You are on page 1of 205

Reliability Evaluation of Electric Power

Systems Integrating Cable Design and Ageing

A thesis submitted to The University of Manchester for the Degree of

Doctor of Philosophy

in the Faculty of Science and Engineering

2018

Shuran Liu

School of Electrical and Electronic Engineering

Electrical Energy and Power Systems Group


Table of Contents

TABLE OF CONTENTS

Table of Contents ............................................................................................... 3


List of Figures ..................................................................................................... 8
List of Tables .................................................................................................... 11
Nomenclature .................................................................................................. 13
Abstract.............................................................................................................. 20
Declaration ....................................................................................................... 21
Copyright Statement...................................................................................... 22
Acknowledgements ....................................................................................... 23
1 Introduction ............................................................................................. 24
1.1 Power System Reliability ............................................................................................ 24
1.1.1 Reliability Evaluation Methods ............................................................................. 26
1.1.2 Reliability & Flexibility ............................................................................................. 29
1.1.3 Asset Management & Ageing.................................................................................. 31
1.1.4 Reliability Cost-Worth Consideration ................................................................ 33
1.2 Review of Existing Cable Rating Methods ........................................................... 35
1.2.1 Continuous Rating ...................................................................................................... 35
1.2.2 Cyclic Rating ................................................................................................................. 36
1.2.3 Probabilistic Rating ................................................................................................... 37
1.2.4 Emergency Rating ...................................................................................................... 37
1.2.5 Dynamic Rating ........................................................................................................... 39
1.3 Cable Uprating Scenarios: Needs and Risks ....................................................... 40
1.4 Past Research on Network Reliability and Cable Ageing ............................. 41
1.4.1 Power System Reliability......................................................................................... 42
1.4.2 Cable Thermal Modelling ........................................................................................ 42
1.4.3 Cable Ageing ................................................................................................................. 43
1.4.4 Summary of Past Research ..................................................................................... 44
1.5 Research Aims and Objectives.................................................................................. 45
1.6 Main Contributions of This Research .................................................................... 46
1.7 Thesis Structure .............................................................................................................. 48

3
Table of Contents

2 Reliability Evaluation Framework Integrating Cable Design


and Ageing ......................................................................................................... 51
2.1 Introduction...................................................................................................................... 52
2.2 Overview of the Proposed Reliability Evaluation Framework .................. 54
2.3 Initial Network Component Status Mapping ..................................................... 57
2.4 Network Optimisation Calculations....................................................................... 60
2.5 Thermal-Ageing Modelling of UGCs ....................................................................... 61
2.5.1 UGC Steady-state Thermal Modelling ................................................................. 63
2.5.2 UGC Transient Thermal Modelling ...................................................................... 66
2.5.3 UGC Thermal Ageing Modelling ............................................................................ 69
2.6 Thermal-Ageing Modelling of OHLs....................................................................... 70
2.6.1 OHL Steady-state Thermal Model ........................................................................ 70
2.6.2 OHL Transient Thermal Model ............................................................................. 70
2.6.3 OHL Thermal Ageing Estimation .......................................................................... 71
2.7 Branch and Network Performance Indices ........................................................ 71
2.7.1 Branch Indices ............................................................................................................. 72
2.7.2 Transmission Network Indices ............................................................................. 72
2.7.3 Distribution Network Performance Indices ..................................................... 73
2.8 Summary ............................................................................................................................ 74

3 Impact of Cables’ High-loading Risks on Network Reliability . 76


3.1 Introduction...................................................................................................................... 77
3.2 Methodology Considering Cables’ High-Loading Risks ................................ 79
3.2.1 Outline of a Double Sequential Monte Carlo Simulation ............................. 79
3.2.2 1st SMC loop Component Status Mapping Calculation ................................ 81
1) Initial network component status mapping – Normal failures ............. 81
2) High-loading component status mapping – Emergency conditions ... 82
3.2.3 2nd SMC-loop Component Status Mapping Computations ........................ 85
3.3 Case Study Formation .................................................................................................. 87
3.3.1 Selection of Test Cable Properties ....................................................................... 88
3.3.2 Modelled Scenarios .................................................................................................... 89
3.4 Effect of Cable High-loading Risks on Network Flexibility.......................... 90
3.5 Cable Life-Cycle Analysis ............................................................................................ 93
3.6 Summary ............................................................................................................................ 97

4
Table of Contents

4 Impact of Extensive Emergency Ratings on Network Reliability


and Plant Ageing ............................................................................................. 99
4.1 Introduction ....................................................................................................................100
4.2 Methodology ...................................................................................................................103
4.2.1 Outline of the Proposed Methodology ..............................................................103
4.2.2 Cascading Failure Modelling ................................................................................105
4.2.3 Modelling of the Post-contingency Operation of Circuits .........................107
4.3 Description of Case Studies......................................................................................110
4.3.1 Test Network – 14-bus, 24-bus OHL & UGC network.................................111
4.3.2 Modelled Scenarios ..................................................................................................114
4.4 Impact of Extensive Emergency Ratings in the OHL Networks...............115
4.5 Impact of Extensive Emergency Ratings in the UGC Networks ...............120
4.6 Summary ..........................................................................................................................124

5 Increasing Wind Power Integration through Flexible Cable


Current Ratings ............................................................................................. 126
5.1 Introduction ....................................................................................................................127
5.2 Proposed Methodology ..............................................................................................129
5.2.1 Outline of the Proposed Methodology ..............................................................129
5.2.2 Cable Rating Computation ....................................................................................131
1) DTS Modelling ........................................................................................................131
2) Dynamic, Flexible and Forecasting Rating Calculation ..........................132
5.2.3 Wind Power Forecasting using ARMA .............................................................135
5.2.4 Metrics of Long-term Wind Integration Risk-Worth Analysis ................136
5.3 Case Study Design ........................................................................................................138
5.3.1 Design of Test Network – wind farm cable circuit ......................................138
5.3.2 Rating Scenarios Considered for the Test Network ....................................139
5.3.3 Modelled Rating Scenarios ...................................................................................141
5.4 Impact of the Excursion Times of Probabilistic Cable Rating ..................141
5.5 Impact of Flexible Cable Rating on Wind Integration ..................................143
5.6 Impact of Forecasted Flexible Cable Rating on Wind Integration ..........146
5.7 Long-term Economic Assessment of FCR and FFCR .....................................147
5.8 Summary ..........................................................................................................................149

