reactance, the permitted harmonic distortion levels, and the AC network
harmonic impedance. The design will also depend on economic factors, physical layout constraints and
outage requirements. All this implies that it is especially difficult to incorporate a future HVDC LCC
station in a network harmonic model at the planning stage. Very general assumptions regarding
harmonic generation and filtering have to be made.
The following two sections discuss modelling aspects from the point of view of an Owner, TSO,
consultant or other third-party who may not have access to all the detailed design information available
to HVDC suppliers. These users may need to model an HVDC LCC system for a harmonic study or may
simply wish to acquire a basic understanding of detailed design studies performed by technology
providers.
121
TB 766 - Network modelling for harmonic studies
PCC
I LCC_n Z AC Z AC
Filters System
Figure 4-14 Simplified Equivalent Circuit for Harmonic Emission Assessment at PCC (HVDC LCC)
Normally it is necessary to study different HVDC LCC configurations and output levels. For each of these
possible scenarios, the following considerations must be made:
Converter Harmonic Source (ILCC_n): Typically, the complete HVDC LCC converter, including
converter transformers, is treated as an ideal current source. The level of detail that can be attained
when constructing a model is limited by the data provided by the HVDC LCC supplier. In general,
the technology providers are required to make available, as a minimum, the highest levels for each
individual harmonic over the full range of operation – known as a “non-consistent” set of harmonics.
This may be a conservative approach since the worst levels of each harmonic do not always occur
at the same operating level or configuration, so the total will give an unrealistically high THD.
Additionally, the value of all individual harmonics at a defined set of operating conditions may be
provided (e.g. at 10% steps of transmitted power.) If such more detailed data is available, it must
be decided whether it is required to model the specific harmonic levels for each desired scenario or
if the assumption of coincident worst levels is sufficient. For example, each harmonic source can be
modelled with a look-up table to produce the right output based on the DC transfer level. If no data
is available, typical idealized characteristic harmonics can be calculated for the desired operating
122
TB 766 - Network modelling for harmonic studies
conditions using well-known equations for the converter, as provided in many references (e.g. Part
1, Annex B of [84]).
AC Filter Impedance (ZAC Filters): This impedance must be adjusted depending on the number of
filters in service for a given HVDC LCC configuration and DC transfer level. The different impedance
values are usually available in the Harmonic Performance Study report provided by the supplier.
The calculations carried out to determine these values typically already account for filter detuning
and frequency variations, but it must be verified if the worst-case conditions assumed in the
supplier’s study are still valid. Alternatively, the values of all filter components and the detuning
coefficients may be provided, which would enable an accurate model to be created.
The external system impedance (ZAC System) is normally provided by the system operator in the form of
harmonic impedance envelopes (see Section 5.4).
PCC
Z Z
Transformer AC System
Z Ufn
LCC Z AC
Conv. Filters
Figure 4-15 Simplified Circuit to Assess Magnification (or Damping) of Background Voltage Distortion at
PCC (HVDC LCC)
For most studies, it may suffice to model the AC filter impedance (ZAC Filters) under the assumption that
the impedances of the converter and the transformer are much higher than those of the filters and can
be assumed to be an open–circuit in the model. Typically, this simplified approach might be used to
calculate the harmonic voltage distortion at the PCC due to background harmonics, and the component
rating stresses due to background harmonics.
This simplified approach, ignoring the converter harmonic impedance, would provide a general
assessment of the converter station as a harmonic load, but the results may not be very accurate when
considering low order harmonics. More precise calculations may be desirable, not only for normal design
or compliance studies but also for analysis of particular behaviour in the low-order harmonic range, and
can be achieved through detailed consideration of the internal impedance of the converter, which
consists of two elements:
Converter Transformer (ZTransformer): Information on the transformer impedance should be available
in the design documentation. Tap-changer range should be included. See section 3.3 for guidelines
on transformer modelling.
Converter impedance (ZLCC Conv): This includes the impedances of the DC line, DC filters, converters
valves, smoothing reactor, etc., transferred to the AC side by the cross-modulation function of the
conversion process. For cases in which the DC side can be considered decoupled, it may be
sufficient to model the time average commutating impedance per bridge considering overlap. If the
123
TB 766 - Network modelling for harmonic studies
converter impedances for different conditions are not provided by the supplier, it is possible to
calculate the required impedance knowing certain parameters such as the commutating reactance
per bridge and the overlap angle [85], [86].
The external system impedance (ZAC System) is normally provided by the system operator in the form of
harmonic impedance envelopes (see Section 5.4). The pre-existing background distortion (Ufn) is also
provided by the system operator and it is normally obtained from measurement campaigns or
simulations (see Section 5.8).
PCC
Z
Transformer
Z
I LCC_n LCC Z AC
Conv. Filters
124
TB 766 - Network modelling for harmonic studies
voltage. This configuration permits using lower switching frequencies compared to PWM, which also
leads to lower losses.
Figure 4-17 shows the typical configuration of an MMC VSC converter, as well as the two main types of
modules used. The half-bridge configuration has the lowest number of components (which also means
lower cost and losses), but its output voltage only has one possible polarity. The full-bridge configuration
allows an output voltage with either polarity, which makes it possible to block the current from DC faults
and provides more flexibility at the expense of using more switching devices.
Figure 4-17 a) MMC VSC Configuration, b) Half-Bridge Module (c) Full-Bridge Module
125
TB 766 - Network modelling for harmonic studies
4.3.2.2 Cross-Modulation
Harmonic interaction between the AC and DC sides of an HVDC VSC occurs but is generally less
significant and more complex than for HVDC LCC. The existence of capacitive energy storage within
the converter tends to partially decouple the two sides of the converter in terms of harmonic interaction.
126
TB 766 - Network modelling for harmonic studies
PCC
Z
Z
converter
Transformer
reactor
Z VSC
Conv.
Z AC Z AC
Filters System
U VSC_n
PCC
Z
Z
converter
Transformer
reactor
Z VSC
Conv.
Z AC Z AC
Filters System
U VSC_n
Figure 4-18 Simplified Equivalent Circuit for Harmonic Emission Assessment at PCC (HVDC VSC)
The main aspects to be considered for each element of the circuit are discussed below.
Voltage harmonics source (UVSC_n): HVDC VSC-generated harmonics are in general of very low
magnitudes. With VSC technology we do not have the option to calculate characteristic harmonics
using well-known equations. The representation of the voltage source will solely depend on the
available data provided by the supplier, since the generated harmonics are unique to their
proprietary solution. The supplier can provide the harmonic voltages produced for each operating
point or, at the least, the maximum individual harmonic voltages considering a range of operating
points. Once again, it is important to consider that the maximum distortion for each frequency does
not occur at the same VSC operating point, so modelling the voltage source based on these values
will lead to conservative results and will not match measurements.
Converter impedance (ZVSC Conv): This impedance represents the VSC module elements, such as
the valves and the capacitors and the influence of the control system. In order to model this
impedance, the supplier must provide values for different operating points.
Converter reactor impedance (Zconverter reactor): Values for the reactor impedance are readily available
in the design documentation.
AC filters (ZAC Filters): AC filters are not required in many HVDC VSC installations using the latest
technologies. Where a filter is used, it normally consists of a single high-pass branch which remains
connected at all power transfer levels. Its impedance data may be readily obtained. Section 3.6.3
provides guidelines for the modelling of harmonic filters.
127
TB 766 - Network modelling for harmonic studies
Transformer impedance (ZTransformer): It may suffice to use the leakage reactance of the converter
transformer considering its tap-changer range. However, depending on the study objectives it may
also be necessary to include imbalances between phases and converter bridges. At high
frequencies (above 50th harmonic order), the transformer stray capacitances may become
significant and should be modelled. See section 3.3 for detailed guidelines on transformer modelling.
The external system impedance (ZAC System) is normally provided by the system operator in the form of
harmonic impedance envelopes (see Section 5.4). As mentioned above, HVDC VSC systems may
generate harmonics of very high order. If the effects of these harmonic flowing into the AC network are
of concern, then the AC system harmonic impedance beyond the 50 th harmonic must be taken into
consideration.
Amplification of background harmonic distortion caused by the HVDC VSC
Figure 4-19 shows an equivalent passive circuit diagram suitable to assess the modification of
background voltage distortion at PCC caused by the HVDC VSC harmonic impedance.
In this case, the source Ufn is used to represent the pre-existing AC harmonics to be studied. This is
provided by the system operator and it is normally obtained from measurement campaigns or
simulations (see Section 5.8).
The modelling aspects of the other circuit elements have been covered in the previous subsection.
PCC
Z
Z Z
converter
Transformer AC System
reactor
Z VSC Z AC Ufn
Conv. Filters
PCC
Z
Z Z
converter
Transformer AC System
reactor
Z VSC Z AC Ufn
Conv. Filters
Figure 4-19 Simplified Circuit to Assess Magnification (or Damping) of Background Voltage Distortion at
PCC (VSC HVDC)
128
TB 766 - Network modelling for harmonic studies
PCC
Z
Z
converter
Transformer
reactor
Z VSC
Conv.
Z AC
Filters
U VSC_n
PCC
Z
Z
converter
Transformer
reactor
Z VSC
Conv.
Z AC
Filters
U VSC_n
129
TB 766 - Network modelling for harmonic studies
4.3.3 Summary
Figure 4-21 summarises the basic guidelines to follow when modelling HVDC systems for harmonic studies.
HVDC Modelling
Guidelines for Harmonic
Studies
Use Use
Utility to provide Option to Use digital
manufacturer manufacturer
harmonic limits calculate LCC model capable
data to model data to model
and network characteristic of representing
generated converter
impedance data harmonics system controls
harmonics impedance
Model imbalance:
AC phases, More accuracy
transformer and needed
bridge impedance
130
TB 766 - Network modelling for harmonic studies
FACTS devices generate harmonics depending on their configuration, control strategy and switching
scheme. Therefore, certain device parameters must be taken into consideration and adequately
modelled for harmonic studies.
A comparison of the harmonic performance of the most common LCC and VSC FACTS devices (i.e.
SVC and STATCOM) is given below as a brief overview of the main aspects to consider with each
technology:
Table 4-3 Comparison of Harmonic Performance in Typical FACTS Devices
STATCOM SVC
Harmonic emission related to VSC Harmonic emission related to thyristor-
topology. based converter topology.
Harmonics related with frequency Harmonic emission is dependent on
converter modulation. operating point.
Converter control determines also the Significant power system characteristic
harmonic profile, i.e. harmonic harmonics due to non-linear nature of
contribution / attenuation and impedance semiconductor thyristors.
shaping. Possible stability issues for frequencies
Power system characteristic harmonics below the fundamental component.
due to nonlinear components. Possible resonances can be seen
Possible stability problems within between TSCs, TCRs and other wind
harmonic range as well as frequencies power plant components.
below the fundamental component.
STATCOM SVC
Optimized modulation strategy. Harmonic filters tuned to low-order power
Small harmonic filters tuned to system characteristic harmonics.
frequencies related with switching Possible extra need of harmonic filters or
operation. detuning of TSCs.
Possible active filtering in terms of 12-pulse (or higher pulse) configurations
compensation and damping. possible to limit filter size.
Can contribute to active damping of
harmonics generated by other
components
131
TB 766 - Network modelling for harmonic studies
The following subsections provide harmonic modelling guidelines for selected FACTS devices. Due to
their popularity, SVCs and STATCOMs are discussed in detail. The other FACTS devices are not as
widely used in power systems, but they are also briefly covered, and can be modelled using similar
approaches.
A TCR consists of a reactor in series with a switch often made of antiparallel thyristors. The reactor is
typically split into two parts, with the valves placed in between to protect them from short-circuit currents.
The TCR can be switched to smoothly vary the susceptance seen by the power system, effectively
controlling the reactive power injected to the network to maintain the required voltage set point. This is
achieved by delaying the firing angle of the thyristors to reduce the inductive current drawn by the
reactor.
The TSC/TSR is made of groups of capacitor/reactor banks that can be switched in and out of the
system, typically using a bidirectional thyristor switch. They usually include a current limiting reactor to
damp inrush current, reduce switching transients and prevent resonances with the AC system. SVCs
only having a TSC/TSR can simply provide stepwise control of the reactive power. However, the TSC
can be used to increase the capacitive range of the SVC, thereby providing a faster response compared
to mechanically-switched capacitors.
For SVCs having both a TCR and a TSC, the control of these two devices can be coordinated to provide
linear control of the reactive power support. In addition, the SVC has the capability to control
mechanically-switched elements to meet the network’s steady-state needs while saving the use of its
thyristor-based elements for a fast dynamic response. SVCs also usually require the use of permanently
connected filters to meet harmonic performance requirements.
Figure 4-22 shows a typical configuration of an SVC. The TSC and TCR are usually delta-connected to
trap the triplen harmonics in the delta. The filters and fixed shunts are generally connected in an
ungrounded wye arrangement to block zero-sequence and third harmonic currents from entering these
components. The SVC transformer is typically delta connected on the secondary side to prevent the
flow of zero sequence components into the grid.
132
TB 766 - Network modelling for harmonic studies
The following sections focus on the harmonic generation mechanism and characteristic harmonics of
an SVC as well as their modelling aspects. For a more in-depth overview of SVC types, characteristics
and overall performance capabilities, the reader is encouraged to consult CIGRE Technical Brochure
25 [87].
0.06
I5
0.05
I7
Harmonic Current (p.u.*)
0.04 I11
I13
0.03
I17
0.02
I19
0.01
0
90 110 130 150 170
Delay Angle (deg)
* pu of nominal (fundamental frequency) current
Since power system conditions are never ideal, there may be non-characteristic harmonics produced
by the TCR depending on the following conditions:
Even harmonics can be created by the asymmetrical firing of the antiparallel thyristors. These
harmonics are generally of much smaller magnitude than the odd characteristic harmonics, but in
some cases it may necessary to include the 2nd harmonic in the performance calculations.
133
TB 766 - Network modelling for harmonic studies
Triplen harmonics could flow into the network due to system unbalances, such as firing angle or
inductance differences between phases. The negative sequence component in the system
voltage will also result in triplen harmonics escaping the delta, with the 3rd harmonic having the
most significant impact. The magnitude of these harmonics depends on the severity of the
unbalance, so it is recommended to assess these harmonics taking into consideration the specific
network conditions and the TCR equipment tolerances.
PCC
Z(h)
Transformer
Figure 4-24 Simplified circuit for Harmonic Emission Assessment at PCC (SVC)
As discussed above, SVC configurations vary so not all the components shown in the generic equivalent
circuit (Figure 4-24) may be present in a given project. For example, an SVC may not include a TSC
and may only have filters on the secondary side of the transformer.
The maximum and minimum harmonic impedance values for each of the SVC branches are generally
provided in the Main Component Design report, which is part of the standard documents provided by
the manufacturer as part of the engineering design and review process. These values typically include
the worst-case combination of component tolerances and frequency variations.
The TCR-generated harmonic currents are not always provided by the supplier for different operating
points, but the Harmonic Performance Study should at a minimum provide the highest level observed
for each individual harmonic over the whole range of operation, typically at least up to the 25 th harmonic
134
TB 766 - Network modelling for harmonic studies
order. These maximum levels do not occur at the same operating point or SVC configuration, so the use
of this non-consistent set of harmonics will lead to a conservative THD assessment. Table 4-1 shows a
template that can be used to gather the required data to model the TCR generated harmonics.
Alternative formats, such as definition of current relative to the fundamental component of the AC line
side, are also possible. If no data is available, the ideal characteristic harmonics can be calculated for
the desired scenario using the equations for the TCR available in many references [3], [87].
The external system impedance (Z(h)AC System) is normally provided by the system operator in the form
of harmonic impedance envelopes (see Section 5.4).
Amplification of background harmonic distortion caused by the SVC
Figure 4-25 presents a simplified passive model that can be used to analyse the modification of the
background voltage distortion at the PCC caused by the SVC harmonic impedance interacting with the
power system impedance. A voltage source (Ufn) is used to represent the existing background
harmonics on the power system.
PCC
Z(h) Z(h)
Transformer AC System
Z(h) AC Z(h) AC
Z(h) Z(h) Filters Filters Ufn
TCR TSC (LV (HV
Side) Side)
Figure 4-25 Simplified Circuit to Assess Magnification (or Damping) of Background Voltage Distortion at
PCC (SVC)
As discussed in the previous subsection, the Main Component Design report should at least provide a
maximum and minimum value for the impedances of each SVC component. However, depending on
the objective of the study, it may be necessary to assess the impedances over the whole SVC operating
range; and this information may not be included in the report.
The TCR impedance varies depending on the firing angle and these values can be manually calculated
using the TCR equations [88], or they can be obtained using a software model of the SVC that includes
a representation of the controls.
The TSC may form a resonance with the system impedance that can cause large harmonic currents in
the SVC components and voltage distortions in the network. In the SVC design process, these
resonances must be taken into consideration for all harmonic orders of interest to ensure that the
equipment is rated to withstand the resulting currents. In some cases, it is necessary to include a filter
to damp these resonances.
135
TB 766 - Network modelling for harmonic studies
PCC
Z(h)
Transformer
Z(h) AC Z(h) AC
ITCR_n Z(h) Filters Filters
TSC (LV (HV
Side) Side)
A practical arrangement of the TCSC may consist of several units connected in series and includes
other elements, protection and control.
There are three modes of TCSC operation: (i) bypassed-thyristor mode when the thyristors fully conduct;
(ii) blocked-thyristor mode where the thyristor valves are blocked (do not conduct at all); and (iii) Vernier
mode which allows the TCSC to behave either as a continuously-controllable capacitive reactance or
as a continuously-controllable inductive reactance.
In capacitive Vernier control mode, the thyristors are fired when the capacitor voltage and capacitor
current have opposite polarity which causes a TCR current to have a direction opposite to the capacitor
current resulting in an increase of the equivalent capacitive reactance.
In inductive Vernier mode, the direction of the current circulating between the reactor and capacitor is
reversed and the controller presents a net inductive impedance.
The resulting equivalent reactance of the TCSC is given in Figure 4-28.
136
TB 766 - Network modelling for harmonic studies
TCSC operation with partial conduction will cause some harmonics to be injected into the power system.
The TCSC will generate all odd harmonics due to the operation of the TCR, which are well known. Most
of the harmonic currents will flow through the parallel capacitor bank and the effect on the power system
will be greatly attenuated – see Figure 4-29.
Harmonics are not expected to be an issue with TCSC applications, especially if several modules are
connected in series. Typically, only the lower-order harmonics (the 3rd, 5th, and 9th) dominate and need
to be considered.
Because of parallel combination of the TCR and a capacitor bank, the harmonic generation from a TCSC
appears as a low impedance voltage source in series with the line which drives the harmonic current
through the system. As TCSCs are likely to be applied to long lines, the impedance of the line results in
relatively small harmonic current.
The exception to the situation described above is when there is no capacitor and the installation is
reduced to a variable thyristor-controlled reactor. In this case, there will be significant harmonics injected
137
TB 766 - Network modelling for harmonic studies
into the power system. Based on existing TCSC installations, this type of installation is more theoretical
than practical.
As the TCSC would be connected to two electrically separate busbars of the power system, both
busbars will be affected by the harmonic injection of the device. To correctly model and calculate the
harmonic propagation and distortion in such a case, it would be necessary to have a detailed model of
the network or alternatively, a model of the network that will represent impedance at both busbars where
TCSC is connected as well as the mutual impedance between the busbars. Such an equivalent model
must represent all feasible outage conditions on the system.
SVC TCSC
Harmonic emission related to thyristor- Harmonic emission related to TCR
based converter topology. switching.
Harmonic emission is dependent on Harmonic emission is dependent on
operating point. operating point.
Significant power system characteristic Significant power system characteristic
harmonics due to non-linear nature of harmonics due to non-linear nature of
semiconductor thyristors. semiconductor thyristors.
TCR emissions can be represented as an TCR emissions can be represented as an
ideal current source. ideal current source.
Harmonic impedance is determined by Harmonic impedance is determined by
filters, TCR (as a function of the firing TCR (as a function of the firing angle)
angle), TSC and transformer impedance. and capacitors.
Possible resonances between TSCs, Typically, only the lower-order harmonics
TCRs and other power system (the 3rd, 5th, and 9th) dominate and need
components can drive the need for to be considered.
additional filters.
138
TB 766 - Network modelling for harmonic studies
VSC
VS ule
module
C
o d
m
VS ule
C
d
VSC
mo
module
VSC
mo SC
le
module
du
V
mo SC
le
du
V
The following sections focus on the harmonic generation mechanism and characteristic harmonics of a
STATCOM as well as their modelling aspects. For a more in-depth overview of STATCOM
characteristics and overall performance, the reader is encouraged to consult CIGRE Technical
Brochures 144 [90] and 663 [91].
139
TB 766 - Network modelling for harmonic studies
The input impedance model of the STATCOM may need to consider the PLL impedance shaping
effect when the Short-Circuit Ratio (SCR) is near or below 3. In this case, detailed information
needs to be provided by the vendor to represent cross-sequence and cross-frequency coupling.
These effects are difficult to capture in standard frequency domain analysis. If it becomes a
concern, more specialised tools and detailed converter models are required, but these are
normally only within the control of the STATCOM vendor
140
TB 766 - Network modelling for harmonic studies
harmonic voltages produced for each operating point or, at the least, the maximum individual harmonic
voltages considering a range of operating points. Table 4-1 provides a template for gathering the
required information.
As the harmonic generation is typically very low, it is normally sufficient to calculate a non-consistent set
of maximum harmonic voltages over the complete range of operation of the converter, and use this as
the modelled harmonic generation, represented as a voltage source behind the converter internal
harmonic impedance.
Harmonic Impedance
With reference to Figure 4-20, the harmonic impedance model must include not only the passive
components of the STATCOM (e.g. harmonic filter, series reactor, step-up transformer) but also the
VSC internal impedance defined by the operational mode, the controller transfer functions and its
corresponding parameters. The passive components normally dominate the equivalent input impedance
within the medium to high frequency range. However, for low frequencies within the bandwidth of the
converter inner current control loop, the influence converter impedance may dominate. Therefore, the
harmonic impedance is operating point dependent in the low frequency range (typically up to the 5 th
harmonic order).
In order to assess the interactions between the STATCOM harmonic impedance and the power system,
the vendor needs to supply the harmonic impedance for a range of operating points.
141
TB 766 - Network modelling for harmonic studies
As the device utilises two converters, both need to be accounted for in harmonic studies. Harmonics
generation is dependent on the type and topology of the converters, as described in previous sections.
STATCOM SSSC
Represented as a harmonic voltage Harmonic voltage source connected in
source behind the converter internal series with the line impedance.
harmonic impedance.
Harmonics related with frequency UPFC
converter modulation. Harmonics can be generated by either
Converter control determines also the the series or the shunt connected
harmonic profile, i.e. harmonic converter (or both).
contribution / attenuation and impedance For simplicity, harmonic generation can
shaping. be represented by a single harmonic
Harmonic impedance is operating point voltage source combining the effect of
dependent and needs to be supplied by both converters.
the vendor for a range of conditions.
Harmonic generation is typically low. IPFC
Assessment of a non-consistent set of
maximum harmonic voltages (to be Harmonic voltage sources connected in
provided by the vendor) is normally series with each of the lines’ impedance.
sufficient.
142
TB 766 - Network modelling for harmonic studies
Thus, the grid converter can improve the harmonic distortion either by active harmonic current
contribution (e.g. harmonic current compensation) or by impedance shaping (e.g. resonance damping).
Whether the active harmonic filter is series- (e.g. WT, HVDC) or shunt- (e.g. STATCOM) connected, it
can be represented for harmonic studies as an equivalent Thévenin or Norton circuit where the voltage
or current source (respectively) will represent APF harmonic contribution and the associated impedance
will reflect the converter control scheme.
Typically, APFs can be modelled in the harmonic or frequency domain to perform standard harmonic
propagation studies. However, some harmonic controllers are difficult to linearize, due to their highly
non-linear behaviour (e.g. limiters, instability detection, auto tuning). For such systems, detailed
modelling using time-domain or real-time simulations are more adequate. APF’s can fulfil a given control
function by minimizing the error between measurements and a harmonic reference. In this case, they
act as a controlled source for that specific frequency, e.g. harmonic current compensation with defined
reference current can approximately be considered at that harmonic frequency as a current source with
current magnitude equal to the setpoint / reference value.
It should be noted that active filtering is an area of ongoing work and development. For further
information the reader is referred to [92] to [98].
143
TB 766 - Network modelling for harmonic studies
One of the main challenges when assessing the harmonic distortion of electric railways is the distribution
of the harmonic generation along the railway route in both time and space. Electric railways can be very
long and as such supplied from the transmission/distribution network at many points which to some
degree have to be assessed as one system, not as individual independent connections. One way of
achieving this is to represent the traction power system by a number of equivalents which together
represent the entire railway effect on the transmission/distribution system [99].
The railway system representation must be adequate and it has to include suitable representation of all
trains operating on the system to a given schedule or timetable. Detailed representation of the trains in
harmonic studies can be very complex and it is common to employ some method of a simplified, yet
adequately accurate representation. Such simplification can utilise network equivalents of each feeder
station or traction supply point which accounts for the effects of the ac traction system impedance on
the harmonic propagation through the system and the aggregated traction load.
144
TB 766 - Network modelling for harmonic studies
Finally, the harmonic currents generated by all trains operating on the railway are aggregated using a
suitable aggregation rule. Power electronic converters typically used in trains may require different
summation law to that generally recommended [100]. A classification of the train generated harmonics
the point of view of synchronization and frequency components distribution can be grouped as follows
[101]:
Unsynchronised Emission Sources with Fixed Frequency
Synchronised Emission Sources with Fixed Frequency
Uncoupled Emission Sources with Variable Frequency
Coupled Emission Sources with Variable Frequency
Modulation Products Between Fixed and Variable Frequency Sources
145
TB 766 - Network modelling for harmonic studies
Harmonic performance limits applicable to the HV AC side of the supply point, which will generally be
part of the public network and be required to conform to applicable harmonic standards.
146
TB 766 - Network modelling for harmonic studies
Figure 4-35 V-I-characteristic of AC-Arcs (1 to 6 for Low Currents and 7 to 8 for High Currents) [102] and
Example of a Measured Harmonic Spectrum of the Primary Currents [103]
The arc resistance is non-linear. Figure 4-35 shows the measured current harmonics representative of
a large AC EAF. The result is a harmonic distortion of the currents on the primary side of the furnace
transformer and consequently a corresponding voltage distortion that depends on the network
impedance. AC furnaces are important sources of odd and even harmonics, mostly below order 11 th
(Figure 4-35), but generate also a continuous spectrum in the harmonic range.
Figure 4-36 Typical Electrical Diagram of a DC Arc Furnace with Classic 12-Pulse Operation and
Harmonic Spectrum (Measured) of the Primary Currents (Maximum and Average Values)
Figure 4-36 shows the measured current harmonics representative of a large DC EAF. All even and odd
harmonics are present. Orders 11th and 13th are dominant since the converter is in 12-pulse operation.
They are, however, significantly smaller than the theoretical calculated harmonics and they become
even smaller if the furnace operation is more unstable. This is a well-known reaction of converters under
such conditions called the “principle of the air mattress” [103].
147
TB 766 - Network modelling for harmonic studies
148
TB 766 - Network modelling for harmonic studies
When frequency domain models are not sufficient to capture the complex behaviour of the EAF, time
domain simulations with detailed models of the installation, might be needed. For the case of submerged
arc furnaces, it means a detailed representation of the converter components (thyristors) and their
control laws. These time-domain models can also be used to develop or validate simplified models for
frequency-domain analysis according to the structure shown in Figure 4-38.
The reader is warned that performing measurements of voltages and currents at the EAF facilities, as
an alternative to the model development, is probably impractical if not impossible in many cases. There
are multiple difficulties when considering direct measurements: (1) the great problem of calculating the
influence of normal background harmonics on the measurements, (2) risk of an unintended trip-out of
part or all of the EAF installation when external personnel (unfamiliar with the facilities) are on site and
cause something to go wrong, and (3) arranging for a suitable time with the furnace operator/owner to
perform the measurements (which might only be possible after annual maintenance and startup).
Harmonic Impedance
For the Thévenin (or Norton) equivalent model of the EAF (Figure 4-38), one equivalent impedance per
harmonic frequency for each stage of the process (or each relevant operating point of the furnace) is
needed. This impedance represents the passive behaviour of the furnace components (including any
power factor correction device, harmonic filters, etc.) and accounts for the influence of the furnace on
the network harmonic impedance. In addition to all passive assets, the influence of the controller (if
applicable) also has to be covered in the harmonic impedance.
When this complex behaviour cannot be captured with frequency-domain models, time domain
simulations may be necessary. Once again, the detailed time-domain models must reflect all the process
states that need to be included in the studies. For the case of the AC submerged arc furnace, the
harmonic impedance means a detailed representation of the converter components (thyristors) and their
controls is required.
149
TB 766 - Network modelling for harmonic studies
load variations affect the input AC current spectrum as well as the passive behaviour of the furnace, as
seen from the grid.
Similar to the AC EAF type, a Thévenin or Norton representation (Figure 4-38) capturing active emission
sources and passive harmonic impedance components is recommended.
Harmonic Sources
The generation of harmonic currents by DC furnaces strongly depends on the rectifier type. Assuming
ideal commutation and smoothing conditions, the harmonics are generated at orders h=n*p±1 (p is the
pulse number and n is an integer) and can be estimated using the well-known equation based on the
ideal square wave or stepped square wave input current:
𝐼
𝐼ℎ = 1. Equation 4.1
ℎ
If a rectifier without special features such as free-wheeling diodes or shift control is used, then Equation
4.1 can be used to estimate the harmonic output of a DC arc furnace. The purpose of the special features
and controls is to minimize flicker, not harmonics. In real operation, commutation and smoothing are not
ideal and the firing angles are continually changing. This leads to a spectrum of harmonics that is time-
varying, does not follow the ideal results of Equation 4.1, and cannot be calculated easily. Due to this
complexity, the model for the DC arc furnace as a source of harmonic currents must be furnished by the
specialist supplier of the power quality control equipment, such as STATCOM, SVC and any associated
filtering.
In order to assess the harmonic impact of a DC arc furnace when planning new plants or evaluating the
performance of compensating devices such as passive or active filters, time domain models may be
more suitable. For the modelling of the 12-pulse converter, the reader should refer to LCC HVDC
modelling in Section 4.3.1. However, the representation of the DC arc dynamics is quite challenging. A
deterministic approach is usually insufficient [107]. However, chaos-based methods using Chua or
Lorentz equations [108] can also be used.
Harmonic Impedances
It is a very complex problem to build a frequency domain model to represent a DC furnace or a DC
submerged arc furnace. Such models must be supplied to the utility owner or operator and, in reality,
will normally originate with the specialist contracted by the furnace manufacturer to specify the power
quality control equipment. Similar to the AC EAF case, the harmonic impedance model must capture
the passive behaviour of the furnace components (including any power factor correction device,
harmonic filters, etc.) and account for the influence of the furnace on the network harmonic impedance.
In addition to all passive assets, the influence of the controller also has to be captured in the harmonic
impedance.
Key Points
The electric arc furnace, in whichever form, is a large and often dominating source of harmonics
which deserves careful modelling for harmonic studies. It is essential that accurate models are
made available to the utility or TSO to accurately represent their effects in the power system.
It is equally important that the utility or TSO specify which furnace processes are of most interest to
be captured in the models. This implies close cooperation between the utility or the TSO and the
customer/furnace owner, especially for future new installations.
Measurements of voltages and currents at the arc furnace has severe drawbacks, not limited to the
difficulty to calculate the influence of the background harmonics.
A Thévenin or Norton representation of the EAF capturing both, active emission sources and
passive harmonic impedance components, is recommended.
150
TB 766 - Network modelling for harmonic studies
implemented in the drive. For purposes of modelling for harmonic studies, it is thus important to
recognize the differences in the spectrum and magnitudes of the harmonic currents that they generate.
Five types of drives are discussed in this section. Their technologies are briefly described below and
their harmonic modelling is discussed.
Figure 4-39 Classical 6-Pulse Diode Bridge Rectifier and 6-Pulse PWM Inverter
151
TB 766 - Network modelling for harmonic studies
The potential worst case in terms of Amps (highest magnitude of any harmonic component) is expected
to correspond to the maximum load.
In reality, the AC input inductance, due to impedance of the feeding network and supply transformers
as well as the configuration of the converter transformer winding and load current through the drive will
have a significant effect on the AC input current waveform [109]-[111]. The percent distortion currents
for an actual diode rectifier can be significantly higher than the theoretical values [109]-[112]. Figure
4-41 shows examples of AC input current spectrum for different 6-pulse VFDs at different load
conditions, with AC source inductance accounted for [113].
Figure 4-41 Examples of Input Current Spectrums for 6-Pulse VSI-PWM Drives ([114], [115], [116])
For more realistic adjustments of harmonic current magnitudes, a look-up table based on time domain
simulation results can be used [111]. Different load conditions can then be considered, as well as
different AC grid inductance or DC link inductance values. However, this approach does not consider
the rectifier operational mode (continuous or discontinuous conduction mode), on which harmonic
current emissions strongly depend. To do so, an analytical model can be used such as the one proposed
in [111].
The effects of unbalance in the supply voltage could also be considered [117], [118]. Indeed, unbalance
in the AC input voltages to the diode rectifier will result in non-characteristic triplen harmonics (3rd, 9th,
etc.) appearing in the input currents [117], [119]. The magnitudes of these harmonics will increase with
increase in degree of unbalance and there will be a corresponding decrease in the magnitudes of the
characteristic harmonics [117].
Alternative Models:
For a more accurate approach than the ideal harmonic current source, a frequency-domain harmonic
matrix model can be used. This method allows consideration of the interaction between the grid and the
VFD. An analytical model is presented in [120] and is based on an accurate determination of operational
mode. This approach, for which voltage distortion at the connection point is considered, is recommended
for harmonic penetration studies.
As a last step in the accuracy of the model, time domain simulations with circuit-based representation
can be used. With this approach, the influence of unbalance and voltage distortion is included in
calculations. However, as for any simulation of this kind, it requires much more parameters and longer
calculation time.
152
TB 766 - Network modelling for harmonic studies
constant DC bus voltage necessary for voltage source converter operation [121]-[123],
can be used to actively slow down the motor through the capability of dynamic braking by feeding
energy back into the AC system from the motor through the front end of the drive.
With some modifications, the drive with AFE has the capability of bidirectional power flow [124]
Additional information on modelling the source as well as the operational impedance of the PWM rectifier
at the relevant harmonic frequencies can be found in 4.2.1.2.1.
If passive filters are used to mitigate the effects of the sidebands of the PWM carrier frequency, they
must be included with the Norton/Thévenin equivalent circuits. Below the frequencies of the sidebands,
the drive should not produce emissions that would require filtering. PFC capacitor banks also should not
be normally required if an AFE is used.