5
Table of Contents

6 Quantifying the Risk of Emergency Ratings on Distribution


Cable Network with EVs............................................................................. 151
6.1 Introduction.................................................................................................................... 152
6.2 Methodology................................................................................................................... 154
6.2.1 Overview of the Proposed Methodology ......................................................... 154
6.2.2 Asset Failure and Repair Modelling .................................................................. 156
6.2.3 Electric Vehicle Charging and Restoration Modelling ................................ 158
3) EV charging modelling ........................................................................................ 158
4) EV demand restoration modelling ................................................................. 160
6.2.4 Network Restoration Modelling ......................................................................... 161
6.3 Case Study Formation ................................................................................................ 164
6.3.1 Selection of the Feeder Cables............................................................................. 164
6.3.2 Modifications of the RBTS bus 4 network ...................................................... 165
6.3.3 Modelled Scenarios .................................................................................................. 168
6.4 Impact of EVs on Network and Cable Performance ..................................... 169
6.5 Impact of Remote Switching on EV Penetration ............................................ 170
6.6 Impact of Emergency Ratings on EV Penetration and Cable Ageing .... 173
6.7 Summary .......................................................................................................................... 176

7 Conclusion & Future Work ............................................................... 178


7.1 Conclusion ....................................................................................................................... 178
7.2 Future Work ................................................................................................................... 182

8 References ............................................................................................. 185


Appendix A: Network Data ....................................................................... 194
A.1. IEEE 14-bus Test System ..................................................................................... 194
A.1.1 Bus Data ....................................................................................................................... 194
A.1.2 Branch Data ................................................................................................................ 195
A.1.3 Generator Data ......................................................................................................... 195
A.2. IEEE 24-bus Reliability Test System (RTS-96) ....................................... 196
A.2.2 Branch Data ................................................................................................................ 197
A.2.3 Generator Data ......................................................................................................... 198
A.2.4 Branch Reliability Data ........................................................................................ 199
A.3. Roy Billinton Reliability Test System (RBTS bus 4) ............................ 200

6
Table of Contents

A.3.1 Load Data .....................................................................................................................200


A.3.2 Reliability Data .........................................................................................................200
A.4. Chronological Load Profile .................................................................................201

Appendix B: Cable Design and Rating Data ......................................... 203


Appendix C: Publications ........................................................................... 205

7
List of Figures

LIST OF FIGURES
Figure 1-1 Hierarchical Levels ....................................................................................................... 25

Figure 1-2 Reliability evaluation types, network levels, methods (the approach used
in the thesis in bold) ........................................................................................................................... 27

Figure 1-3 Drivers of needed flexibility and selected flexibility solutions ................... 30

Figure 1-4 Factors influencing asset management decision .............................................. 32

Figure 1-5 Reliability cost-worth concept ................................................................................. 34

Figure 2-1 Flowchart of the network reliability evaluation methodology ................... 55

Figure 2-2 Flow Chart for Initial Network Component Status Mapping ........................ 58

Figure 2-3 an example of modelling the component state transition ............................. 59

Figure 2-4 Cable thermal calculation flowchart at Δt [62] .................................................. 61

Figure 2-5 TEE model for a single core cable with n soil layers. ...................................... 62

Figure 2-6 Steady-state TEE model for single core cable. ................................................... 63

Figure 3-1 Flowchart of the double SMC methodology [62] .............................................. 80

Figure 3-2 Operation states of a cable i considering its normal (i') and high-loading
(i") conditions [62] ............................................................................................................................. 82

Figure 3-3 An example of modelling the operating state and corresponding current
flow transitions [62]........................................................................................................................... 85

Figure 3-4 Single line diagram of IEEE 14-bus network ...................................................... 87

Figure 3-5 EENS, EFNF of Sc-4 and base case across a cable design life of 30 years 93

Figure 3-6 EENASS and EENATS of Sc-4 and base case from a steady and transient state
thermal model across a cable design life of 30 years ............................................................ 94

Figure 3-7 EDEL results of C10, C11, C13 cables throughout their life cycle. .............. 95

Figure 3-8 EFEL results of C10, C11, C13 cables throughout their life cycle. .............. 96

Figure 4-1 Flowchart of the network evaluation integrating cascading failure,


emergency rating operation and circuit ageing ....................................................................103

8
List of Figures

Figure 4-2 Flowchart of fault chain search model to capture cascading failures .... 106

Figure 4-3 Flowchart of the ‘post-contingency operation of circuits’ block ............. 108

Figure 4-4 An example of modelling the state transition and corresponding current
flows ...................................................................................................................................................... 110

Figure 4-5 Single line diagram of IEEE RTS-96 24-bus network ................................... 111

Figure 4-6 Time-varying ambient and soil temperature in a simulation year ......... 114

Figure 4-7 EENS, EDLC output in the time domain of DSTE for the 14-bus (left) and 24-
bus OHL networks (right) ............................................................................................................. 115

Figure 4-8 EFCCF and EENA of the modelling scenarios in the time domain of DSTE for
the 14-bus (left) and the 24-bus OHL networks (right) .................................................... 117

Figure 4-9 EAIC of the modelling scenarios in the time domain of DSTE for the 14-bus
(left) and 24-bus OHL networks (right) .................................................................................. 119

Figure 4-10 EENS/EDLC of Sc-1, Sc-3 and Sc-5 across the different ER uprating factors
U in the 14-bus (left) and 24-bus UGC networks (right). .................................................. 120

Figure 4-11 EFCCF, EENA of Sc-1, Sc-3 and Sc-5 across different ER uprating factor U
in 14-bus (left) and 24-bus UGC network (right). ............................................................... 122

Figure 4-12 EAIC of Sc-1, Sc-3 and Sc-5 across the different ER uprating factors U in
the 14-bus (left) and 24-bus UGC networks (right). ........................................................... 123

Figure 5-1 Method to evaluate different cable utilisation on wind farm circuits.... 130