153
TB 766 - Network modelling for harmonic studies
Figure 4-43 CSI Variable Speed Drive with PWM CSR and CSI
154
TB 766 - Network modelling for harmonic studies
commutation. Relatively large DC smoothing reactors are used in the DC link. The rectifier (6-pulse,
12-pulse or 24-pulse) controls the DC current while the inverter controls the speed of the motor [125],
[131], [132]. LCI drives are used for large motors, typically over 15 MW (e.g. for the compression of
natural gas for storage and shipping), and LCI converter blocks can even be paralleled to deliver up to
about 120 MW. If the driven motor is asynchronous, capacitors are required in the inverter circuit to
assist commutation (Figure 4-44). If the motor is synchronous (Figure 4-45), then the excitation is
adjusted to provide slightly leading power factor and the inverter phase currents are commutated using
the back EMF of the motor. Power factor at the input to an LCI drive, before application of harmonic
filters and/or power factor correction capacitor banks, can range from 0.5 to 0.92 for a variable torque
load, thereby requiring a significant amount of correction in terms of capacitive Mvar. Additionally, the
LCI has large input current harmonics, with I-THD as high as 12%, requiring substantial filtering.
Utility
Grid A
Ia
B
Ib
IM
C Ic
Induction
Motor
Harmonic
Filters and PFC
155
TB 766 - Network modelling for harmonic studies
into account. The reader is referred to 4.3.1 on LCC HVDC for some additional information on modelling
of current-source converters.
It is interesting to note that LCI drives do not have the inherent capability to mitigate any of their
harmonics, as with some PWM-based drives which use SHE. As an example, Figure 4-46 shows the
spectrum of harmonic emissions from a particular 12-pulse LCI drive for a 14,200 hp (about 13 MW)
synchronous motor for two different loading levels [132].
Figure 4-46 Input Current Harmonics for a 12-Pulse LCI Drive for a 12.47kV 14,200 hp Synchronous Motor
156
TB 766 - Network modelling for harmonic studies
where fi is the input frequency to the rectifier and f 0 is the drive output frequency. “m” is constrained by
the following requirement:
(n·p ± 1) ± m = odd integer Equation 4.5
where p = pulse number, m = integer, and n = 0, 1, 2…. Due to the complexity of the input spectrum for
a cycloconverter, it is suggested the reader use reference [125] or [133] for a more detailed treatment
of harmonic emission.
As an example, Figure 4-48 shows the cycloconverter harmonic current spectrum for an 18 MW motor
operating at full speed (11.6 RPM), taken from Table 1 of Reference [134] and based on manufacturer’s
data. It is interesting to compare Table 2 (operation at 8.5 RPM) of [134] to Table 1, showing some
shifting in the spectrum of the interharmonics. The significant feature of the emissions from
cycloconverters compared to other VFDs is the relatively significant magnitudes at some of the higher
harmonics.
157
TB 766 - Network modelling for harmonic studies
4.8 Summary
This chapter has addressed the modelling of non-linear devices as source of harmonic distortion in the
power system. The most common devices are discussed, namely converter-based generation, HVDC
converters, FACTS devices, traction systems, electric arc furnaces and variable speed drives.
For each of these devices, the chapter explains the mechanisms for generation of harmonics and
provides recommendations for adequate modelling in frequency domain for harmonic studies. When
possible, limitations of the models are highlighted and recommendations are provided for more
sophisticated methods of analysis.
A common feature of most non-linear devices is the complexity of operation and the dependency of
many factors (e.g. operating mode, control loops, converter topology, external grid, etc.) for their
harmonic performance. As such, it is not possible to develop generic models for these devices. Instead,
a common model structure is proposed in this chapter based on a Norton/Thévenin equivalent circuit
(Figure 4-1). This recommendation aligns with the recent IEC TR 61400-21-3 ED1 [62], applicable to
the modelling of Wind Turbines (WT). The same principles can be extended to model all other non-linear
devices, especially those based on power electronic converters. In addition to the obvious advantages
of standardisation, the proposed structure offers the benefit of providing a comprehensive
characterization of the non-linear device without the need to disclose proprietary design features.
158
TB 766 - Network modelling for harmonic studies
2. Assessment of compliance with the emission limits issued by the System Operator.
This study is normally performed by the Customer as a pre-connection requirement for non-linear
installations. The new facility is represented in detail and the external system is represented by the
envelopes provided by the System Operator. The objective of the study is to demonstrate compliance
with the allocated emission limits for all operating points encompassed by the envelopes. Investigation
of mitigation options is also part of the study in case of non-compliance.
4. Trend investigation.
This study is intended to investigate trends in measurements from a network of monitoring devices and
to identify areas of the grid or operating conditions with increasing levels of harmonic distortion. The
objective is to understand the drivers behind the trends and to proactively implement actions that will
prevent breaches of statutory limits.
159
TB 766 - Network modelling for harmonic studies
The extent of the network represented in detail must be sufficient to incorporate all contingencies that
need to be studied. The reader is directed to chapters 3 and 4 of this TB for guidelines on accurate
representation of individual network components.
Accurate representation of the external network with a network equivalent can be a challenging task.
Various approaches for accurate representation of network equivalents are discussed in [135]. As a rule
of thumb, the use of Power Frequency Network Equivalents (short circuit impedance or other variations)
is only adequate when the external network does not exhibit any resonances within the frequency range
of interest (see section 5.5) or when the external network is electrically so far from the area of study that
does not have significant impact on the harmonic impedance at the buses of interest. In other cases, a
frequency dependent network equivalent (FDNE) is required.
If the frequency response of the network is known (from measurements or from a complete network
model used for a frequency scan), a network equivalent can be built that reproduces its frequency
response within the frequency range of interest.
The use of the FDNE instead of the full network model saves computation time and facilitates the
exchange of data between different companies as the network topology and its parameters are hidden
[136]. It should be noted that multiple FDNE may be required in order to capture a range of scenarios
and operating conditions. This is normally implemented in the form of harmonic impedance envelopes
capturing a wide range of credible operating conditions, as described in section 5.4.
Most commercial harmonic analysis tools based on frequency domain techniques include functionality
to represent the frequency variation of external grids as look-up tables or matrices. Alternatively, if EMT
160
TB 766 - Network modelling for harmonic studies
tools are selected for the study, the frequency domain response of the external network can be fitted
with a rational function-based model which could be obtained using a technique such as vector fitting
[137] with passivity enforcement [138].
To get accurate results for the frequency response of the system, it is preferable to calculate the FDNE
from a detailed network model. However, if the detailed network model is not available in a simulation
tool, the network can be modelled in detail up to a certain distance from the point of interest and with
power frequency network equivalent, like Thévenin voltage source, beyond this distance. To determine
the minimum critical distance, CIGRE working group C4.307 [135] has suggested to “progressively
increase the distance (horizontal and vertical) of the detailed modelling until the results do not change
in a significant manner”. In other words, the detailed network is progressively extended horizontally,
starting from the highest voltage level, until the variation of the network impedance is not significant
anymore in the frequency range of interest. The network is then extended vertically to lower voltage
levels until the variation of the network impedance is not significant anymore. When extending the
network, priority should be given to the electrical nodes containing the highest capacitive components.
This method is extensively used by the French [136] and Spanish [139] transmission system operators.
Many harmonic studies have been recently conducted on the French HV grids. As a result, the detailed
network model has been progressively extended and it now includes the complete 400 kV and 225 kV
grids in an EMT tool. As an illustration, the complete 400 kV detailed network is displayed in Figure 5-1.
Other TSOs, such as those in Denmark, Ireland and the UK, have reported the use of full system models
in frequency domain for their harmonic studies.
One drawback of such large and complex network models is the continuous requirement to keep large
volumes of data up-to-date and to maintain coherency between the studies in different simulation
platforms for the same project. However, the quick accessibility to all parts of the model and increased
confidence on simulation results normally outweighs the data management overhead.
161
TB 766 - Network modelling for harmonic studies
the engineer should be cognisant of the increasing levels of uncertainty associated with the later years.
Concerning the evolving system configuration during this period, it is recommended to capture
significant development phases that could impact harmonic impedance, especially near the connection
point.
In the case of an operational study, where the installation and the system configuration are well defined,
it is important to replicate as much as possible the conditions that can be the cause of trouble or
investigation. In addition, the expected system configuration several years ahead may be considered to
investigate the risk of reoccurrence.
0,70%
Vy ph spring 2014
% Harmonic voltage of fundamental
0,60%
Vy ph summer
0,50% 2014
Vy ph winter 2015
0,40%
0,30%
0,20%
0,10%
0,00%
5
6
2
3
4
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
Harmonic order
Figure 5-2 Harmonic Voltage Distortion Measurements (95th Percentile) in a Transmission Station in the
UK for Three Levels of System Demand: Summer, Spring and Winter
162
TB 766 - Network modelling for harmonic studies
An example of harmonic impedance variation due to network contingencies is illustrated in Figure 5-3.
This graph shows frequency scans at a transmission node using a full detailed system model of the Irish
system. A single circuit outage introduces a large parallel resonance at approximately 800 Hz and a
double-circuit outage shifts this resonance towards 750 Hz.
1000
Intact (N) Condition
900
Single contingency (N-1) condition
800
600
500
400
300
200
100
0
0 200 400 600 800 1000 1200 1400 1600 1800 2000
Frequency [Hz]
Figure 5-3 Harmonic Impedance in a Transmission Station in Ireland under Intact (N), Single Contingency
(N-1) and Double Contingency (N-2) Conditions
Depending on the complexity of each network and the range of harmonic frequencies of interest, it is
not straightforward to provide general guidance on the number and type of contingencies to be studied
for each system configuration or demand level. This choice should be made through engineering
judgement, based on knowledge and understanding of each network. To reduce number of cases, it is
163
TB 766 - Network modelling for harmonic studies
generally suggested to restrict the contingencies up to the third node away from the point of interest
unless there is evidence that other contingencies beyond that boundary could result in more onerous
conditions within the range of harmonic frequencies of interest. The need to check the operational
feasibility of each contingency is emphasised.
It is therefore worthy to be noted that attention must be given to considering only realistic conditions
(practically feasible combination of system configuration / demand level / contingencies / capacitor
banks / SVCs and HVDC), in order to limit the computational burden and, more importantly, to avoid
overly-pessimistic results leading to investment in mitigation or over-design of harmonic filters.
Any network condition that can be deemed unrealistic, particularly in terms of generation and load
scenarios (i.e. those conditions that imply impossible operating scenarios or which might fail to provide
a convergent fundamental frequency load flow), should not be included in the harmonic study. If the
software does not allow for load flow calculations, it should at least be verified that the fundamental
frequency short-circuit impedance, calculated for the same cases as the harmonic impedance, is within
the anticipated range.
164
TB 766 - Network modelling for harmonic studies
The concept of harmonic impedance loci and envelopes is illustrated with an example in Figure 5-4,
which relates to the impedance at a 110 kV substation in the Irish transmission system. Figure 5-4 (a)
shows the well-known frequency scan representation in which the magnitude of impedance is plotted
against frequency for a defined network operating condition. Peaks in frequency scan relate to parallel
resonances (e.g. 635 Hz) while valleys relate to series resonances (e.g. 698 Hz). Figure 5-4 (b) shows
the same harmonic impedance data but plotted in an R-X plane. This representation reveals the path
traced in the complex plane by a single network operating state with increasing frequency, also known
as impedance locus. It can be seen that the harmonic impedance exhibits a characteristic spiral-like
path with multiple loops.
Figure 5-4 (b) shows that the harmonic impedance at this transmission node for the assumed network
condition presents inductive behaviour for most of the frequency spectrum. However, it changes to
capacitive and then back to inductive at the major resonant frequencies (highlighted in the figure), which
is reflected as a major loop in the R-X plane. Major resonances can be identified in the R-X plane as
those points where the impedance path crosses the horizontal axis (i.e. where the impedance is purely
resistive). The impedance locus also shows a number of minor loops in which the sign of the impedance
does not change. It is worth mentioning that the characteristic behaviour around major resonances is
typically defined by large changes in impedance associated with small changes in frequency, as can be
seen in Figure 5-4 (b) in the frequency range from 625 Hz to 640 Hz.
Figure 5-4 (c) displays a group of frequency scans corresponding to different scenarios and operating
conditions selected as per guidelines described in section 5.3.2. It can be seen that the network
impedance exhibits a number of series and parallel resonances whose amplitude and location can vary
significantly with different operating conditions. Figure 5-4 (d) displays the same harmonic impedance
data in the R-X plane, comprising the loci for all operating conditions. The red dashed line is the
harmonic impedance envelope. The network harmonic impedance can present any value on the
boundary or within this envelope.
The harmonic impedance data can be subdivided into smaller envelopes covering narrower frequency
ranges as shown in Figure 5-4 (e) for 5th ≤ h ≤ 7th and Figure 5-4 (f) for 26th ≤ h ≤ 35th. Figure 5-4 (g)
compares the reduced frequency range envelopes with the larger single frequency range. This
subdivision allows a more realistic power quality assessment, especially in the low order harmonic
range.
It should be noted that this section concentrates on harmonic impedance envelopes as the most
commonly used approach in industry to define system impedance for compliance with harmonic
emission limits and/or for filter design. However, some utilities may adopt other approaches, such as
the provision of families of trajectories for defined frequency ranges or the provision of frequency scans
covering the frequency range of interest for all scenarios. Independently of the format selected to present
the data, the practical considerations described in section 5.4.2 still apply.
165
TB 766 - Network modelling for harmonic studies
Z [ohm] X [ohm]
400 350
350 300
Parallel 635 Hz
300 Resonance 250
625 Hz
200
250
150 630 Hz
200
100
150 R [ohm]
50
100 0
0 50 100 150 200 250 300 350
50 -50
Series
Resonance 698 Hz -100 635 Hz
0
640 Hz
0 200 400 600 800 1000 1200 1400 1600 1800 2000
-150
Frequency [Hz]
(a) Impedance vs frequency for one operating condition (b) Harmonic impedance locus from 2nd to 40th harmonic orders
Z [ohm]
X [ohm]
1000
800
900
800 600
700
400
600
500
200
400
R [ohm]
300 0
0 100 200 300 400 500 600 700 800 900
200
-200
100
0
0 200 400 600 800 1000 1200 1400 1600 1800 2000 -400
Frequency [Hz]
(c) Impedance vs frequency for multiple operating conditions (d) Harmonic impedance loci from 2nd to 40th harmonic orders
X [ohm] X [ohm]
250 600
500
200
400
150 300
200
100
100
R [ohm]
0
50 0 100 200 300 400 500 600 700 800 900
-100
R [ohm]
0 -200
0 50 100 150 200 250
(e) Harmonic impedance loci from 5th to 7th harmonic orders (f) Harmonic impedance loci from 26th to 35th harmonic orders
X [ohm]
800
2 ≤ h ≤ 40
600
26 ≤ h ≤ 35
400
5≤h≤7
200
R [ohm]
0
0 100 200 300 400 500 600 700 800 900
-200
-400
Figure 5-4 Example of Harmonic Impedance in a Transmission Network: Locus, Loci and Envelope (5 Hz
Calculation Resolution)
166
TB 766 - Network modelling for harmonic studies
200
Zh(max)/2
100
h(max)
0 R [ohm]
0 100 200 300 400 500 600
h(min)
-100
Rh(min)
-200
-300
Sectors
This is a variation of the circle envelope that restricts the area by setting upper and lower limits to the
magnitude and angle of the harmonic impedance as shown in Figure 5-6. Each sector is defined by the
following four parameters: Zh(max), Zh(min), h(max), h(min), applicable to the frequency range captured by
the envelope. Similar to the circle representation, sectors are very quick and easy to calculate.
X [ohm]
300
250
200
Zh(max)
150
100 h(max)
50 Zh(min) h(min)
R [ohm]
0
0 50 100 150 200 250 300
Figure 5-6 Generic Sector Envelope
167
TB 766 - Network modelling for harmonic studies
Polygons
Discrete polygons with variable number of vertices can provide a close delimitation of the impedance at
specific harmonic orders without including excessive empty areas with no calculated impedance points.
A basic definition of a polygon envelope is shown in Figure 5-7, based on maximum and minimum values
of resistance, reactance and impedance angles within the frequency range captured by the envelope.
More complex and irregular polygon shapes can also be defined if needed to provide a better fit to the
calculated impedance values (see Figure 5-4 and Figure 5-8), for example by means of convex-hull.
Polygons are typically defined by the co-ordinates of the vertices, provided in tabular format. In general,
this type of envelope requires the most effort to produce, however it also provides the best definition for
the harmonic behaviour of the system impedance.
4
A frequency step of (fn/100) is reported in [140] as typically used by a European TSO.
168
TB 766 - Network modelling for harmonic studies
frequency bands, with some overlapping as discussed next. It should be noted that the
recommendations below are of general nature and mostly applicable to typical power systems
dominated by overhead lines. Systems with large share of cable circuits will typically exhibit resonances
at low harmonic orders and the engineer will need to take account of that characteristic behaviour when
defining the most appropriate frequency width and overlap of harmonic envelopes.
The choice of frequency bands is very much related to the frequency behaviour of the network. In
general, since low harmonic orders do not tend to exhibit large resonances, the impedance can be
adequately defined with individual harmonic order envelopes or with narrow frequency bands (e.g. two
to four harmonic orders); see Figure 5-4 (e). The use of envelopes covering a wide frequency range is
not advised for low harmonic orders unless resonances are observed in that frequency range as it will
typically result in unrealistic large impedance areas and, hence, too onerous conditions for assessment
(see Figure 5-4 (g)).
At higher frequencies, numerous resonances tend to appear at multiple frequencies and it is not possible
to capture this complex behaviour with narrow frequency band envelopes. In order to confidently capture
possible resonance shifts and/or modelling uncertainties it is necessary to widen the size of the
frequency bands. As far as practicable, the boundaries of the bands should avoid any resonant
frequency; see Figure 5-4 (f) and (g).
The introduction of some overlapping between subsequent envelopes is advised as a safety margin to
account for uncertainties such as small variations of system frequency, equipment tolerances,
parameter variation with temperature, modelling errors, etc. At low frequencies the uncertainties tend to
be low, therefore, widening the size of the envelope by ±½ harmonic order is normally adequate if there
are no significant resonances. For example, an envelope covering the 4th and 5th harmonic orders could
include data from the 3.5th to the 5.5th. At higher frequencies, the modelling uncertainty tends to increase
and accuracy tends to decrease, hence higher overlapping between envelopes is justified. As a general
guideline, a ±10% tolerance of the harmonic orders covered in the envelope can be used (although
larger or smaller tolerance may be required in different systems depending on their frequency response
and the modelling accuracy). This means that an envelope covering the harmonic range from 21 st to
34th should be extended to cover from 18.9th to 37.4th. As there will be an intrinsic overlap with the bands
for the higher and lower harmonic groups, the analysis of frequencies within the overlap range needs to
consider both envelopes (or a composite of the two).
The shape of the envelopes should be selected to minimise “empty” areas without realistic impedance
points. Polygons are very widely used as they can provide a very close fit to a set of calculated
impedance points. Typically, polygons can be defined by six to ten vertices, but higher number may be
required in cases. The larger the number of vertices, the best fit is achieved. The use of convex-hull
polygons provides the closest possible fit (i.e. minimum “empty” areas) at the expense of large number
of vertices, which can increase the computation burden for the assessing party.
An illustrative example comparing various envelope shapes is shown below. Figure 5-8 (a) compares
the fit of the generic envelopes described in section 5.4.1 for a set of calculated harmonic impedance
values in a defined frequency range (from 5th to 7th harmonic orders). In this example, it is clear that the
full-circle envelope produces an area that is too large and does not reflect the behaviour of the system
impedance in this frequency range (e.g. the circle suggests a large area of capacitive behaviour that is
not present in the calculated impedance data set). The size of “unrealistic” impedance area reduces
sequentially with the use of sector, basic polygon and convex-hull fit polygon.
The possibility of adapting polygons to reduce empty areas is shown in Figure 5-8 (b). In this example,
the basic polygon has been enhanced by cutting the empty top left corner, thus introducing an additional
vertex. The top-left area has been selected because it is normally one of the most critical regions for
compliance due to the low resistance values. The same approach can be adopted in all other corners if
necessary. In this example, the resulting number of vertices is 6, 7 and 27 for basic polygon, enhanced
polygon and convex-hull, respectively. The user needs to balance the requirements for accuracy (i.e.
minimum empty areas) and computational burden (i.e. number of vertices).
169
TB 766 - Network modelling for harmonic studies
X [ohm] X [ohm]
300 250
200
200
X [ohm]
100 120
150 100
R [ohm]
0 80
0 100 200 300 400 500 600
100 60
-100
40
50 20
-200 R [ohm]
0
0 50 100
R [ohm]
-300 0
0 50 100 150 200 250
(b) Harmonic impedance envelopes: Basic Polygon
(a) Harmonic impedance envelopes: Circle,
(continuous line), Enhanced Polygon (dashed line),
Sector, Basic Polygon, Convex-Hull Polygon
Convex-Hull Polygon
For the purpose of harmonic filter design and definition of resonance between de-tuned filters and the
network impedance, the critical points on the harmonic impedance envelope always lie on the left-hand
boundaries of the envelope (TB553 – appendix 2.1 [23]). Thus, most care must be taken in the definition
of this boundary – especially with regard to the maximum angle and the maximum and minimum values
of impedance.
In some instances, different configurations of the power system may create clouds of impedance points
which lie in distinctly different areas of the R-X plane (for example, in the event of a major line outage).
In such cases it may be expedient to create two envelopes, each containing a distinct area of the R-X
plane, as being the best way to minimize the inclusion of unrealistic harmonic impedances. See example
in Figure 5-9.
170
TB 766 - Network modelling for harmonic studies
Future-proofing
While it is possible to derive harmonic impedance envelopes to quite accurately represent the existing
or future planned states of the power network, treatment of developments in the more distant future is
more difficult.
Electrical power systems are highly likely to change significantly in nature over the lifetime of any new
plant being installed now. Renewable energy sources, distributed generation, electric vehicle charging,
and the introduction of very large numbers of associated power electronic converters may have a
significant and largely unpredictable impact on the network harmonic impedance.
Future changes in the network should not necessarily imply larger impedance envelopes – as networks
increase in capacity the impedances may well decrease and the harmonic impedance envelopes shrink
in some cases (note that the use of insulated cables can have the opposite effect by means of
introduction of low order harmonic resonances). This has often been the case with what are now
regarded as conventional power networks. But with new technological paradigms entering the industry,
future predications become more uncertain.
The expected life of the newly-connected plant and associated filters may be in the range of 30-40 years.
It is reasonable to expect that harmonic performance of the plant should continue to be acceptable
during much or all of that period. But if the network harmonic impedance changes substantially in the
future, such that it transgresses the impedance envelopes used for the design, then the harmonic
performance may exceed the agreed limits – through no fault of the plant owner. Any consequential
measures required for mitigation could be problematical – not only in terms of contractual/legal
discussions regarding responsibility for the cost, but also in terms of physical space for additional filtering
and the impact on the plant availability during modification works.
Consequently, there is an argument for including additional margins in the network impedance
envelopes to be used for design purposes, to create a degree of “future-proofing”. Such additional
margins, implying expansion of the areas covered by envelopes in the Z-plane, may well result in
additional filtering capacity being required and a higher degree of damping in the filter design, with
concomitant higher losses. Both imply an increased cost for the connection.
Whether or not to include margins for future-proofing in the definition of harmonic envelopes is not a
question on which there can be any general recommendation. Decisions must be made on a case-by-
case basis, depending on the contractual relationship of the parties involved and the assessed risk and
cost of any future modifications.
171
TB 766 - Network modelling for harmonic studies
model and also a linear representation derived from the short-circuit impedance. The results obtained
with both approaches deviate before the main parallel resonance (at approximately 450 Hz) and do not
converge again. The limitations of this approach are obvious and can lead to over- or under-estimation
of the actual network harmonic impedance depending on the system characteristics and frequency
range.
500
Impedance in Ohm
400
Short-Circuit Equivalent
300
200
Detailed Frequency Dependent
100 Model
0
0 500 1000 1500 2000 2500
Frequency in Hz
100
75
50
Angle in degree
Short-Circuit Equivalent
25
-50
-75
-100
Frequency in Hz
Figure 5-10 Comparison of Harmonic Impedance Derived from a Power Frequency Short-Circuit
Equivalent and from a Detailed Frequency-Dependent Model
It can be concluded that linear short-circuit equivalents are only suitable for frequencies well below the
first network resonance. In most cases, this frequency can be estimated using Equation 5.1 where f1 is
the fundamental frequency, SSC is the short-circuit power at the point of interest and Q is the reactive
power of the capacitive components directly (or close-by) connected to the point of interest. A more
detailed representation of the network is required if the scope of the harmonic analysis includes
frequencies close to (or above) f1.
𝑆𝑆𝐶
𝑓𝑟 = 𝑓1 √ Equation 5.1
𝑄
Reference [8] includes a discussion on this topic concluding that that the use of a simple inductance
based on the fault level contribution gives unsatisfactory results when applied to transmission systems.
A simple equivalent based on a “T circuit” is shown to give acceptable results reproducing the harmonic
impedance of a generic transmission system with a dominant resonance. While this approach is an
improvement from the basic simple inductance representation, caution is advised when applying it to
systems with more complex behaviour showing multiple dominant resonances (e.g. Figure 3-30, Figure
3-75, Figure 5-3 and Figure 5-10). The use of FDNE is recommended in those situations (see section
5.3.1).
Key Point:
The practice of using short-circuit impedance to derive harmonic impedance at the PoC must be
discouraged for applications in meshed power systems.
172
TB 766 - Network modelling for harmonic studies
173
TB 766 - Network modelling for harmonic studies
Figure 5-11 Impact of the Offshore Cable and Transformer Harmonic Losses Modelling on System
Resonance Damping at Offshore Substation (source [145])
Numerous publications (e.g. [69], [146], [147], [148]) have shown the dependency of harmonic distortion
to certain operational factors such MV collector network topology (e.g. radials in or out of service),
connection of WTGs, generation output, etc. Thus, it is recommended to incorporate these variables in
the studies in order to capture all practical combinations of resonance shifts and harmonic emissions
from the installation.
Once the customer’s installation has been accurately represented in the model, the external grid
information needs to be considered as well. Here, the customer will need to adapt to the amount of
information the network owner/operator can provide. However, it is worth noting that insufficient network
data will translate into a poor model and thus into a risky analysis. As an example, the study documented
in [141] demonstrates that the use of simplified grid impedance models based on short-circuit power
and X/R ratio can lead to very different amplitude and location of resonances when compared to a
detailed assessment using frequency dependent harmonic impedance envelopes. Therefore, it is
important to discuss with the relevant system owner/operator the need for accurate representation of
the external grid before investing in mitigation equipment.
Similar to the new installation, the external grid needs to be considered with all possible operation
scenarios that can have an impact on harmonic behaviour (see section 5.4 for a detailed discussion on
this topic).
It is demonstrated in TB553 (Appendix 2.1) [23] that the critical resonances between the external grid
and the customer’s installation always lie in the perimeter of the external grid harmonic impedance
envelope. Thus, it is generally acceptable to analyse interactions for Grid Code compliance considering
only the points in the perimeter of the grid impedance envelope. It should be noted that all points within
the perimeter need to be assessed, not only the vertices.
One part of a harmonic study is the evaluation of amplification of background harmonics caused by
interaction between the new installation and the external grid. As this is often more critical than harmonic
emission, it is important to consider realistic background distortion at the connection point. In cases
where representative measurements are not available (for example, the substation has not been built
yet), the system operator needs to estimate these values. Depending on the network topology and/or
available monitoring devices in the area, this may be difficult, and it is a topic requiring further research.
This is discussed in detail in section 5.8.
Harmonic studies are often performed to capture worst case operating conditions. As long as harmonic
limits are not exceeded this approach causes minimal workload and gives confidence in operation with
regards to harmonics. However, if limits are exceeded further considerations are needed to ascertain
whether mitigation measures are required. To justify this decision, a more detailed study on harmonics
would be beneficial. It may be the case that harmonic limits will only be exceeded if the new installation
is in an operating point which is rarely used. To perform detailed studies considering different operating
points, a full EMT model or frequency domain harmonic models considering different operating points
(e.g. with power steps of 10 %) are needed. This will normally result in a very large number of scenarios
to consider, but it becomes manageable when only applied to the areas of “non-compliance”. While
174
TB 766 - Network modelling for harmonic studies
performing a detailed harmonic study, it is advisable to use accurate data for the phase angle of single
harmonics instead of general summation laws.
As it may be concluded from the paragraphs above, harmonic studies developed by a customer for a
new installation carry a certain amount of risk, mainly due to some data being quite preliminary. As a
result, it is easy to overestimate the emitted distortion and even to oversize mitigation solutions, such
as filters, that may even not be required later. While preliminary studies give a good indication of
potential harmonic issues, in some cases it is advisable to consider the level of uncertainties and to
discuss more detailed analysis or even a conditional connection with the system operator (as described
in Stage-3 of IEC 61000-3-6 [100]) to verify harmonic performance before committing to design a
particular mitigation solution.
𝛼
𝛼
𝑈ℎ = √∑ 𝑈ℎ𝑖 Equation 5.2
𝑖
Where
𝑈ℎ is the voltage magnitude at harmonic order ℎ
𝑈ℎ𝑖 is the magnitude of the various emission levels at order ℎ to be combined
𝛼 is an exponent as defined in Table 5-1
Harmonic Order 𝜶
ℎ<5 1
5 ≤ ℎ ≤ 10 1.4
ℎ > 10 2
Key Point:
The IEC 61000-3-6 general summation law may be applicable in many cases but is only
recommended for use when no better information is available. In addition, the alpha exponents
included in [100] were recommended based on information available at the time of publication,
mostly through analysis of rectifier bridges, and it is suggested that they may not be suitable
nowadays to capture the behaviour of modern harmonic sources [149].
The usefulness and practicality of the IEC 61000-3-6 general summation law is presented in [150], as
no phase angle information is required. Other contributions, such as [143] and [146], show that the use
of this summation law may yield very misleading results, and the application of the alpha exponent to
two harmonic sources with differing amounts of randomness, has been shown to be inconsistent with
measurement data. Furthermore, CIGRE TB 672 [79] concludes that if larger PV plants are built using
multiple individual PV inverters of the same model, the standard summation exponents (e.g. according
175
TB 766 - Network modelling for harmonic studies
to IEC 61000-3-6) are not suitable and the harmonic currents of individual PV inverters should be added
up arithmetically independent of the harmonic order.
Results of extensive synchronised harmonic measurements performed in an off-shore wind farm (type
4 WTG) in Denmark are reported in [151]. It is shown that the use of the IEC general summation law
leads to incorrect results. In this particular case, harmonic current emissions between the 5 th and the
20th order are under-estimated at the PCC whereas harmonic currents between the 2nd and 4th and
above the 45th are over-estimated. A mismatch factor of 4 was observed for high frequency harmonic
currents. This work concludes that the use of the default exponents in the IEC general summation law
can lead to incorrect assessment of harmonic emissions with respect to Grid Code compliance and
recommends further work to better understand the phase angle behaviour of the power converter
harmonic emissions.
A recent contribution to the literature, [149], proposes that the general IEC 61000-3-6 summation law
be revised in order to adequately consider modern harmonic sources. This proposal suggests the use
of a uniform distribution of the difference in phase angles to calculate the probability that the magnitude
of the summation vector of two harmonics may exceed a given value. New alpha exponents are
recommended based on this work [149].
Another contribution [152] presents various methods for the summation and propagation of random
harmonic phasors, deriving analytical explicit expressions of currents and voltages for simple cases
where the injected currents exhibit uniform or Gaussian distributions and are independent. For more
complex cases, other techniques such as Monte Carlo simulation are required.
In summary, the application of the IEC 61000-3-6 general summation law is reiterated in [150] and [152],
with the advice that other exponents be used in specific cases where more detailed information is
available. A recommendation for a revision of the alpha exponents is included in [149]. For now, it is
recommended to apply caution and/or consult with converter manufacturers on adequate exponents to
be applied for the aggregation of harmonic emissions from multiple non-linear devices.
176
TB 766 - Network modelling for harmonic studies
Figure 5-12 Comparison of Two Different Weeks of 7th Harmonic Measurement in Germany
Table 5-2: Comparison of Different Statistic Values of Measurements of 7th Harmonic of Two Different
Weeks in Germany
Figure 5-13 shows the measurement of the 7th harmonic at the same location as the example above for
two different months. It illustrates that completely different results can be obtained. This can have
financial and performance implications, e.g. if design is based on measurements of the red curve
problems during operation may occur as equipment will be exposed to higher harmonic distortion than
specified or, alternatively, additional amplification caused by capacitive components may result in
breaches of statutory limits. Table 5-3 summarises the statistical analysis of the measured data. It clearly
indicates that the measured values in month 1 are much higher than in month 2. In extreme, the 99th
percentile of month 1 is 86 % higher than the equivalent for month 2 and thus could lead to inadequate
conclusions.
One observation can be made from this example: the statistical differences between the measured data
reduce as the measurement period increases.
177
TB 766 - Network modelling for harmonic studies
Figure 5-13 Comparison of Two Different Months of 7th Harmonic Measurement in Germany
Table 5-3: Comparison of Different Statistical Values of Measurements of 7th Harmonic for Two Different
Months in Germany
The same trend is seen in long-term measurements. As an example, Figure 5-14 shows the phase-to-
ground 95th percentile values of the 11th and 13th harmonics in a 400 kV Danish substation. The
measuring period is 20 weeks and the 95th percentile levels are determined per week based on 10 min
aggregated values.
178
TB 766 - Network modelling for harmonic studies
Figure 5-14 Phase-to-Ground 95th Percentile Values of the 11th and 13th Harmonics in a 400 kV Danish
Substation over 20 Weeks
The maximum and minimum weekly 95th percentile levels of harmonics are presented in Table 5-4.
Table 5-4 Maximum and Minimum 95th Percentile Levels of 11th and 13th Harmonics in a 400 kV Danish
Substation over 20 Weeks
The ratios of the maximum and minimum weekly 95th percentile levels measured over the 20 weeks are
displayed in Table 5-5.
Table 5-5 Ratios of the Maximum and Minimum 95th Percentile Levels of 11th and 13th Harmonics in a 400
kV Danish Substation over 20 Weeks
179
TB 766 - Network modelling for harmonic studies
The variations in the levels of harmonics and therefore also the ratios between the maximum and
minimum 95th percentile values are highest for the 11th harmonic. The highest ratio over the 20 weeks
occur in phase C and is 2.25 and only 1.43 in phase B for the same harmonic. The 13th harmonic levels
are more constant over the 20 weeks with ratios changing from 1.37 to 1.45.
The reason for the large variation in level of harmonic is the dependency on operation conditions of the
power system, including changes in topology in the main grid (for example circuit outages, sectionalising
stations, etc.) and operating points of e.g. HVDC, wind power plants, arc furnace and production and
consumption schemes at distribution level. It cannot be guaranteed that the grid is operating under
representative or worst-case situations during the measuring period and it can be impossible to bring
the system to a known worst-case situation due to restriction in operating the grid during measurements,
for instance forced contingency situations. Therefore, knowledge of the system’s behaviour for instance
derived from use of simulation models combined with measurements is often required to ensure robust
designs.