Figure 5-2 Measured and calculated duct temperatures based on cable loading ... 131

Figure 5-3 Computational steps for DCR, FCR and FFCR calculations at Δt .............. 133

Figure 5-4 Example of FCR and FFCR application under a short-term high wind power
output .................................................................................................................................................... 134

Figure 5-5 Long-term cost-worth analysis of the different rating methods ............. 138

Figure 5-6 Probabilities of the modelled (left) and recorded (right) wind farm output
being at different levels of the maximum wind farm capacity........................................ 139

Figure 5-7 Frequency distributions of the soil’s annual, winter, spring, summer and
autumn temperatures. .................................................................................................................... 139

9
List of Figures

Figure 5-8 Excursion times and temperature distribution under SCR and PCR. .....140

Figure 5-9 EWPC, EECA and ECUP outputs of SCR, SeCR and PCR 20%, 60% and 100%
methods under a range of wind farm integration capacities. ..........................................142

Figure 5-10 Risk of thermal exceedance and EECA across different PCR-Pex%s with
2.0 pu of wind farm capacity considering 60 years of wind and soil data. .................143

Figure 5-11 Calculated EWPC, EECA and ECUP for standard SCR, DCR, and FCRs with
various predefined soil properties and increased installed wind capacity. ...............144

Figure 5-12 Mean µ, µ±σ (red) and µ±2σ (blue) ranges of the cable’s thermal
exceedance and EECA, with varying soil temperature FCR and 2 pu wind capacity.
..................................................................................................................................................................145

Figure 5-13 FFCR Seasonal performance assessment at different temporary


temperature limits and forecasted durations under 1.5pu wind capacity. ................146

Figure 5-14 Profits generated from various uprating schemes against SCR when a 50%
increase in wind power integration is considered across a 30 yr operation. ............149

Figure 6-1 Methodological flowchart for evaluating the EV impact in the distribution
network .................................................................................................................................................154

Figure 6-2 EV arrival pattern and average charging power demand pattern ...........159

Figure 6-3 An example of power restoration in a two-legged ring network with or


without EV connection ....................................................................................................................161

Figure 6-4 Single line diagram of the modified RBTS bus 4 distribution network ..166

Figure 6-5 EENS, EENA, SAIDI and SAIFI output of Sc-4 when utilising normal, cyclic
and emergency ratings ....................................................................................................................174

Figure 6-6 EDNS and EDEDI output of Sc-4 when utilising normal, cyclic and
emergency ratings.............................................................................................................................175

10
List of Tables

LIST OF TABLES
Table 2-1Thermal Matrix M & Layer Thermal Properties................................................... 67

Table 3-1 Cable Electrical Properties Modelled in an IEEE 14-bus Network .............. 88

Table 3-2 Cable reliability properties modelled in the IEEE 14-bus network............. 89

Table 3-3 Modelling Scenarios for Emergency UGC Risk of Failure ................................ 90

Table 3-4 Summary of network performance indices .......................................................... 91

Table 4-1 Types and Levels of Emergency Rating ............................................................... 107

Table 4-2 Branch Electrical Properties Modelled in the 14 and 24-bus Networks 112

Table 4-3 Branch outage rate and duration in IEEE 14-bus network.......................... 113

Table 4-4 Modelling Scenarios of the Emergency Rating Operations .......................... 114

Table 5-1 Initial Conditions for DCR, FCR and FFCR Calculations................................. 132

Table 5-2 Soil Properties Used for Static and Seasonal cable rating ............................ 140

Table 5-3 Modelling Scenarios of Wind farm Cable Ratings ............................................ 141

Table 5-4 Key Performance Indicators of the Different Rating Methods under 1.5pu
Wind Capacity .................................................................................................................................... 147

Table 6-1 Reliability parameters of the distribution assets in the modified RBTS 158

Table 6-2 The Cable Electrical Properties Modelled in the RBTS bus 4 network ... 164

Table 6-3 Thermal Parameters in XLPE Cables and Soil ................................................... 165

Table 6-4 Number of EV charging slots on each bus at PEV% penetration level .... 167

Table 6-5 Modelling Scenarios of EV impact on the distribution network ................ 168

Table 6-6 Network performance indices of the base case and Sc-1 ............................. 169

Table 6-7 The output of EENS, EDNS, EDEDI and EENA in Sc-1, Sc-2, Sc-3 and Sc-4
across different EV penetration levels ..................................................................................... 171

Table A-1 Bus Data of IEEE 14-bus Test System .................................................................. 194

Table A-2 Branch Data of IEEE 14-bus Test System ........................................................... 195

Table A-3 Generator Data of IEEE 14-bus Test System ..................................................... 195

11
List of Tables

Table A-4 Bus Data of IEEE 24-bus Reliability Test System .............................................196

Table A-5 Branch Data of IEEE 24-bus Reliability Test System ......................................197

Table A-6 Generator Data of IEEE 14-bus Test System ......................................................198

Table A-7 Branch Reliability Data of IEEE 24-bus Reliability Test System ................199

Table A-8 Load Point Data of RBTS bus 4 network ..............................................................200

Table A-9 Reliability Data of RBTS bus 4 network ...............................................................200

Table A-10 Hourly load profile in percentage of peak load ..............................................201

Table A-11 Daily load profile in percentage of peak load ..................................................202

Table A-12 Weekly load profile in percentage of peak load .............................................202

Table B-1 Design properties for cables in 14 and 24-bus network ...............................203

Table B-2 Design properties for cables in RBTS and windfarm circuit ........................204