Key Point:
The levels of harmonics can vary significantly either over short time or according to season. For
some harmonics, the levels can be quite constant is relation to time and season where others
measured at the same location can vary significantly. Therefore, recommendation for a
“representative measuring period” is system dependent, but in general it can be stated that
measurements should be conducted for as long as possible, ideally for not less than three months,
including measurements of all three phases.
Key Point:
Harmonic 95th and 100th percentile levels can be significantly different and the ratio can vary
significantly over time. Depending on the type of study e.g. system planning or component rating
this should be considered.
180
TB 766 - Network modelling for harmonic studies
Figure 5-15 Sorted 10 min. Aggregated Values Measured over One Week for the 11th, 13th and 23rd
Harmonics in a Danish 400 kV Substation
Figure 5-16 Ratios between the 100th and 95th Percentile Levels Evaluated for the 11th and 13th Order
Harmonics Measured over 20 Weeks in a Danish 400 kV Substation
181
TB 766 - Network modelling for harmonic studies
182
TB 766 - Network modelling for harmonic studies
Event #1 Event #2
NODE #1
NODE #2
NODE #3
NODE #4
NODE #5
Figure 5-17 Harmonic Voltage Distortion Measurements in five Substations in Ireland over a 7-
Day Period. Legend: 7th harmonic, 11th harmonic, 13th harmonic
A second example showing harmonic measurements at a wind farm located in the southern end of Brazil
is presented next. Figure 5-18 and Figure 5-19 show the 5th and 7th harmonic voltage measured over a
period of five months at the PCC for each of the three phases. The output power of the wind farm is also
shown for completeness. A sharp increase in distortion at both harmonic orders occurred on February
22nd, coinciding with the commissioning of a new 525 kV transmission line in the vicinity of this wind
farm. For the 7th harmonic the effect is more pronounced and it is probably caused by a system
impedance change and particularly by the “electrical proximity” of the non-linear loads from a
metropolitan and industrialized area to the PCC.
Other event impacting the harmonic distortion level at the PCC of the same wind farm is shown in Figure
5-20, covering a 9 hour period. The measurements in this figure show an abrupt reduction of the 5th
183
TB 766 - Network modelling for harmonic studies
harmonic voltage at approximately 1:35 a.m., when two 525 kV transmission lines were disconnected
for voltage control. At 8:01 a.m., when the first of the two transmission lines was energized, a sharp
increase of voltage distortion at the 5th harmonic order was observed. The most interesting fact about
this event is that the two disconnected transmission circuits are about 1,000 km away from the wind
farm.
Figure 5-18 5th Harmonic: Voltage vs Power Figure 5-19 7th Harmonic: Voltage vs Power
The two examples above illustrate the impact of switching operations over large areas of the power
system and highlights the difficulties in selecting a measurement period unaffected by these variations.
184
TB 766 - Network modelling for harmonic studies
Moreover, the problem becomes even more complex if other customer facilities have been recently
given emission limits but have not connected to the grid yet, or ramped to their contracted capacity,
hence their actual emissions are uncertain. In these cases, it is common practice to assume that the
new customers will eventually use their entire allocation of harmonic headroom. Therefore, the
“measured” background needs to be modified to capture the effect of the new contracted connections,
reducing the available headroom for future customers or network developments.
Accurate estimation of pre-connection background distortion is important for two reasons: (i) over-
estimation of background will result in very restrictive emission limits for the connectee, which may drive
unnecessary cost for mitigation (for example installation of harmonic filters), (ii) under-estimation of
background may result in over-relaxed emission limits to the new connectee that can cause breaches
in Planning Levels and damage to equipment, requiring a mitigation solution to be implemented by the
System Owner/Operator and possibly compensation for damages. Either way, there are costs
implications that are normally carried to the end customer.
The problem of prediction of background distortion can be divided into two parts:
1. Estimation of background in existing nodes where measurements are not available.
2. Prediction of harmonic distortion in future nodes or modification in existing nodes subject to significant
topology changes.
185
TB 766 - Network modelling for harmonic studies
State Estimator (ETAD). An overview of the system is shown in Figure 5-21. The State Estimator module
(ETAD) is described in detail in Appendix F.
Measurements Unknowns
Estimation
Figure 5-21 Power Quality Monitoring and Analysis System by Red Eléctrica de España (REE) [165]
From the literature review and TSO’s experience, it can be concluded that there are feasible solution
methods to tackle the problem of harmonic estimation in nodes without measurements. However, these
methods have not been widely adopted by System Owners/Operators, possibly due to the effort required
in implementing and integrating them with their standard harmonic analysis tools and models. Availability
of HSE functionality in commercial harmonic analysis packages would be a valuable step forward for
the accurate assessment of harmonic distortion in power systems and to assist in fair allocation of
emission limits to new customer connections.
186
TB 766 - Network modelling for harmonic studies
of the method and the fact that results err on the side of safety makes this approach attractive for most
harmonic studies. However, careful consideration is recommended to any series resonance points in
the frequency scan which can lead to the computation of unrealistically large gain values.
A common method used to estimate the modification of background at a given node caused by the
connection of customer’s equipment is the “Constant Voltage behind Thévenin Equivalent” approach
[23], [168]. In this approach, the background harmonic distortion is assumed constant for any network
condition while the harmonic impedance of the system, seen at the PoC, is normally given as a range
of values (see section 5.4) to capture different operating conditions (including circuit outages). This
approach is illustrated in Figure 5-22 below, which represents the connection of an offshore wind farm.
In this example, V_back represents the pre-existing background distortion at the PoC (measured or
estimated), Z_Net represents the harmonic impedance of the transmission grid at the PoC (normally
given as a family of harmonic impedance envelopes for a range of frequencies) and Z_WF represents
the harmonic impedance of the customer facility and comprises all cables, transformers, reactive
compensation devices, etc at the customer facility beyond the PoC. Z_WF can also vary depending on
the operating conditions; for example the number of collector arrays and WTG in service.
Figure 5-22 Thévenin Equivalent Representation for Assessing Modification of Background Distortion
due to a Customer Connection [168]
The modification of the pre-existing harmonic voltage distortion at the PoC caused by the passive
components of the customer’s installation is given by Equation 5-3, where “h” denotes each harmonic
order and the impedances are given as complex numbers. The modification factor (kh) can, theoretically,
take any value depending on the relative values of the system and customer impedances. As such, the
effect of the customer installation on the distortion at the PoC could be: (i) amplification of pre-existing
background if kh > 1, (ii) reduction of pre-existing background if k h < 1 or (iii) no change in pre-existing
background if kh = 1.
𝑉ℎ−𝑏𝑎𝑐𝑘𝑔𝑟𝑜𝑢𝑛𝑑 (𝑝𝑜𝑠𝑡−𝑐𝑜𝑛𝑛𝑒𝑐𝑡𝑖𝑜𝑛) 𝑍ℎ_𝑊𝐹
𝑘ℎ = = Equation 5-3
𝑉ℎ−𝑏𝑎𝑐𝑘𝑔𝑟𝑜𝑢𝑛𝑑 (𝑝𝑟𝑒−𝑐𝑜𝑛𝑛𝑒𝑐𝑡𝑖𝑜𝑛) 𝑍ℎ_𝑊𝐹 + 𝑍ℎ_𝑁𝐸𝑇
The difficulties typically encountered with the “Constant Voltage behind Thévenin Equivalent” approach
relate to the fact that the network conditions leading to the highest amplification factors do not
necessarily coincide with the network conditions for which the background distortion was measured or
estimated. As a result, excessive (and sometimes unrealistic) “post-connection” background distortion
can be calculated leading to investment in mitigation solutions. Caution is advised in these situations to
avoid unnecessary costs. A possible approach, when feasible, can be to estimate the pre-connection
background distortion for each network condition and, therefore, provide a family of Thévenin
Equivalents with linked values of network harmonic impedance and associated background distortion.
This approach, while theoretically correct, may not be feasible in many cases due to the large volume
of data to be processed, high uncertainty in calculation assumptions or the need to reserve headroom
for future growth. A balance must be sought between practicality in the exchange of information between
customer and System Owner/Operator and the justification for costs in mitigation. In some cases, it
might be beneficial to re-examine critical conditions arising from worst-case assumptions.
187
TB 766 - Network modelling for harmonic studies
5.9 Summary
Due to the proliferation of PE device connections in power systems, harmonic studies are gaining
importance as part of the planning and design process for new customer connections. In setting up the
simulation models, it is important to focus on the purpose of the analysis as well as to understand the
limitations of the methods, tools and models. The overall objective of the assessment is to define the
appropriate requirements for the non-linear connection, taking into account foreseeable conditions in
the power system as well as present and future uncertainties over the life time of the installation while
avoiding, insofar as possible, over-conservative requirements that may lead to investment in
questionable mitigation measures or over-design of harmonic filters. This chapter provides a general
discussion and practical guidelines on key aspects that bear consideration during harmonic impact
assessments of new customer connections. It is emphasised that no two power systems are the same
and regulatory and contractual arrangements may differ from country to country, therefore caution and
good engineering judgement is advised in implementation of the above.
An overview of the most common types of harmonic studies is provided, from a network owner/operator
perspective as welll as from a new customer or equipment manufacturer point of view. Key inputs and
outcomes of each type of analysis are highlighted.
Important considerations regarding the representation of the power system and setting up simulation
cases include the extent of the model and the number and type of scenarios to be considered.
General guidelines on the minimum number of nodes to be accurately represented as well as how to
derive accurate equivalents of the rest of the network are provided.
A discussion on the scenarios and contingencies that can be considered in the harmonic studies,
depending on the type and objective, is included in this chapter. It has been shown that the maximum
harmonic distortion does not always correlate with peak or valley system demand levels. Therefore, it is
imporant to incorporate intermediate demand levels in the studies in order to capture the changes in
system resonances and emissions from all non-linear devices. Outages of network components that
can affect the harmonic behaviour of the system at the PoC need to be captured. These can include
most circuits in the vicinity of the PoC as well as harmonic filters, capacitor banks, FACTS devices and
HVDC installations. Detailed guidelines are provided.
The need to capture all possible combinations of key variables than can affect the harmonic performance
of the system normally leads to unmanageable number of cases to run and analyse. Thus, automation
is normally adopted to perform the studies. Caution should be exercised to ensure that all scenarios
included in the assessment are realistic and comply with the utilty operational standards.
The concept of harmonic impedance loci and envelopes is explained and illustrated with a real example.
The most common types of envelopes used in industry are discussed, highlighting their strengths and
limitations. Practical considerations for creating harmonic impedance envelopes are provided.
It has been shown that the use of network equivalents derived from short-circuit impedance calculations
at power frequency are not adequate to reproduce the complex frequency dependent behaviour of
typical trasmission systems. Its use is strongly discouraged.
Some considerations for the representation of the customer installation when performing design and
compliance assessment studies are discussed. The need for detailed models of the power electronic
converters is emphasised while the use of “ideal harmonic current source” approach is shown to be
overly pessimistic. How to aggregate the harmonic emissions from multiple sources within the new
installation is a critical assumption that can have high cost implications in terms of non-compliance.
Detailed discussions with vendors are recommended to gain a better understanding of the prevailing
phase angle and expected cancelations. Furthermore, aggregation of multiple turbines, transformers
and MV cables into a single equivalent can be beneficial in some large installations to reduce
computational burden.
It is concluded that harmonic studies developed by a customer for a new installation carry a certain
amount of risk. As a result, it can be easy to overestimate the prospective emissions and even to
oversize mitigation solutions, such as filters, that may even not be required later on. While preliminary
studies give a good indication of potential harmonic issues, in some cases it is advisable to consider the
level of uncertainties and to discuss further detailed studies or special conditions for the connection with
the System Owner/Operator such as conditional connections.
A review of literature suggests that the commonly used IEC general summation law [100] may not be
adequate to capture the phase angle behaviour of emissions in modern power converters. Caution is
188
TB 766 - Network modelling for harmonic studies
advised as well as consultation with converter manufacturers on adequate exponents to be applied for
the aggregation of harmonic emissions from multiple non-linear devices.
The topic of duration of background measurements is controversial. IEC documents recommend, as a
minimum, seven days of continuous operation. In principle, the objective is to measure for a long enough
period to derive a representative signature of the background distortion at that location. Since harmonic
distortion is continuously changing, depending on a large number of variables, pre-determining the
minimum duration requirements is a challenging (if not impossible) task. Therefore, recommendation for
a “representative measuring period” is system dependent, but in general it can be stated that measuring
should be conducted for as long as possible, ideally for not less than three months, including
measurements of all three phases.
As it is often the case, some nodes cannot be monitored (or have not been built yet), therefore
background estimation is needed. Various estimation methods are discussed, highlighting their
applicability and limitations.
During the discussions leading to the compilation of this chapter, the following areas where future work
is required were identified:
Summation of harmonic sources: achieve a better understanding of the phase angle behaviour of
modern converters (deterministic vs stochastic), effects of converter control, modulation scheme,
operation point, background distortion, etc. The overall objective should be to gather enough
information, based on field measurements and theoretical analysis, to propose a robust and realistic
alternative to improve the IEC summation law currently adopted in most harmonics analysis.
Develop practical methods for accurate aggregation of wind farm components (wind turbine
generators, transformers, cables, etc) into a single frequency dependent equivalent.
Develop practical methods for the estimation of background distortion in meshed network topologies
where availability of measurements is limited.
189
TB 766 - Network modelling for harmonic studies
190
TB 766 - Network modelling for harmonic studies
6. Conclusions
Power systems globally are experiencing a transition towards decarbonisation of electricity production
through large-scale deployment of renewable energy sources (RES), which are gradually displacing
conventional thermal plant. This changing environment is seeing a proliferation of power electronic
converters connecting at all voltage levels in power systems, namely RES, FACTS devices, HVDC
systems, domestic load, etc. These devices are highly non-linear and emit harmonics at the point of
connection, but also modify pre-existing harmonics in the network. In addition, increased installation of
HVAC cables is creating system resonances at frequencies close to the characteristic emissions from
these non-linear devices. As a result, many power systems are experiencing an increase in harmonic
distortion. Power quality issues associated with harmonics in power systems are becoming more
pronounced and are driving a new focus towards the need to undertake detailed analysis at the planning
stages in order to ensure adherence to statutory limits.
This Technical Brochure has been compiled drawing expertise from a worldwide membership base and
provides comprehensive guidelines for practising power system engineers when they need to perform
harmonic distortion assessments. The document covers the modelling of the most common network
components and discusses key features that need to be considered in the assessments. The focus of
this Technical Brochure is on frequency-domain modelling for steady-state AC harmonic analysis in
power systems, typically in the range from power frequency up to the 50th harmonic (2.5 kHz in 50 Hz
systems or 3 kHz in 60 Hz systems), consistent with typical power quality assessments. The approach
and modelling guidelines provided are reasonably valid up to the 100 th harmonic if specialized studies
are required. The document is not intended to cover guidelines for analysing transient issues or
controller harmonic instability, although most of the guidelines for modelling network components
equally apply to those areas also.
The main focus of this Technical Brochure is on practical aspects of modelling for direct application in
the planning process of connecting a new customer (non-linear installation) to the power system, or
when introducing a change to the system as part of asset replacement or system expansion. These
guidelines will be valuable in the definition of harmonic performance specifications for new HVDC
converters, FACTS devices or other non-linear installations. They will also assist connectees when
modelling their installation to assess or demonstrate compliance with the emission limits provided by
the System Owner/Operator and to investigate and specify mitigation measures such as harmonic filters.
Furthermore, this Technical Brochure can also be used post-commissioning for any incident
investigation or to assist resolution of customer complaints via modelling and analysis.
By reviewing existing literature on modelling techniques and approaches for each network component,
complemented with comprehensive analysis and expertise from the members, this Technical Brochure
reflects the current state-of-the-art and best practice in network modelling for harmonic studies.
An overview of available solution methods for analysis of harmonics in power systems, including
frequency-domain, time-domain, harmonic-domain and hybrid methods is given. A comprehensive
analysis of the strengths and limitations of each method is provided. It is concluded that frequency-
domain solution methods provide adequate results for most applications, along with straight-forward
modelling of harmonic injections and frequency dependence, making this study domain accessible and
attractive to many engineers. Some cases where frequency-domain approaches are not suitable are
identified and recommendations are provided for the most appropriate alternative solution method.
Issues associated with asymmetry in network components are discussed and illustrated with an
example. Recommendations for three-phase modelling and application of unbalanced solution methods
are included.
A comprehensive review of models available to represent the most common passive network
components in harmonics studies is given. The relevant input data and level of detail required to
represent each component are presented and discussed. Different modelling options are assessed in a
benchmark model as well as in real system models to illustrate the effects and consequences of each
type of model in the context of system-wide studies. When available, some models have been validated
against measurement data and recommendations are provided. The importance of representing the
frequency dependency of resistive losses is stressed and illustrated with examples. Furthermore, the
importance and difficulties of accurate load modelling are highlighted. The general load composition and
its harmonic behaviour is normally one of the biggest unknowns when performing harmonic
assessments in power systems. Various approaches for load aggregation are presented and discussed,
however this is an area where further work is required to improve and validate load models considering
191
TB 766 - Network modelling for harmonic studies
the on-going changes in load composition, i.e. the move towards more power electronics based, as well
as seasonal and daily variations.
A chapter is devoted to the accurate representation of harmonic generating equipment in the frequency
domain. The most common sources of harmonic distortion are presented, discussing the mechanisms
of harmonic generation for each device, aspects that influence their harmonic performance and
elements that need to be considered in the model. When possible, limitations of the frequency-domain
models are highlighted, and recommendations are provided for more sophisticated methods of analysis.
It is emphasized that the harmonic performance of these devices is generally complex and dependent
on many factors such as converter topology, control strategy, operating point, external grid, etc. As such,
manufacturer-specific models need to be adopted. Recommendations on the structure of the models
and the features that need to be captured are provided. Emphasis is placed on the need to capture not
only the harmonic current/voltage emissions but also the harmonic impedance of the devices. A common
model structure is proposed based on a Norton/Thévenin equivalent circuit. This approach aligns with
the recent IEC TR 61400-21-3 ED1 [62], applicable to the modelling of wind turbines. The same
principles can be extended to model all other non-linear devices, especially those based on power
electronic converters. In addition to the obvious advantages of standardisation, the proposed structure
offers the benefit of providing a comprehensive characterization of the non-linear device without the
need to disclose proprietary design features.
A final chapter is devoted to examining practical aspects, other than the modelling of each individual
component, that need to be considered when performing harmonic studies related to the connection of
non-linear devices to the power grid. In setting up the simulation models, it is important to focus on the
purpose of the analysis as well as to understand the limitations of the methods, tools and models. An
overview of the most common types of harmonic studies is provided, from a network owner/operator
perspective as well as from a new customer or equipment manufacturer point of view. Key inputs and
outcomes of each type of analysis are highlighted. For example, the overall objective of a typical system
owner/operator study is the definition of appropriate emission limit requirements for a new non-linear
connection, considering foreseeable conditions in the power system as well as present and future
uncertainties over the lifetime of the installation. The focus should be on avoiding, insofar as possible,
over-conservative requirements that may lead to investment in questionable mitigation measures or
over-design of harmonic filters while ensuring adequate system performance in all transmission and
distribution nodes, in adherence with statutory limits.
Recommendations on the extent of the model, scenarios and contingencies that can be considered in
the harmonic studies, depending on the type and objective, are provided. It has been shown that the
maximum harmonic distortion does not always correlate with maximum or minimum system demand
levels. Therefore, it is imporant to incorporate intermediate demand levels in the studies in order to
capture the changes in system resonances and emissions from all non-linear devices. The concept of
harmonic impedance loci and envelopes is explained and illustrated with a real example. The most
common types of envelopes used in industry are discussed, highlighting their strengths and limitations.
Practical considerations for creating harmonic impedance envelopes are provided. The use of network
equivalents derived from short-circuit impedance calculations at power frequency is shown to be not
adequate for meshed power systems and, therefore, this practice is strongly discouraged. Finally, the
background distortion at the point of connection is a key input in most harmonic studies.
Recommendations for an adequate period of measurements and data analysis are given. As is often
the case, some nodes cannot be monitored, therefore background estimation is needed. Various
estimation methods are discussed, highlighting their applicability and limitations.
The discussions and recommendations in this Technical Brochure address perspectives as seen from
(i) the system owner/operator; and (ii) the new connectee, with the overall objective of minimising risk
of equipment failure due to excessive harmonic distortion as well as avoiding unnecessary investment
in mitigation. While finding the right balance is never an easy task, it is hoped that the considerations
presented in this document will aid the engineer in making informed decisions.
Future work on the subject of harmonic modelling and analysis may expand the effort already performed
in the following directions:
Load modelling: Characterisation and development of accurate aggregation and representation
methods. Special focus should be placed on capturing the transition in the general load
composition towards a more predominant power-electronic based type. Interactions between the
PE-based load impedance and the system harmonic impedance as well as characterisation of
harmonic emissions need to be accounted for in the enhanced load models. Validation through
lab tests and field measurements is recommended.
192
TB 766 - Network modelling for harmonic studies
Validation of power converter models with field measurements. Accurate methods to separate the
effect of system harmonic impedance and background distortion from the converter emissions.
Summation of harmonic sources: achieve a better understanding of the phase angle behaviour
of modern converters (deterministic vs. stochastic), effects of converter control, modulation
scheme, operation point, background distortion, etc. The overall objective should be to gather
enough information, based on field measurements and theoretical analysis, to propose a robust
and realistic alternative to improve the IEC summation law currently adopted in most harmonics
analysis.
Develop practical methods for accurate aggregation of wind or solar farm components (wind
turbine generators, PV converters, transformers, cables, etc.) into a single frequency-dependent
equivalent.
Develop practical methods for the estimation of background distortion in meshed network
topologies where availability of measurements is limited.
193
TB 766 - Network modelling for harmonic studies
194
TB 766 - Network modelling for harmonic studies
7. Bibliography/references
[1] CIGRE JTF 36.05.02/14.03.03, “AC System Modelling For AC Filter Design – An Overview of Impedance
Modelling”, Electra 164, pp 133-151, February 1996.
[2] CIGRE WG CC-02 (CIGRE 36.05/CIRED 2), “Guide for assessing the network harmonic impedance”,
Electra No. 167, pp 97-131, July 1996.
[3] Arrillaga, J.; and Watson, N. R.; “Power System Harmonic Analysis“, 2 nd Edition, John Wiley & Sons, Ltd,
2003.
[4] IEEE Task Force on Harmonics Modelling and Simulation; Medina, A.; Segundo-Ramirez, J.; Ribeiro, P.;
Xu, W.; Lian, K.L.; Chang, G.W.; Dinavahi, V.; Watson, N.R., “Harmonic analysis in frequency and time
domain“, IEEE Transactions on Power Delivery, 28, (4), 2013.
[5] CIGRE TB 754 “AC Side Harmonics and Appropriate Harmonic Limits for VSC HVDC”, WG B4.67,
February 2019.
[6] Herraiz, S.; Sainz, L.; Clua, J., “Review of harmonic load flow formulations”, IEEE Transactions on Power
Delivery, vol. 18, pp. 1079-1087, July 2003.
[7] Testa, A.; Akram, M.F.; Burch, R.; Carpinelli, G.; Chang, G.; Dinavahi, V.; Hatziadoniu, C.; Grady, W.M.;
Gunther, E.; Halpin, M.; Lehn, P.; Liu, Y.; Langella, R.; Lowenstein, M.; Medina, A.; Ortmeyer, T.; Ranade,
S.; Ribeiro, P.; Watson, N.; Wikston, J.; Xu, W., "Interharmonics: Theory and Modelling," IEEE
Transactions on Power Delivery, vol.22, no.4, pp.2335-2348, Oct. 2007.
[8] IEEE Power Engineering Society Task Force on Harmonics Modelling and Simulation, “Tutorial on
Harmonics Modelling and Simulation“, TP-125-0, 1998.
[9] Dinh, N.; and Arrillaga, J.; “A salient-pole generator model for harmonic analysis “, IEEE Transactions on
Power Systems, 2001, 16, (4), pp. 609-615.
[10] Rittiger, J.; Kulicke, B., "Calculation of HVDC-converter harmonics in frequency domain with regard to
asymmetries and comparison with time domain simulations," IEEE Transactions on Power Delivery, vol.10,
no.4, pp.1944-1949, Oct 1995.
[11] Froebel, A.; Vick, R. "Chosen aspects for harmonic analysis in distribution networks", 22 nd International
Conference on Electricity Distribution, CIRED 13, June 2013.
[12] Harmonic Analysis in Frequency and Time Domain, IEEE Task Force, IEEE Power Delivery, July 2013
[13] J. Arrillaga and C. D. Callaghan, ''Three Phase AC-DC Load and Harmonic Flows,” IEEE Trans. on Power
Delivery, Vol. 6, No.1, January 1991, pp. 238-244.
[14] J.G. Mayordomo, L.F. Beites, R. Asensi, M. Izzeddine, A.H. Bayo, “INTAR, A Powerful Tool to Simulate
Harmonics under Balanced and Unbalanced Conditions”. International Conference on Electricity
Distribution. CIRED, 1997.
[15] IEEE Standard 1124, IEEE Guide for analysis and definition of DC side harmonic performance of HVDC
transmission systems
[16] Fuchs, E.F.; “Power Quality in Power Systems and Electrical Machines”, Elsevier Academic Press, 2008.
[17] C. F. Jensen; Ł. H. Koeewiak; Z. Emin; “Amplification of Harmonic Background Distortion in Wind Power
Plants with Long High Voltage Connections”, CIGRE Paris Session 2016
[18] C. F. Jensen; “Harmonic Background Amplification in Long Asymmetrical High Voltage Cable Systems”,
IPST 2017
[19] Lombard, X.; Mahseredjian, J.; Lefebvre, S.; Kieny, C., "Implementation of a new harmonic initialization
method in the EMTP," IEEE Transactions on Power Delivery, vol.10, no.3, pp.1343-1352, July 1995.
[20] Semlyen, A.; Medina, A.; "Computation of the periodic steady state in systems with nonlinear components
using a hybrid time and frequency domain methodology," IEEE Transactions on Power Systems, vol.10,
no.3, pp.1498-1504, July 1995.
[21] J. de Jesus Chavez, J.; Ramirez, A.I.; Dinavahi, V.; Iravani, R.; Martinez, J.A.; Jatskevich, J.; Chang, G.W.,
"Interfacing Techniques for Time-Domain and Frequency-Domain Simulation Methods," IEEE
Transactions on Power Delivery, vol.25, no.3, pp.1796-1807, July 2010.
[22] Arrillaga, J.; Medina, A.; Lisboa, M.L.V.; Cavia, M.A.; Sanchez, P., “The harmonic domain. A frame of
reference for power system harmonic analysis“, IEEE Transactions on Power Systems, 1995, 10, (1), pp.
433-440.
[23] CIGRE TB 553 “Special Aspects of AC Filter Design for HVDC Systems”, October 2013.
[24] J. Arrillaga, B.C. Smith, N.R. Watson, A. R. Wood, “Power System Harmonic Analysis”. John Wiley & Sons,
195
TB 766 - Network modelling for harmonic studies
1997.
[25] E. Acha, M. Madrigal, Power Systems Harmonics – Computer Modelling and Analysis. John Wiley & Sons,
2001.
[26] CIGRE WG 36-05. “Harmonics, characteristic parameters, methods of study, estimates of existing values
in the network”. Electra 77, July 1981.
[27] IEEE Task Force on Harmonics Modelling and Simulation. Modelling and Simulation of the Propagation of
Harmonics in Electric Power Networks. Part I: Concepts, Models and Simulation Techniques. and Part II:
Sample Systems and Examples IEEE Transactions on Power Delivery, Vol 11, No 1, January 1996.
[28] W. T. Wiechowski, Harmonics in transmission power systems. Aalborg: Institut for Energiteknik, Aalborg
Universitet, 2006.
[29] J. Arrillaga, E. Acha, T.J. Densem, P.S. Bodger. Ineffectiveness of Transmission Line Transpositions at
Harmonic Frequencies. Proc IEE, 123C(2). 1986.
[30] Wenner, F., “A Method of Measuring Earth Resistivity,” Report No. 258, Bulletin of Bureau of Standards,
Vol. 12, No. 3, October 11, 1915
[31] Palmer, L. S., “Examples of geotechnical surveys,” Proceedings of the IEE, Paper 2791-M, vol. 106, pp.
231–244, June 1959
[32] Deri, A., Tervan, A. Deri, G. Tevan, A. Semlyen and A. Castanheira, "The Complex Ground Return Plane.
A Simplified Model for Homogeneous and Multi-Layer Earth Return," IEEE Power Engineering Review,
vol. PER-1, no. 8, pp. 31-32, Aug. 1981.
[33] H. W. Dommel, EMTP Theory Book, Prepared for Bonneville Power Administration, Portland, Oregon,
1995
[34] CIGRE TB 531 “Cable systems electrical characteristics”, 2013.
[35] CIGRE TB 556, “Power system technical performance issues related to the application of long HVAC
cables”, 2013.
[36] Ł. H. Kocewiak, B. Gustavsen, “Impact of Cable Impedance Modelling Assumptions on Harmonic Losses
in Offshore Wind Power Plants”, CIGRE Paris Session, August 2018. Paper C4-309.
[37] F.F. Da Silva, C.L. Bak, “Electromagnetic Transients in Power Cables”, Springer-Verlag London, 2013
[38] A. Ametani, “A general formulation of impedance and admittance of cables”, IEEE Transactions on Power
Apparatus and Systems, vol. PAS-99, no.3, 1980.
[39] Ł. H. Kocewiak, “Harmonics in large offshore wind farms,” PhD Thesis, 2012, pp. 332, 978-87-92846-04-
4.
[40] M.H.J. Bollen, S.M. Gargari, “Harmonic resonances due to transmission cables”, in CIGRE Belgium
Conference, Brussels, 2014.
[41] A. Pagnetti, “Cable modelling for electromagnetic transients in power systems”, PhD thesis, Université
Blaise Pascal – Clermont II, Clermont-Ferrand, France, 2012.
[42] A. Ametani, T. Ohno, N. Nagaoka, “Cable System Transients: Theory, Modelling and Simulation”, Wiley
IEEE Press, 2015.
[43] U. R. Patel and B. Gustavsen and P. Triverio, “An Equivalent Surface Current Approach for the
Computation of the Series Impedance of Power Cables with Inclusion of Skin and Proximity Effects”, IEEE
Transactions on Power Delivery, 2013, volume 28, number 4, pp 2474-2482.
[44] C. H. Chien and R. W. G. Bucknall, “Harmonic Calculations of Proximity Effect on Impedance
Characteristics in Subsea Power Transmission Cables”, IEEE Transactions on Power Delivery, 2009,
volume 24, number 4, pp 2150-2158.
[45] U. R. Patel and B. Gustavsen and P. Triverio, “Proximity-Aware Calculation of Cable Series Impedance
for Systems of Solid and Hollow Conductors”, IEEE Transactions on Power Delivery, 2014, volume 29,
number 5, pp 2101-2109.
[46] CIGRE WG B1.03, TB 272, “Large cross-sections and composite screens design”, Electra, June 2005.
[47] Ł. H. Kocewiak, B. Gustavsen and Andrzej Hołdyk, "Wind Power Plant Transmission System Modelling for
Harmonic Propagation and Small-signal Stability Analysis", International Workshop on Large-Scale
Integration of Wind Power into Power Systems, and Transmission Networks for Offshore Wind Farms,
Berlin, 2017.
[48] Benato and S. D. Sessa, "A New Multiconductor Cell Three-Dimension Matrix-Based Analysis Applied to
a Three-Core Armoured Cable," in IEEE Transactions on Power Delivery, vol. 33, no. 4, pp. 1636-1646,
Aug. 2018.
196
TB 766 - Network modelling for harmonic studies
[49] IEEE Std. 399, IEEE Recommended Practice for Industrial and Commercial Power Systems Analysis,
1997.
[50] G. Funk, T. Hantel: Frequenzabhängigkeit der Betriebsmittel von Drehstromnetzen, etz Archiv Band 9 Heft
11, 1987.
[51] Ivan Arana Aristi, “Switching overvoltages in offshore wind power grids. Measurements, modelling and
validation in time and frequency domain”. PhD Thesis, Technical University of Denmark. November 2011.
[52] CIGRE WG 13-05, “The Calculation of Switching Surges. Part II: Network Representation for Energisation
and Re-Energisation Studies on Lines Fed by an Inductive Source”. Electra No 32, pp 17-42, 1974.
[53] Bjørn Gustavsen, “A Filtering Approach for Merging Transformer High-Frequency Models With 50/60-Hz
Low-Frequency Models”. IEEE Transactions on Power Delivery, Volume: 30, Issue: 3, June 2015, PP1420-
1428.
[54] CIGRE TB 566 “Modelling and Aggregation of Loads in Flexible Power Networks”, WG C4.605, February,
2014.
[55] IEEE Task Force on Harmonic Modelling and Simulation, “Impact of Aggregate Linear Load Modelling on
Harmonic analysis: A comparison of Common Practice and Analytical Methods”, IEEE Transactions on
Power Delivery, Vol 18, No. 2, April 2003
[56] Capasso et al “Representation of large asynchronous Load Areas for Harmonic Penetration studies”,
International Conference on Harmonics in Power Systems, 1992.
[57] R. Lamedica, A. Prudenzi, E. Tironi, D. Zaninelli, “A Model of Large Load Areas for Harmonic Studies in
Distribution Networks”, IEEE Transaction on Power Delivery, Vol. 12, Nº 1, January 1997.
[58] Zhang, X.P.; and Handschin, E.; “Frequency-dependent simple harmonic model of synchronous
machines“, IEEE Power Engineering Review, May 2000, pp. 58-60.
[59] Powertech Labs 1994 report 267D766
[60] Hydro-Québec, Direction MAINTENANCE DES ÉQUIPEMENTS, ET SÉCURITÉ DES BARRABES,
SERVICE ESSAIS ET ÉTUDES TECHNIQUES, DIVISION ÉTUDES TECHNIQUES - TRANSPORT,
Omer Bourgault, et al, September 1995.
[61] J.B. Ward “Equivalent Circuits for Power Flow Studies”, AIEE Winter General Meeting, 1949.
[62] IEC TR 61400-21-3 ED1: Wind energy generation systems – Part 21-3: Wind turbine harmonic model and
its application.
[63] CIGRE TB 719 “Power Quality and EMC issues with future electricity networks”. JWG C4.24/CIRED,
March 2018.
[64] Snyder, M. "Development of Simplified Models of Doubly-Fed Induction Generators (DFIG): A contribution
towards standardized models for voltage and transient stability analysis", Master Thesis, Chalmers
University of Technology, 2012.