12
Nomenclature

NOMENCLATURE

List of Abbreviations
AC Alternative Current

AENS Customer Average Energy Not Supplied

AM Analytical Method

ANM Active Network Management

ARMA Auto-Regressive Moving Average

BC Base Case

CF Cascading Failure

Con Continuous

Cov Covariance

DAL Drastic Action Limit

DC Direct Current

DCR Dynamic Cable Rating

DG Distributed Generation

DLR Dynamic Line Rating

DNO Distribution Network Operator

DSMC Double Sequential Monte Carlo

DSR Demand Side Response

DTR Dynamic Thermal Rating

DTS Distributed Temperature Sensing

EAIC Expected Ageing & Interruption Cost

EANL Expected Annual Network Losses

ECA Equivalent Cable Ageing

ECUP Expected Cable Utilization Profit

EDEI Expected Duration of EV charging demand Interruption

13
Nomenclature

EDEL Expected Duration of Emergency Loading

EDLC Expected Duration of Load Curtailment

EDNS EV Demand Not Satisfied

EECA Expected Equivalent Cable Ageing

EELA Expected Equivalent Line Ageing

EENA Expected Equivalent Network Ageing

EENS Expected Energy Not Supplied

EFCCF Expected Frequency of Circuit Cascading Failure

EFCF Expected Frequency of Cable Failure

EFEI Expected Frequency of EV charging demand Interruption

EFEL Expected Frequency of Emergency Loading

EFLC Expected Frequency of Load Curtailment

EFLF Expected Frequency of Line Failure

EFNCF Expected Frequency of Network Cascading Failure

EFNF Expected Frequency of Network Failure

EIC Expected Interruption Cost

ELA Equivalent OHL Ageing

ENS Energy Not Supplied

ENWL Electricity North West Limited

ER Emergency Rating

ETC Electro-Thermal Coordination

EV Electric Vehicle

EWPC Expected Wind Power Curtailed

FACTs Flexible AC Transmission System

FCR Flexible Current Rating

FEM Finite Element Model

FFCR Forecasted Flexible Current Rating

14
Nomenclature

HL Hierarchical Level

HS Health Score

HLF High-Loading Failure

HVDC High Voltage Direct Current

IEC International Electrotechnical Commission

IEEE Institute of Electrical and Electronic Engineers

LTE Long-Term Emergency

MC Monte Carlo

MCS Monte Carlo Simulation

MCS Monte Carlo Simulation

NLF Normal Line Failure

NLR Normal Line Rating

OHL Overhead Line

OPF Optimal Power Flow

PCR Probabilistic Cable Rating

PF Power Flow

RBTS Roy Billinton Test System

RES Renewable Energy Sources

RTS Reliability Test System

RTTR Real Time Thermal Rating

SAIDI System Average Interruption Duration Index

SAIFI System Average Interruption Frequency Index

Sc Scenario

SCR Static Cable Rating

SeCR Seasonal Cable Rating

SMC Sequential Monte Carlo

SMCS Sequential Monte Carlo Simulation

15
Nomenclature

SPEN Scottish Power & Energy Networks

SS Steady State

STE Short-Term Emergency

TEE Thermo-Electric Equivalent

TS Transient State

TSO Transmission System Operator

TTF Time-to-Failure

TTR Time-to-Repair

UGC Underground Cable

UK United Kingdom

USA United States of America

V2G Vehicle-to-Grid

XLPE Cross-linked Polyethylene insulators

List of Symbols
fDlp (∙) Damage cost function for curtailed demand at point lp

fPg (∙) Real power generation cost function of generator g

fQg (∙) Reactive power generation cost functions of generator g

b Network branch number ∀b∈Nbranch

bus Network bus number running from 1 to Nbus

CAr Armour thermal capacitance

CC Conductor thermal capacitance

Cd Insulation thermal capacitance

CJ Jacket thermal capacitance

Costlp Cost function of customer power interruption

CS Screen thermal capacitance

16
Nomenclature

CSL(k) Thermal capacitance of soil layer k (1≤k≤n)

Csoil Thermal capacitance of the soil

g Network generator number running from 1 to Ng

i Network cable, ∀i∈ NUGC

i' Normal-loaded cable, ∀i'∈ NUGC

i" High-loading cable, ∀i"∈NUGC

IDAL Current rating at drastic action limit

Iflow Current flow

ILTE Long-term emergency current rating

Inormal Normal current rating

ISTE Short-term emergency current rating

j Overhead line, ∀j∈ NOHL

k Layer number for cable and soil running from -m to n

lp Network load point number running from 1 to Nlp

Nbranch Total number of branches in the network, Nbranch=NUGC+NOHL

NCus Total number of customers served in the network

Ng Total number of generators in the network

Nlp Total number of load points in the network

NOHL Total number of overhead lines in the network

NUGC Total number of cables in the network

Nx Total number of components in the network


Maximum number of repetitions of annual simulations within
NY
the 1st SMC-loop
TAir Thermal resistance of the air between cable and duct

TArm Thermal resistance of the armour

Tduct Thermal resistance of the duct

TIns Thermal resistance of the insulation

TJ Thermal resistance of the jacket

17
Nomenclature

TSL(k) Thermal resistance of soil layer k (1≤k≤n)

Tsoil Thermal resistance of the surrounding soil

TTFE Emergency time-to-fail

TTFN Normal time-to-fail

TTRN Normal time-to-repair

TTRz Time-to-repair for cascading failure z

tΔt Time vector for time-step analysis (1st SMC-loop)

tθ Time vector for cable thermal modelling

UN Normal unavailability

Vwind wind velocity

WAr Armour losses

WC Conductor losses

Wd Dielectric losses

WS Screen losses

x Network component number running from 1 to Nx

Y Simulation year running from 1 to NY

YL Simulation life-cycle year running from 1 to YLife

YLife Maximum number of life years for network life-cycle analysis

α Design risk coefficient

β Ageing acceptability risk coefficient

Δt Time-step of SMC annual computation, ∀Δt∈Y

θAir Air temperature within the duct

θAmb Ambient temperature

θAr Armour temperature

θC Conductor temperature

θC,Max Maximum conductor temperature, usually 90°C for XLPE cables

θDTS External duct temperature measured by DTS

18
Nomenclature

θS Screen temperature

θSL(k) Soil temperature at the layer k

θSoil Ambient soil temperature

λE Emergency failure rate

λN Normal failure rate

μN Normal repair rate

ρSoil Soil thermal resistivity

19
Abstract

ABSTRACT
Reliability Evaluation of Electric Power Systems Integrating Cable Design and Ageing
Shuran Liu, The University of Manchester, Dec 2018

The restructuring of the electricity sector forces utilities to operate their existing
power networks with increasing flexibility and efficiency. Alternative cable rating
methods, such as emergency and dynamic ratings, improve the utilisation of existing
network cables, and provide cable-based networks with additional flexibility and
resilience against uncertainties. However, such approaches may pose more stress on
network cables and increase the risk of thermal overloading and ageing.
This thesis advances the existing power system reliability assessment methodology
by integrating a detailed cable system thermal-ageing model that considers cable
design properties and environmental conditions into a traditional sequential Monte
Carlo simulation. The main outcome of this integration was the comprehensive
probabilistic evaluation of cable networks’ reliability, together with each individual
cable’s thermal performance and ageing risk. The methodological framework
integrating cable design and ageing provides operators with the additional tools to
identify key network cables based on a quantified ageing index in order to perform
optimised cable replacement and maintenance strategies. The framework has also
helped utilities and operators to propose alternative cable operating practices to
increase the loadability and flexibility of cable-based networks for better network
performance, but also to avoid excessive risk of cable ageing.
The methodological framework was applied on four test networks for four specific
case studies. The first study presented an innovative methodology to capture the
additional risk of cable failure and ageing when cables are operated beyond their
limits. The second study further quantified the network reliability benefits and cable
ageing risks when extended durations and increased ampacity of the emergency
ratings are implemented. These two studies will inform the utilities of the benefits
and risks of implementing emergency rating of UGCs and to help them to make
decisions related to optimising their operational flexibility. The third study proposed
an alternative flexible cable current rating method to optimise the loading capability
of existing cable tie lines between wind farms and host utilities. It allows the wind
farm developers to increase their generation capacities without upgrading their
existing cable tie circuits and risking the cable lifetime. The last study evaluated the
benefits and potential risks of emergency ratings of UGCs could give in a distribution
cable network with EV connections. The proposed emergency rating approach
enables for a higher penetration level of EVs while an acceptable level of cable ageing
is generated.

20
Declaration

DECLARATION
No portion of the work referred to in the thesis has been submitted in support of an
application for another degree or qualification of this or any other university or other
institute of learning.

21
Copyright Statement

COPYRIGHT STATEMENT
The author of this thesis (including any appendices and/or schedules to this thesis)
owns certain copyright or related rights in it (the “Copyright”) and s/he has given
The University of Manchester certain rights to use such Copyright, including for
administrative purposes.

Copies of this thesis, either in full or in extracts and whether in hard or electronic
copy, may be made only in accordance with the Copyright, Designs and Patents Act
1988 (as amended) and regulations issued under it or, where appropriate, in
accordance with licensing agreements which the University has from time to time.
This page must form part of any such copies made.

The ownership of certain Copyright, patents, designs, trademarks and other


intellectual property (the “Intellectual Property”) and any reproductions of copyright
works in the thesis, for example graphs and tables (“Reproductions”), which may be
described in this thesis, may not be owned by the author and may be owned by third
parties. Such Intellectual Property and Reproductions cannot and must not be made
available for use without the prior written permission of the owner(s) of the relevant
Intellectual Property and/or Reproductions.

Further information on the conditions under which disclosure, publication and


commercialization of this thesis, the Copyright and any Intellectual Property and/or
Reproductions described in it may take place is available in the University IP Policy
(see http://documents.manchester.ac.uk/DocuInfo.aspx?DocID=24420), in any
relevant Thesis restriction declarations deposited in the University Library, The
University Library’s regulations (see
http://www.library.manchester.ac.uk/about/regulations/) and in The University’s
policy on Presentation of Theses.

22
Acknowledgements

ACKNOWLEDGEMENTS
First, I would like to express my gratitude to my supervisor Dr Konstantinos Kopsidas
for his patience, guidance and support during my PhD. I sincerely appreciate his
helpful advice, comments and discussion which have contributed a lot to my research
and my personal development. Thanks him to bring me to Manchester, encourage me
to become a better man and complete my PhD. Although he has strict requirements
on research, Kostas is really a nice guy deep in his heart.

I also would like to thank my colleagues and friends in the Ferranti building and our
research group at The University of Manchester, Dr Jingli Guo, Dr Nan Liu, Dr Wei
Zheng, Dr Shahnurriman Abdul Rahman, Mr Yucong Zhao, Mr Carlos Cruzat
Hermosilla, Mr Mohamed Galeela, Mr Mohammed Abdulaziz Al Aqil. Thanks for the
fruitful discussions and their encouragement throughout the period of my PhD study.
Wish you all have a bright and successful future.

Finally, I really would like to express my thanks to my parents, Deyou Liu and Jinhua
Zha, for their unconditional love and support throughout my life. I believe you would
be very happy and pound of me.

Equation Chapter (Next) Section 1

23
Chapter 1: Introduction

1 Introduction

Modern society has a growing reliance on electricity. Electric power utilities must
provide a reasonable assurance of quality and the continuity of service to their
customers. Hence, an accurate, rapid and comprehensive reliability evaluation is
important in relation to the planning and operations of a power system.
Underground transmission networks are becoming increasingly important,
especially for urban areas. There is a growing necessity to conduct a
comprehensive reliability assessment for underground cable networks to use to
consider their electrical and thermal performance.

This research has been conducted to integrate cable design properties and ageing
characteristics into reliability evaluations for cable-based power grids. More
importantly, it was conducted to explore the potential improvements and
challenges that cable networks can face. More flexible operating schemes for cable
networks can provide more flexibility to the power system’s operations,
contributing to network reliability. However, this may also lead to additional risks,
such as cable ageing and failure. These benefits and risks need to be quantified in
order to provide utilities with more informed suggestions related to their optimal
planning, operation and asset management strategies.

This chapter provides a brief literature review of the related past research,
together with the overview, motivation and aims and objectives of this research,
and the contributions of the PhD studies referenced.

Reliability is an inherent characteristic of a component, and it describes its ability to


accomplish the intended functions within the required time under certain operating
conditions. Power system reliability defines the capability of a power system to
continuously provide qualified electricity to the customers under both static and
dynamic conditions [1].