[65] Muller, S.; Deicke, M.; De Doncker, R.W., "Doubly fed induction generator systems for wind
turbines," IEEE Industry Applications Magazine, vol. 8, no. 3, pp.26-33, May/Jun 2002.
[66] Larose, C.; Gagnon, R.; Prud’Homme, P.; Fecteau, M.; Asmine, M.; “Type-III Wind Power Plant Harmonic
Emissions - Field measurements and aggregation guidelines for adequate representation of harmonics”,
Proceedings 11th International Workshop on Large-Scale Integration of Wind Power into Power Systems
as well as on Transmission Networks for Offshore Wind Power Plants (Wind Integration Workshop),
Lisbon, 2012.
[67] Bradt, M.; Badrzadeh, B.; Camm, E.; Mueller, D.; Schoene, J.; Siebert, T.; Smith, T.; Starke, M.; Walling,
R., "Harmonics and resonance issues in wind power plants," Transmission and Distribution Conference
and Exposition (T&D), 2012 IEEE PES, vol., no., pp.1-8, 2012.
[68] Lui, S.Y.; Pimenta, C.M.; Pereira, H.A.; Mendes, V.F.; Mendonca, G.A., "Aggregated DIFG wind farm
harmonic propagation analysis", Anais do XIX Congresso Brasileiro de Automática, CBA 2012.
[69] King, R.; Ekanayake, J.B., "Harmonic modelling of offshore wind farms," IEEE Power and Energy Society
General Meeting, July 2010.
[70] Hernandez, E.; Madrigal, M., "A Step Forward in the Modelling of the Doubly-fed Induction Machine for
Harmonic Analysis" IEEE Transactions on Energy Conversion, vol.29, no.1, pp.149-157, March 2014.
[71] E. Guest, T. Rasmussen, K. H. Jensen, “An Impedance-Based Active Filter for Harmonic Damping by
Type-IV Wind Turbines”. 17th Wind Integration Workshop. 17 - 19 October 2018. Stockholm, Sweden.
Submission-ID WIW18-20.
[72] M. Lehmann, M. Pieschel, M. Juamperez, K. Kabel, Ł. H. Kocewiak, S. Sahukari, “Active Filtering with
Large-Scale STATCOM for the Integration of Offshore Wind Power”. 17 th Wind Integration Workshop. 17
197
TB 766 - Network modelling for harmonic studies
198
TB 766 - Network modelling for harmonic studies
199
TB 766 - Network modelling for harmonic studies
Under Input Voltage Unbalance and Sag Conditions:, IEEE Transactions on Power Delivery, Vol. 21, No.
2, pp. 567-576.
[120] Y. Sun, C. Dai, J. Li, J. Yong, “Frequency-domain harmonic matrix model for three-phase diode-bridge
rectifier”, IET Gener. Transm. Distrib., 2016, Vol. 10, Iss. 7, pp. 1605–1614
[121] T. Alexander, S.M. Lingham, T.S. Davies, P. Iwanciw, “Experience With 3-Phase Sinusoidal Regenerative
Front-Ends”, Fifth International Conference on Power Electronics and Variable Speed Drives, 1994, pp.
246-250.
[122] J. Kikuchi, T.A. Lipo, “Three-Phase PWM Boost-Buck Rectifiers With Power-Regenerating Capability”,
IEEE Transactions on Industry Applications, Vol. 38, No. 5, September/October 2002, pp. 1361-1369.
[123] L. Moran, J. Espinoza, M. Ortiz, J. Rodriguez, J. Dixon, “Practical Problems Associated With the Operation
of ASDs Based on Active Front End Converters in Power Distribution Systems”, Record of the IEEE
Industry Applications Conference, 39th Annual Meeting, 2004, Vol. 4, pp. 2568-2572.
[124] J.R. Rodriguez, J.W. Dixon, J.R. Espinosa, J. Pontt, “PWM Regenerative Rectifiers: State of the Art”, IEEE
Transactions on Industrial Electronics, Vol. 52, No. 1, Feb. 2005, pp. 5-22.
[125] B.K. Bose, “Modern Power Electronics and AC Drives”, 2002, The University of Tennessee, Knoxville,
ISBN 0-13-016743-6.
[126] L. Li, D. Czarkowski, Y. Liu, P. Pillay, “Multilevel Selective Harmonic Elimination PWM Technique in Series-
Connected Voltage Inverters”, IEEE Transactions on Industry Applications, Vol, 36, No. 1,
January/February 2000.
[127] M. Chomat, L. Schreier, J. Bendl, “Control of Active Front-End Rectifier in Electric Drive Under Unbalanced
Voltage Supply in Transient States”, PRZEGLᾼD ELECTROTECHNICZNY (Electrical Review), ISSN
0033-2097, R. 88, No. 1a, 2012
[128] X.H. Wu, S.K. Panda, J.X. Xu, “Design of Plug-In Repetitive Control Scheme for Eliminating Supply-Side
Current Harmonics of Three-Phase PWM Boost Rectifiers Under Generalized Supply Voltage Conditions”,
IEEE Transactions on Power Electronics, 2010, Vol. 25, Issue 7, pp. 1800-1810
[129] A.V. Stankovic, K. Chen, “A New Control Method for Input-Output Harmonic Elimination of the PWM Boost-
Type Rectifier Under Extreme Unbalanced Operating Conditions”, IEEE Transactions
[130] A.V. Stankovic, T.A. Lipo, “A Novel Control Method for Input Output Harmonic Elimination of the PWM
Boost Type Rectifier Under Unbalanced Operation Conditions”, IEEE APEC 2000, Feb. 6-10, pp. 413-418.
[131] B. Wu, J. Pontt, J. Rodriguez, S. Bernet, S. Kouro, “Current-Source Converter and Cycloconverter
Topologies for Industrial Medium-Voltage Drives”, IEEE Transactions on Industrial Electronics, Vol. 55,
No. 7, July 2008, pp. 2786-2797
[132] K.A. Puskarich, W.E. Reid, P. Hamer, “Harmonic Experiences With a Large Load Commutated Inverter
Drive”, IEEE Paper No. PCIC-99-18
[133] B. Ozpineci, L.M. Tolbert, “Cycloconverters”, An on-line tutorial for the IEEE Power Electronics Society,
2001
[134] A. Symonds and M. Laylabadi, “Cycloconverter Drives in Mining Applications”, IEEE Industry Applications
Magazine, November/December 2015, pp 36-46.
[135] CIGRE TB 568. “Transformer Energization in Power Systems: A Study Guide”, WG C4.307, February
2014.
[136] Y. Vernay, B. Gustavsen, “Application of Frequency-Dependent Network Equivalents for EMTP Simulation
of Transformer Inrush Current in Large Networks”. 2013 International Conference on Power Systems
Transients (IPST2013) in Vancouver, Canada July 18-20, 2013.
[137] B. Gustavsen and A. Semlyen, “Rational approximation of frequency domain responses by vector fitting”,
IEEE Trans. Power Delivery, vol. 14, no. 3, pp. 1052-1061, July 1999.
[138] B. Gustavsen and A. Semlyen, “Enforcing passivity for admittance matrices approximated by rational
functions”, IEEE Trans. Power Systems, vol. 16, pp. 97-104, Feb. 2001.
[139] G. Alvarez-Cordero, A. Bachiller Soler, A. Gomez-Exposito, J. A. Rosendo Macias, C. Gomez-Simon, “A
Methodology for Harmonic Impedance in Large Power Systems. Application to the Filters of a VSC”.
CIGRE Paris Session 2012. Paper C4-112.
[140] R. de Groot, F. van Erp, K. Jansen, J. van Waes, M. Hap, L. Thielman. “Method for Harmonic and TOV
Connection Impact Assessment of Offshore Wind Power Plants – Part I: Harmonic Distortion”. 17th Wind
Integration Workshop. 17 - 19 October 2018. Stockholm, Sweden. Submission-ID WIW18-124.
[141] V. Myagkov, L. Petersen, S. Burutxaga Laza, F. Iov, L. H. Kocewiak, “Parametric Variation for Detailed
Model of External Grid in Offshore Wind Farms”. Proceedings of the 13 th International Workshop on Large-
scale Integration of Wind Power Into Power Systems As Well As on Transmission Networks for Offshore
200
TB 766 - Network modelling for harmonic studies
201
TB 766 - Network modelling for harmonic studies
Quality (ICREPQ’14) Renewable Energy and Power Quality Journal (RE\&PQJ), 2014.
[163] Z.-P. Du, J. Arrillaga and N. Watson, "Continuous harmonic state estimation of power systems," IEE
Proceedings- Generation, Transmission and Distribution, vol. 143, no. 4, pp. 329-336, July 1996.
[164] K. Yukihira, “Development of a Program for Estimating Harmonic Current Sources”, CRIEPI Report
T89043, May 1999.
[165] A. Diaz García, L. F. Beites, M. Alvarez and L. Soto Cano, “Power Quality Monitoring and Assessment in
the Spanish Transmission System” CIGRÉ Paris, 2016. C4-108.
[166] Y. Fillion, S. Deschanvres, “Background harmonic amplifications within offshore wind farm connection
projects”, presented at the International Conference on Power Systems Transients (IPST2015) in Cavtat,
Croatia June 15-18, 2015.
[167] D.H. Mills, Z. Emin, D. O’Brien, M. Val Escudero and C. F. Jensen, “Calculation Method Selection for
Harmonic Voltage Distortion Gains”. CIGRE Symposium in Aalborg (Denmark) 4th – 7th June 2019. Paper
65.
[168] A.J. Hernandez M., S. Wijesinghe, A. Shafiu, “Hamonic Amplification of the 576 MW Gwynt-y-Môr Offshore
Wind Power Plant”, presented at the 11th International Workshop on Large-Scale Integration of Wind
Power into Power Systems, Lisbon, Portugal, 2012.
202
TB 766 - Network modelling for harmonic studies
where 𝑍0 (ℎ), 𝑍1 (ℎ), and 𝑍2 (ℎ) are the zero, positive and negative sequence impedances at the h’th
harmonic.
The decoupled sequence representation shown in Equation A.1 is the preferred choice for the
representation of electrical components in larger networks for many utilities.
No component in the transmission network is, however, perfectly balanced. Depending on the design of
the component, the self- and mutual impedances between phases can be significantly different. For
such a system, a full phase representation is needed and will have the following form:
Only in the case where the three self- and six mutual impedances of Equation A.2 are equal to each
other, the off-diagonal elements in the sequence impedance matrix will be zero and a decoupled
sequence representation can be used for modelling with no loss of accuracy.
The off-diagonal terms in Equation A.2 are related to the coupling between sequences (inter-sequence
coupling). For the decoupled sequence representation, the inter-sequence coupling is either neglected
by choice or naturally zero because of a balanced system. Therefore, an applied positive sequence
voltage can result in the flow of positive sequence current only. Conversely, by using a phase
representation (or a coupled sequence representation), the inter-sequence coupling is taken into
consideration and therefore an applied positive sequence voltage can result in the flow of both negative
and zero sequence current. It can be concluded that when the component under study is no longer
balanced, a decoupled sequence model is no longer sufficient to correctly represent the true electrical
behaviour of the component. The effect of inter-sequence coupling is especially pronounced at and
around harmonic resonance.
Only the positive, negative and zero sequence impedances are needed to construct a model in the
decoupled sequence domain. This is very beneficial for systems where specific component data is
unavailable. To construct a true phase-domain transmission system model, a geometrical description of
all transmission lines and their electrical properties must be known, as well as the phase impedances
of transformers, loads and others equipment. Depending on the desired level of detail, specific
information pertaining to individual line section lengths, cable system bonding, OHL conductor
transposition and substation grounding are some of the parameters that may be required. Such
requirements for data availability and the high level of detail can make this type of study very time
consuming which in some cases causes the engineer to choose decoupled sequence modelling despite
its apparent limitations. The system response is very sensitive to component parameters near
resonance. Therefore, the mutual- and hence inter-sequence coupling is highly sensitive as well. Using
phase-domain models, the influence of specific component parameters can be examined which is a
benefit as some parameters are more prone to change than others. For instance, specific design
parameters vary within bands and it is only after the component has been constructed that specific data
can be delivered. In the case of planning studies assumptions must therefore be made.
203
TB 766 - Network modelling for harmonic studies
Further to the example presented in Section 2.2.2.6, the sensitivity of the per-phase amplification on the
single-core 220 kV cable system is examined by varying the relative permittivity of the main insulation
and the system length by ± 5 % (expected variation under design). The results are shown for the three
phases in Figure A.1 (a) and (b).
App Figure A.1 Phase Amplification as Function of the Harmonic Order and (a) the Relative Permittivity
and (b) the Cable System Length
The results show that near resonance even a minor change in some of the many component parameters
can have a large influence on the harmonic amplification. The shunt capacitance of the cable is linear
dependent on the relative permittivity of the main insulation material. Therefore, this parameter has a
strong influence on the resonance frequencies of the cable system. The cable system length determines
the resonance frequencies due to the influence on all electrical cable parameters. It is therefore
recommended to undertake a sensitivity analysis if one wants to ensure that resonance points are not
missed if only a single fixed parameter value is used.
The long, unloaded radial is subjected to strong inter-sequence coupling due to low damping and
resonance behaviour in the frequency range of interest. This phenomenon is also found in a loaded
meshed grid. In Figure A.2 the harmonic voltages obtained at two Danish 400 kV buses is represented
by their phase- and sequence voltage quantities after injection of positive sequence harmonic current
of equal magnitude at all harmonics from the 2 nd to the 20th harmonic. The system is at low load (45%)
but under normal operating conditions (n-0).
204
TB 766 - Network modelling for harmonic studies
App Figure A.2 Phase and Sequence Voltage in an Inter-Sequence Coupled Meshed Transmission
Network
The frequency scans obtained at Substation 1 and 2 are presented in Figure A.3. The frequency scans
indicate unbalanced conditions at the harmonics where Figure A.2 shows strong inter-sequence
coupling. This is the case around the 16th harmonic at Substation 1 and at the 18th harmonic at
Substation 2. The results show that the highest phase voltage at the 16th harmonic (measured at
Substation 1) is 41% higher compared to the positive sequence voltage, while the difference is 84% at
the 18th harmonic for Substation 2.
App Figure A.3 Frequency Scans in Inter-Sequence Coupled Meshed Transmission Network at Two
Unique Substations
205
TB 766 - Network modelling for harmonic studies
206
TB 766 - Network modelling for harmonic studies
207
TB 766 - Network modelling for harmonic studies
Circuit Geometry
Phase A co-ordinates (-8.05, 16.4) m
Phase B co-ordinates (0, 16.4) m
Phase C co-ordinates (8.05, 16.4) m
Phase Transposition Yes (assumed perfect symmetry)
Soil Resistivity
Soil Resistivity 400 Ω.m
220/110 kV Transformers
The basic parameters of the 220/110kV transformers shown in Figure B.1 are included in Table B.3
below.
App Table B.3 220/110kV Transformers Parameters
208
TB 766 - Network modelling for harmonic studies
2 3
2 3
App Figure C.1 Sequential Steps followed in the 220 kV Network Model Development
Figure C.2 shows the calculated impedance, as seen from Bus 5, for each step of model development.
It can be seen that the frequency response of the system becomes more complex as more circuits are
added to the model. In particular, the number of resonances increases with the number of circuits, as
209
TB 766 - Network modelling for harmonic studies
there are more components interacting with each other. These interactions lead to addition or
cancellation of resonances depending on the characteristics of the circuits and their lengths. The most
important observation from this figure is that the lumped parameter model (blue traces) fails to reproduce
the complex frequency behaviour of this basic system. In the best of the cases illustrated, the lumped
model approximates the first and/or second resonant peaks, but the high-frequency responses are
always missed. This observation supports the recommendation to always use a distributed parameter
model.
210
TB 766 - Network modelling for harmonic studies
16000 80
14000 60
20
10000
Harmonic Order
0
8000
0 5 10 15 20 25 30 35 40
-20
6000
-40
4000
-60
2000
-80
0
0 5 10 15 20 25 30 35 40 -100
1 OHL - LUMPED MODEL 1 OHL - DISTRIBUTED MODEL Harmonic Order 1 OHL - LUMPED MODEL 1 OHL - DISTRIBUTED MODEL
9000 80
8000 60
Sending End Voltage [V]
7000 40
80
6000
60
Sending End Voltage [V]
5000
40
2000 -40
-60
1000
-80
0
0 5 10 15 20 25 30 35 40 -100
3 OHL - LUMPED MODEL 3 OHL - DISTRIBUTED MODEL Harmonic Order 3 OHL - LUMPED MODEL 3 OHL - DISTRIBUTED MODEL
4000 80
3500 60
40
Phase Angle [deg]
3000
Sending End Voltage [V]
20
2500 Harmonic Order
0
2000 0 5 10 15 20 25 30 35 40
-20
1500
-40
1000
-60
500
-80
0
0 5 10 15 20 25 30 35 40 -100
4 OHL - LUMPED MODEL 4 OHL - DISTRIBUTED MODEL Harmonic Order 4 OHL - LUMPED MODEL 4 OHL - DISTRIBUTED MODEL
80
3500
60
Sending End Voltage [V]
3000
40
Phase Angle [deg]
2500
20
Harmonic Order
2000 0
0 5 10 15 20 25 30 35 40
1500 -20
-40
1000
-60
500
-80
0
-100
0 5 10 15 20 25 30 35 40
5 OHL - LUMPED MODEL 5 OHL - DISTRIBUTED MODEL Harmonic Order 5 OHL - LUMPED MODEL 5 OHL - DISTRIBUTED MODEL
80
3500
60
Sending End Voltage [V]
3000
40
Phase Angle [deg]
2500
20
2000
0
Harmonic Order
0 5 10 15 20 25 30 35 40
1500 -20
1000 -40
-60
500
-80
0
0 5 10 15 20 25 30 35 40 -100
6 OHL - LUMPED MODEL 6 OHL - DISTRIBUTED MODEL Harmonic Order 6 OHL - LUMPED MODEL 6 OHL - DISTRIBUTED MODEL
80
2500
60
Sending End Voltage [V]
2000 40
Phase Angle [deg]
20
1500
Harmonic Order
0
0 5 10 15 20 25 30 35 40
-20
1000
-40
500
-60
-80
0
0 5 10 15 20 25 30 35 40 -100
7 OHL - LUMPED MODEL 7 OHL - DISTRIBUTED MODEL Harmonic Order 7 OHL - LUMPED MODEL 7 OHL - DISTRIBUTED MODEL
App Figure C.2 Harmonic Impedance at Bus-5 under Incremental 220kV Network Model Development
211
TB 766 - Network modelling for harmonic studies
The harmonic impedance has been calculated as shown in Figure C.3. Frequency scan results for bus
5 are included in Figure C.4.