24
Chapter 1: Introduction

Power system reliability can be separated into two fundamental fields: Adequacy and
Security. Adequacy, known as static reliability, is the ability of the power system to
transmit sufficient electricity to its customers. Therefore, it is necessary to include
enough generation, transmission and distribution facilities into the power network
to meet customer demands. Adequacy is related to static conditions rather than any
disturbances that occur during the routine operations of the system. On the other
hand, Security, known as dynamic reliability, defines the ability of a power network
to withstand any disturbance caused by accidental events, for example, a sudden
short circuit or an unexpected failure of components, while still being able to supply
electricity in an uninterrupted manner.

Generation
Facilities
Hierarchical Level I (HL1)

Transmission
Facilities
Hierarchical Level II (HL2)

Distribution
Facilities
Hierarchical Level III (HL3)
Figure 1-1 Hierarchical Levels

Generally, we divide power systems into three zones: Generation, Transmission and
Distribution. Reliability studies are conducted in these three zones or in combination
as shown in Figure 1-1. Hierarchical Level 1 (HL1), often termed ‘generating capacity
reliability assessment’, only focuses on the generation facilities in order to determine
their generating capacity to satisfy the whole demand requirements. The process of
transporting generation to each load point has been neglected in the examined HL1
studies. Hierarchical Level 2 (HL2), also known as ‘bulk transmission system
assessment’, contains power generation and transmission facilities, evaluating the
abilities of the composite system to deliver electricity to bulk load points.
Hierarchical Level 3 (HL3) includes the entire system, from the generation facilities
to the individual load points. However, it is difficult and complicated to directly

25
Chapter 1: Introduction

perform HL3 studies due to their large and complex scale. Thus, HL3 studies are
usually performed separately related to the distribution functional zone, and HL2
outputs are utilised as the input of the distribution zones in order to reduce the
modelling complexity and increase the computation speed.

HL2 and HL3 studies are associated with load flow calculations, contingency analyses,
load curtailment, and therefore consider the structure, operating constraints of the
network, and the reliability and operations of the network facilities. At present,
scholars have achieved a lot regarding algorithms and the application of adequacy
evaluations. However, the thermal-electrical modelling of internal components
within a network-wide reliability study has not been studied a lot.

In this thesis, we have focused on the thermal-ageing modelling of cables within


transmission and distribution networks. Thus, our studies were performed at the
HL2 level for bulk transmission networks, and at the HL3 level for distribution
networks.

1.1.1 Reliability Evaluation Methods

Modern society has a growing reliance on electricity. Electric power utilities must
provide a reasonable assurance of quality and the continuity of service to their
customers. Hence, an accurate, rapid and comprehensive reliability evaluation is
necessary and essential for the planning and operations within power systems.

Network reliability assessment techniques could be grouped into two categories:


deterministic and probabilistic methods. Although both methods are applied by
utilities, most bulk systems utilise probabilistic methods. The relationship between
the different reliability evaluation approaches has been described in Figure 1-2,
where this thesis’s approach has been marked in bold.

The traditional deterministic method utilises a deterministic contingency analysis


that include all network contingencies when any one line, transformer or generator
is out of service. This deterministic criterion, usually named as the N-n criterion,
represents that a N-components system can withstand the removal of any n-

26
Chapter 1: Introduction

components [2]. The N-1 criterion has been largely applied in transmission
expansion planning by utilities.

Power System Reliability

Adequacy Security

HL1 HL2 HL3

Probabilistic Deterministic

Analytical Monte Carlo Simulation

Sequential Non-Sequential
Figure 1-2 Reliability evaluation types, network levels, methods (the
approach used in the thesis in bold)

However, the assumption of only a single contingency is dubious, considering that


the failures hidden in protection systems are able to lead to multiple contingencies
and cascading outages. Extreme weather conditions may also increase the possibility
of multiple simultaneous but independent failures. In conclusion, deterministic
methods cannot take into account the random behaviour of the system and
components, therefore, the deterministic methods cannot consider their impacts.

Probabilistic reliability assessments include two main techniques: analytical


methods (AM) and Monte Carlo simulation (MCS) techniques [3]. The analytical
method evaluates the reliability indices using numerical solutions by representing
the network using mathematical models. The Monte Carlo simulation method
simulates the actual random behaviour and then calculates the system reliability.

Although conceptually simple analytical methods only require low computational


capability, three main disadvantages of AM limit its further application in the power
system reliability evaluation. First, the number of system operating states increases

27
Chapter 1: Introduction

exponentially with the increase of system size. For example for an N-component
network, the number of system states is 2N. Therefore, assumptions are needed to
reduce the number of states, particular for bulk systems. Second, the frequency and
duration indices are difficult to obtain from AM because the time-series relationship
is not considered. However, they do contain important information of the power
system. Lastly, AM cannot simulate stochastic events, such as the load fluctuations,
weather variations and the operators’ control actions.

On the other hand, the Monte Carlo simulation requires harder and heavier
computation, however it is much more versatile to be used for simulating all random
behaviours and contingencies in the system. One of its advantages is the sampling
number which has a given accuracy that is independent to the system size.
Furthermore, by simulating the stochastic incidents and the operators’ control
actions, MCS will provide dependable and informative reliability indices. Therefore,
there is an increasing interest in applying MCS in bulk system reliability evaluations
to model the system’s behaviour more comprehensively.

There are two basic simulation algorithms within the Monte Carlo method:
sequential simulation and non-sequential simulation.

Non-sequential MCS, also known as the ‘state sampling approach’, perform random
state sampling for each system’s component based on the probability that the
component exists in that state. Non-sequential simulations involve a simpler model,
a lower computer memory requirement and a higher convergence speed, which is
more suitable for large power system evaluations and calculations. However, this
state sampling method does not consider the timing sequence information and state
transfers between non-contingency and contingency states. Thus, the frequency and
duration indices are difficult to accurately capture. For this reason, the non-
sequential MCS approach has not been used in this thesis.

Sequential Monte Carlo simulation (SMCS), also known as the ‘state duration
sampling approach’, includes the time sequence information about the system’s
operating states, and thus can accurately simulate the duration time in each state and

28
Chapter 1: Introduction

the shift frequency between states. Thus, all frequency and duration indices can be
easily calculated by modelling actual operations and random aspects in a sequential
time axis. Furthermore, more accurate probabilistic models can be established using
SMCS for their reliability assessments by considering stochastic renewable sources,
chronological load demands and time-varying line ratings. The main disadvantage of
SMCS is that it requires more computation time and memory, because the time
sequential calculation is more complex with a large amount of data and computations.

This thesis considers many time-varying elements, such as renewable generations,


load demand, weather conditions and line ratings in relation to the network
reliability evaluations. More important, this thesis considered the detailed process of
emergency operations in the analysis of the contingency consequences. This was as
well as considering the transient thermal behaviour of underground cables during
emergency operations in their thermal-ageing calculation. For these reasons, the
SMCS approach was selected as the default method for further reliability studies in
this thesis.

1.1.2 Reliability & Flexibility

Power system flexibility is defined in a more general way by EPRI as ‘a power


system’s ability to adapt to dynamic and changing conditions’. For example,
generation and demand balancing, or the deployment of new generation sources and
transmission facilities [4]. The GB energy regulator, Ofgem, focuses on generation
and demand flexibility, by defining flexibility as ‘altering generation and/or
consumption patterns in response to an external signal’ [5].

At present, power utilities provide their networks with flexibility on the ‘supply side’,
ensuring that enough flexible generation is available to always match the variable
load demand. This is because network operators usually have no control schemes
over how and when the customers use electricity. Enough transmission and
distribution facilities are built, ensuring that electricity can always be transported to
the consumers.

29
Chapter 1: Introduction

However, generation is more variable in a low carbon network when renewable


sources are connected, making it impossible to continue to solely rely on supply-side
solutions. Power utilities are challenged to enhance flexibility on the ‘transmission
side’ and ‘demand side’ while providing reliable, secure and affordable electricity to
customers. Therefore, new flexibility methods are emerging, such as demand side
response (DSR), electricity energy storage, distributed generation (DG), FACTS,
HVDC, dynamic ratings and distribution automation.

These new technologies are used to enhance the operational flexibility of power
systems, which helps utilities to connect more renewable energy sources (RES), such
as wind and solar power, into the supply system. Increased flexibility also
strengthens the ability to match supply and demand under various operating
conditions, thus improving overall network reliability.

Solutions to Enhance Flexibility


Distributed generation

Supply Side Wind for active power control

……
Flexibility Drivers Improved
Reliability
Growth in renewable generation Emergency rating

Probabilistic rating
Higher load variability
Enhanced Transmission Side Dynamic rating
Flexibility
Environmental regulations
HVDC

Uncertainties & hidden failures FACTS


Enable more
RES
Smart Switching

Demand Side Demand side response

Electricity energy storage

……

Figure 1-3 Drivers of needed flexibility and selected flexibility solutions

Figure 1-3 illustrates the drivers of flexibility, along with the technologies and
techniques used for enhancing power system flexibility, with the thesis solutions
marked in bold. This thesis focuses on the operational flexibility of transmission &
distribution (T&D) systems, which represents the ability of the power network to

30
Chapter 1: Introduction

manage more uncertain and variable power flows. This T&D network flexibility is
extremely critical to accommodate stochastic renewable generation and sudden
demand changes, as well as being reliable and resilient under situational
uncertainties and network contingencies.

Rapidly demanded growth and increasing renewable integration poses more stress
on T&D networks, forcing the components to operate at their design limits more
frequently, particularly under emergency operations. To overcome this challenge,
utilities are investigating economic and flexible solutions such as emergency ratings,
probabilistic ratings and dynamic ratings, in order to increase the power flow
capacities of the network instead of building new assets. Improved line ratings
provide operators with more flexibility to balance generation and demand without
exceeding the network’s limit. On the other hand, the lack of T&D flexibility may lead
to the inability of the network to satisfy demand from time to time, resulting in
reduced network reliability.

However, these solutions maximise the line and cable performance by increasing
their utilisation in a manner for which they were not originally designed. This poses
potential thermal ageing issues in relation to these assets, and this requires an
improved operation and maintenance strategy. The research in this thesis was
conducted to understand the implications of increased flexibility related to the
potential ageing risks and network reliability benefits, as well as the operation and
maintenance of T&D networks.

1.1.3 Asset Management & Ageing

In the power industry, there are a significant proportion of assets operating beyond
their design life. For example, in urban areas, some distribution cables are more than
50 years old and a great amount of transformers are operating beyond 40 years [6].
At the same time, new challenges are posing more stress to the existing power
networks, such as electric vehicles, renewable supply and distributed energy. Any
ageing asset leads to a reduction in its reliability, thus hampering overall network
reliability performance.

31
Chapter 1: Introduction

There is a growing necessity to enable the optimum usage, maintenance and


replacement of assets, in response to ageing plants, changing generations and loads,
as well as the growth of emerging technologies. The practice of asset management
helps utilities to perform remedial actions to economically maintain their reliability,
prolong the useful lives of the existing assets, and to allow for effective asset
replacement. Figure 1-4 shows the factors affecting the asset management decision,
and four potential actions that could influence network reliability performance.

Asset management is described as ‘the evaluation of assets’ life cycle costs against
supply quality’ [7]. The current industry relies on field tests and condition
measurements for the power equipment, as well as damage models estimating
condition and failure probability [8]. Traditional asset management methods are
performed on an individual basis, which only considers the condition of the asset
itself without considering the system’s operational aspects [9]. The existing research
on asset management strategy based on whole-system analyses is still very limited
[10-12].

History: Environment: Ageing Model


Loading history Ambient temp Field Test
Failure history Cycling load Condition Monitoring
Switching surge Laying condition Measurement
…….. …….. ……..

Asset
Management
Decision

Replacement Change Perform


Do Nothing Environment Maintenance

Impact on Network
Reliability Performance
Figure 1-4 Factors influencing asset management decision

Conservative decisions are usually made by utilities due to their limited


understanding of real insulating conditions and uncertainties in future operating

32
Chapter 1: Introduction

conditions, which leads to higher asset maintenance and replacement costs. This
illustrates the need for a tool to accurately estimate the ageing condition of assets
considering their real operating stresses, environmental conditions, load history and
other system aspects.