13 14
100 km
80 km
110 kV 60 km 85 km
12
11 10
70 km 50 km
35 km 45 km
6 9
I = 1 (f) Amp
220 kV (pos. sequence)
1
5 V5-SC(f) 4
250 km
100 km
120 km 175 km
180 km 250 km
200 km
2 3
App Figure C.3 Setup for Calculation of Harmonic Impedance at Bus 5 including 110 kV System
212
TB 766 - Network modelling for harmonic studies
1800
1600
Voltage at Bus #5 [V]
1400
1200
1000
800
600
400
200
0
0 5 10 15 20 25 30 35 Harmonic Order 40
All 110kV circuits with lumped model One 110kV bus away distributed Two 110kV bus away distributed All 110kV circuits as distributed model
100
80
60
40
Phase Angle [ deg]
20
0
0 5 10 15 20 25 30 35 40
-20
-40
-60
-80
-100
Harmonic Order
All 110kV circuits with lumped model One 100kV b awasy as distributed Two 110kV bus away as distributed All 110kV circuits as distributed
App Figure C.4 Harmonic Impedance at Bus-5 with 220kV and 110kV Network Models
213
TB 766 - Network modelling for harmonic studies
214
TB 766 - Network modelling for harmonic studies
PE outer sheath r3
r2
Aluminium sheath b
To model the given cable, some commercial software requires that corrections be made, as described
below.
Conductor resistivity: In some software it is not possible to model stranded conductors hence a solid
copper conductor was used with corrected resistivity value (taken from IEC 60228) to account for the
spaces between the stranded/segmented wires of the conductor.
215
TB 766 - Network modelling for harmonic studies
The cable had a length of 100km and was cross-bonded as shown in Figure D.3. It comprised 15 major
sections, each of which was divided into three minor sections. The sheaths were cross-bonded between
minor sections (cross-bonding joints) and grounded at the end of the major sections (earthing joints).
The minor sections should have a similar length to keep the system as balanced as possible.
Cable Impedance
Frequency scans were performed on the base case cable configuration with the ends open-circuited or
short-circuited. The results obtained were the magnitudes and phases of the sequence impedances. In
Figure D.4 and Figure D.5 both magnitude and phase of the positive and zero sequence impedances
are shown for both configurations.
It is evident that a duality exists between open-circuited and short-circuited cable terminations where
the series resonance peaks in the open-circuit case become parallel resonance peaks in the short-circuit
case and vice versa. For the base case, the first resonance peak occurs at 255Hz, very close to the fifth
harmonic, and for frequencies up to 2500Hz, twelve resonant points can be observed (six parallel and
six series resonances).
For the remainder of the sensitivity analysis the short-circuited termination was used.
216
TB 766 - Network modelling for harmonic studies
217
TB 766 - Network modelling for harmonic studies
Harmonic Amplification
218
TB 766 - Network modelling for harmonic studies
Cable Length
The harmonic impedance for five different cable lengths (100km (base case), 30km, 50km, 70km and
150km) can be seen in Figure D.9 and Figure D.10. As the cable length increases both the frequency
and the magnitude of the resonances decrease and the number of resonances within the frequency
range increases. Similar behaviour is observed in both the positive- and zero-sequence impedance.
Key point:
Cable length plays a significant role in the frequency, magnitude and number of harmonic resonance
points in both the positive- and zero-sequence harmonic impedance.
App Figure D.9 Positive-Sequence Harmonic Impedance Comparison for Different Cable Lengths
219
TB 766 - Network modelling for harmonic studies
App Figure D.10 Zero-Sequence Harmonic Impedance Comparison for Different Cable Lengths
Conductor Radius
The harmonic impedance comparison for variations in the conductor radius is presented in Figure D.11.
The thicknesses of all other layers remain fixed (i.e. representing an outer radius change). As the
conductor radius increases so does the cable capacitance, resulting in a downward shift (decrease) in
resonant frequencies. This variation decreases the inductance of the cable and hence the overall
frequency shift is small. The positive-sequence impedance magnitude is unchanged. Zero-sequence
impedance resonant frequencies are hardly affected; however the zero sequence impedance
magnitudes are reduced as slightly more damping is introduced with increasing conductor radius. This
is shown in Figure D.12.
Key point:
Conductor radius has a minor influence on the frequencies at which resonances occur in the
positive-sequence and has a minor impact on the zero-sequence impedance magnitude.
220
TB 766 - Network modelling for harmonic studies
App Figure D.11 Positive-Sequence Harmonic Impedance Comparison for Variations in the Cable
Conductor Radius
App Figure D.12 Zero-Sequence Harmonic Impedance Comparison for Variations in the Cable Conductor
Radius
221
TB 766 - Network modelling for harmonic studies
Insulation Thickness
The harmonic impedance comparison for deviations in the insulation thickness is presented in Figure
D.13 and Figure D14. For the positive sequence impedance, as the insulation thickness decreases, the
cable capacitance increases resulting in resonance peaks shifting to lower frequencies. In this example,
the magnitude slightly decreases for smaller insulation thickness. For the zero-sequence impedance,
only the resonance magnitudes are affected. When the insulation thickness increases, so does the
impedance magnitude.
Key point:
Insulation thickness affects resonances seen in the positive-sequence harmonic impedance in terms
of frequency and magnitude, and in the zero-sequence only the magnitude is affected.
App Figure D.13 Positive-Sequence Harmonic Impedance Comparison for Variations in the Cable
Insulation Thickness
222
TB 766 - Network modelling for harmonic studies
App Figure D.14 Zero-Sequence Harmonic Impedance Comparison for Variations in the Cable Insulation
Thickness
Cable Layout
Cable installation and layout refer to where and how individual cables are positioned with respect to one
another. Example layouts include trefoil, flat and triangle as shown in Figure D.15. In the planning stage
of a cable connection the exact layout is unknown, making it important to study how different choices of
cable layout can affect the harmonic impedance.
The harmonic impedance comparison for four cable layouts (trefoil, flat, triangle and touching trefoil) is
illustrated in Figure D.16 and Figure D.17. Variations in the cable layout greatly affect both the frequency
and the magnitude of the positive-sequence impedance resonance peaks. More specifically, the flat and
triangle formations present a greater number of resonance peaks compared with the base case where
the trefoil formation was used. This is due to the higher asymmetry present in the first two layout
configurations compared to the trefoil formation. For the touching trefoil formation, a shift of the
resonance peaks to higher frequencies is observed. When the distance between the cables decreases,
the mutual inductance increases, and the total positive-sequence inductance decreases. The zero-
sequence impedance is identical in all four cases.
Key point:
Cable layout significantly affects the frequency, magnitude and number of resonances seen in the
positive-sequence harmonic impedance.
223
TB 766 - Network modelling for harmonic studies
App Figure D.16 Positive-Sequence Harmonic Impedance Comparison for Different Cable Layouts
App Figure D.17 Zero-Sequence Harmonic Impedance Comparison for Different Cable Layouts
224
TB 766 - Network modelling for harmonic studies
Sheath Bonding
The bonding configuration as well as the number of major sections used can greatly influence the
frequency response of a cable connection. Generally, for long (several km and more) cables, to prevent
large currents from circulating in the sheaths giving rise to high losses, whenever the two ends of the
sheaths are grounded, cross-bonding must be used.
Type of Bonding
The base case described in Figure D.3 where the cross-bonding configuration with 15 major sections
was used, is compared with a solidly-bonded 100km cable. The different bonding configurations will not
result in changes to the shunt admittance matrix (as no alteration is made to the distance between the
conductor and the sheath) but the series impedance matrix may differ [37].
The harmonic impedance comparison for these two bonding configurations is shown in Figure D.18 and
Figure D.19. The positive sequence impedance is significantly affected by the type of cable bonding
used with the cross-bonded cable yielding higher magnitudes and lower resonant frequencies than for
the solidly-bonded cable. In the case of solid bonding there is a 177 Hz frequency shift to higher
frequencies at the first parallel resonance and the magnitude is approximately ten times lower.
The type of bonding has only a negligible effect on the zero-sequence impedance magnitudes as shown
in Figure D.19. The zero-sequence impedance is largely independent of the type of bonding used, as
approximately no current should flow to ground at the grounding points.
Generally, for large cable connections (several km) the only viable option (aside from transposition of
phases) is sheath cross-bonding due to the large circulating sheath currents and associated higher
losses when solid bonding is used. However, the result of this comparison is valuable for the simulation
of cable systems where it is evident that the use of a simplified solid bonding configuration can result in
a high deviation of the harmonic impedance.
Key point:
The type of bonding significantly affects the positive-sequence harmonic impedance but has only a
minor effect on the zero-sequence harmonic impedance.
App Figure D.18 Positive-Sequence Harmonic Impedance Comparison for Different Cable Bonding
Configurations
225
TB 766 - Network modelling for harmonic studies
App Figure D.19 Zero-Sequence Harmonic Impedance Comparison for Different Cable Bonding
Configurations
Key point:
The number of major sections affects the frequency, magnitude and number of resonant peaks in
the positive-sequence.
226
TB 766 - Network modelling for harmonic studies
App Figure D.20 Positive-Sequence Harmonic Impedance Comparison for Different Numbers of Major
Sections
App Figure D.21 Zero-Sequence Harmonic Impedance Comparison for Different Numbers of Major
Sections
227
TB 766 - Network modelling for harmonic studies
Cable Model
In this section, the use of single or cascaded PI -sections are compared with the frequency-dependent
(FD) distributed parameter phase model. Since modelling of the sheath bonding is not possible when
the PI model is used and the positive- and zero-sequence impedances from the line constants routine
are calculated for a single cable section (and not for the complete cross-bonded cable) the comparison
between the PI and FD models was performed when the latter was configured to use solid bonding.
The harmonic impedance comparison of the models can be seen in Figure D.22. When a single PI
section was used, only the first parallel resonance is present which is shifted to a lower frequency. As
the number of PI sections increases, more resonance peaks are represented, and their frequencies
increasingly shift towards those of the FD model. However, due to the representation of the multiple PI
sections at 50Hz, the magnitude of the resonance peaks is significantly higher than those of the FD
model.
Key point:
Model type should be selected carefully depending on the study. In the case of the PI model, the
number of PI sections used influences the number of resonance points able to be modelled.
App Figure D.22 Positive-Sequence Harmonic Impedance Comparison for Different Cable Models
228
TB 766 - Network modelling for harmonic studies
Load: 95 + j 31 MVA
Transformer: 230 / 69 kV (Y / Δ), 150 MVA, X = 10%, QF=50
The load models selected for the studies are (Figure E.2):
CIGRE model described in Section 3.4.1.3
o Similar to Method 1 (CIGRE JTF 36.05.02/14.03.03); Passive Load – domestic
o Illustrated in Figure 3-37
o Referred to in this Appendix as “Rs X”
o It should be noted that this model uses the following expressions to define R and X:
𝑈2
𝑅=𝑃
𝑃2 + 𝑄2
𝑈2
𝑋=ℎ𝑄 2
𝑃 + 𝑄2
229
TB 766 - Network modelling for harmonic studies
Rs Rs Rm
Rp Xp Xp
Xs Xs Xm
In case of reactive power compensation, a shunt capacitor is added in parallel to the load.
The following cases were studied for load represented at 230 kV, 69 kV, 220 V and 220 V through a
system connection equivalent of 69 kV and 13.8 kV systems. Results are provided in the next pages.
Key Points:
The study example demonstrated that the load representation at 220 V with a detailed
representation of 69 kV and 13.8 kV network and through a system connection equivalent can
predict the system resonance. The choice of load model did not significantly change the system
response, as seen at the 230 kV bus.
Load representation at 230 kV and 69 kV (that did not include the downstream impedance) do not
capture resonance.
Results confirmed the importance of connecting the load model through an equivalent downstream
impedance.
230
TB 766 - Network modelling for harmonic studies
App Figure E.3 Frequency Scan at 230kV Bus - Load Represented at a 230 kV Bus
231
TB 766 - Network modelling for harmonic studies
App Figure E.4 Frequency Scan at 230kV Bus - Load Represented at 69 kV Bus
232
TB 766 - Network modelling for harmonic studies
App Figure E.6 Frequency Scan at 230kV Bus - Load Represented at a 220 V with Explicit
Representation of 69kV and 13.8kV Systems
233
TB 766 - Network modelling for harmonic studies
App Figure E.8 Frequency Scan at 230kV Bus - Load Represented at a 220 V through a Reduced
Equivalent Representation of the 69kV and 13.8kV Systems
234
TB 766 - Network modelling for harmonic studies
13.8 kV HVDC
+
C15 735 kV GND
GND
315 kV
0.25uF 6nF
C4 5nF B6P_1
+
0.01472,0.1472Ohm
C6
0,5.888Ohm
+
GND 6-pulse bridge
GND
+
350 MVA CVT
Electra 167 735/315/15 3-phase
RL2
+
80.96
2
R1
RL1
1 YgYgD_np1
+
C16
+
315/120
BUS5 15.5 km 2
+
+
C8
200nF Electra 167 YD_1 4.7nF 1
+
2 1 3 C3
YYD_1 gates -
4.7nF
GND
4.7nF
a FDline1 a
+
C12 C11 GNDGND
10nF 3nF b + FD b GND 3 B6P_2
315/120 c Electra 167
GND GND c 735/315/15 735/315/15 6-pulse bridge
GND
Electra 167 YD_2 + 2
+
C14
2 1 1 3-phase
+
C13 C10
10nF 3nF
+
3 C2
YYD_2
GND 4.7nF
GND
+
+
C17 C9
+
0.25uF 4.7nF C1 GND C7 gates -
4.7nF GND 5nF
+
C5
350 MVA
0.01472,0.1472Ohm
0,5.888Ohm
Z(f) measured by
R2
RL3
1650 MVA
+
C18
+
+
315/120
Z(f)
Electra 167 YD_3
2 1
+
0,5.888Ohm
R3
RL5
350 MVA
+
C20
+
+
200nF
App Figure F.1 HQ Field Measurement Setup to Measure Z(f) Seen at 315 kV
Measurement data can be seen in Figure F.2 [60] and corresponding simulation results in Figure F.3.
235
TB 766 - Network modelling for harmonic studies
Measurements of the machine impedance of a 50 MVA 13.8 kV hydraulic, salient pole machine (X’’d=0.3
pu and X’d=0.4 pu) over a range of frequencies are shown in Figure F.4.
Table F.1 shows the variation of R and X variation vs. frequency of this machine.
236
TB 766 - Network modelling for harmonic studies
It is worth noting that T’’d should normally be used to estimate the resistance of the machine at power
frequency (IEEE assumes X/R at 60 Hz as being around 10 which is somewhat high for 50 MVA
machines but low for 370 MVA). This resistance is then increased by the factor ℎ. The Electra model is
more pessimistic because it considers √ℎ.
The harmonic impedance of a 120MVA hydraulic machine (based on measurements and manufacturer
design data) is presented in Figure F.5. In conjunction with Table F.1, this shows that the harmonic
reactance can either increase or decrease with frequency.
Sensitivity Analysis
A sensitivity analysis using 13.8 kV machines (illustrated in Figure F.6) shows the importance of
modelling of high frequency (HF) losses which have an impact on the magnitude of resonant peaks.
237
TB 766 - Network modelling for harmonic studies
Model Simplification
This case study shows that once the model has been validated against measurement data, it is possible
to represent the power station from the HV side as a simplified Rs-Ls||Rp equivalent. This is illustrated
in Figure F.7. The validation shows that the curves are almost identical, proving that the approximation
is highly accurate.
238
TB 766 - Network modelling for harmonic studies
239
TB 766 - Network modelling for harmonic studies
Ie = Z’m (ZmZ’m)-1Vm
Validation
To check the accuracy of this method, a set of random harmonic current injections has been simulated
on an REE transmission network model. For a small subset of buses, the calculated voltages are
considered as measured voltages. The previously-described method is then applied to these harmonic
voltages, the network and the network data in the model. The results are shown below:
Even when the harmonic current sources are poorly estimated, the values of the resulting harmonic
voltages are more precise. The application of this method to a large number of different scenarios can
increase its accuracy.
An example of the results for the estimation of the 5th harmonic voltage using real scenarios and real
measurements for October and November 2015 is shown below:
App Figure G.2 Example of Results of the Harmonic State Estimation Method
Another important issue is the location of the new PQ monitoring devices. These should maximize the
visibility of the PQ estate of the whole network. It is also important to analyse the location of new devices
in which the errors could be minimal. More PQ monitoring devices will result in the program yielding
more accurate results.
240
CIGRE
21, rue d'Artois
75008 Paris - FRANCE
© CIGRE
ISBN : 978-2-85873-468-9