In this thesis, a novel reliability framework incorporating the cables’ design and
ageing aspects has been developed for utilities as a useful simulation tool to estimate
thermal ageing and the lifetime of cables based on realistic cable design, actual
network operating data and variable weather conditions. More importantly, this
work evaluates the assets’ ageing and failure cost against the cost of having a reduced
supply quality. Thus, it allows for a reliability-based or cost-based asset management
strategy to be developed, providing a good supplement to traditional condition-based
strategy.

1.1.4 Reliability Cost-Worth Consideration

Power system aims to provide reliable electricity to its customers, but more reliable
systems usually involve more financial investment and it is unachievable to design a
system with 100% reliability. Therefore, the level of reliability should depend on the
needs of the customers and the associated costs of providing the service. Power
system planners have always attempted to achieve a reasonable level of reliability at
an affordable cost. The cost-worth evaluation is also an important aspect in network
planning, operation and asset management.

The cost-worth evaluation depends on balancing the utilities’ cost and the profits
obtained from customers. Thus, the utilities’ cost and the customers’ interruption
costs must be considered in current operation and planning practices. As discussed
in the previous two sections, technologies for enhancing power system flexibility will
contribute to the improvement of reliability but also involve with a significant initial
investment in equipment. Under normal circumstances, a conservative asset
management strategy usually leads to better reliability performance, in addition to
higher asset maintenance and replacement costs. Therefore, a cost-worth evaluation
would be useful to compare the costs incurred from such approaches against the
additional benefits from improved reliability and reduced interruption costs.

33
Chapter 1: Introduction

Customer
interruption Total cost
cost
Cost

Utility cost

Ropt Reliability
Figure 1-5 Reliability cost-worth concept

A reliability cost-worth planning method compares the cost of unreliability against


the cost of providing additional reliability with aim to determine whether the system
reinforcements are economically justified. Figure 1-5 shows how the utility cost
(such as more expensive asset management, or investment in flexible solutions), the
operating cost (i.e. extra ageing cost, and additional power losses) and the customer
interruption cost (i.e. reduced reliability) are combined into the “Total cost”.

As shown in Figure 1-5, the utility cost (red line) increases significantly along with
the increase in reliability, due to the high marginal costs required to obtain higher
reliability levels. The cost–worth approach tries to find the minimum cost of the
‘Total cost’, which contains utility investment cost, system operating cost, and
demand interruption cost. The customers will receive service at the optimal
reliability, Ropt, when the total cost is minimised. The minimum cost is displayed as
the lowest point of the ‘Total cost’ curve, as indicated by a blue line in Figure 1-5.
Therefore, a specific reliability level is able to be determined considering utilities’
costs and customers’ worth when providing electricity service using various flexible
solutions and asset management strategies [13].

This thesis has used a cost-worth evaluation to compare the utility cost to enhance
network flexibility (i.e. investment cost, ageing cost) against the reliability gains (i.e.
reduction on customer interruption cost). This helps to provide utilities with more
economically competitive solutions to sustain the system’s reliability, and to
overcome emerging challenges such as renewable generation.

34
Chapter 1: Introduction

Cable thermal/current ratings depend on the cable design, load pattern, laying
condition and required system performance. Several standardised rating methods
for power cables (i.e. ER P17, IEC60287) are available on the market, as well as some
commercial items of software being capable of performing rating calculations for
different types of installation, such as CRATER.

Current power cable ratings can be divided into two categories; static rating and
dynamic rating. The static rating of a power cable is termed so due to there being a
constant current which can be carried continuously for an infinite period of time.
Dynamic cable rating is defined as a time-varying current which is dependent on the
variation of the surrounding thermal conditions. In this section, the existing cable
rating methods have been reviewed, including three static rating methods
(continuous rating, cyclic rating, probabilistic rating) and two dynamic rating
methods (emergency rating, dynamic rating).

1.2.1 Continuous Rating

Continuous rating, also known as normal rating, is defined as the maximum current
which will allow a cable to operate continuously without exceeding its normal
operating temperature. This represents the continuous, steady state load that the
cable can carry based on the given design with a negligible loss of life. Operating the
cable system at or below this level will not result in any degradation of the cable
insulation or shorten the nominal life of the cable system.

The continuous rating of UGCs is largely influenced by the ambient soil temperature,
thermal properties of the surrounding soil and laying conditions. This calculation is
based on either Neher-McGrath [14] or IEC 60287 [15, 16]. Typical environmental
parameters used for continuous rating calculation in UK have been stated in ENA ER
P17:

 Conductor temperature Limit: 85°C (paper cables), 90°C (XLPE cables)


 Soil temperature: 10°C (Winter), 15°C (Summer)

35
Chapter 1: Introduction

 Soil thermal resistivity: 1.2km/W (conservative), 0.9km/W (normal)

1.2.2 Cyclic Rating

Cyclic ratings are frequently used in distribution cables when an intrinsic load
pattern exists. This rating includes a current profile for the cable, which typically
varies in a time-repeating cycle on a daily basis. The cyclic rating will, in general,
allow for a higher loading of the cable, considering the continuously variable nature
of the network demand and loading while ensuring that the temperature is kept
below their limits. Thus, in the standard context, cyclic rating is defined as the
maximum current during daily load cycles, which the cable can sustain constantly
without exceeding its thermal limit.

This deployment assumes that the cable is not loaded at its continuous rating
constantly. Therefore it allows for a rating that is higher than the continuous rating
loaded on the cable during load cycles. This is due to the thermal inertia of the cable
surroundings. It takes a substantial amount of time for UGCs to heat and cool. When
the cyclic current is over the continuous rating for a short time (for example, a couple
of hours), the cable is still not hot enough to exceed its thermal limit.

The presently used rating standard for UGCs, ER P17, applies a 24-hour profile of
normalised hourly load values (the industry standard load curve G) in order to
produce the cyclic ratings [17]. Scottish power & energy networks (SPEN) use a cyclic
rating factor for their 11kV and 33kV cables which is calculated based on a typical
winter daily load curve. For low voltage cables, the factor is derived from a typical
domestic housing 24 hour load cycle. The cyclic ratings need to be modified
consequently if the actual load cycle is significantly different from the standard cycle
[18]. Electricity North West (ENWL) calculate their cyclic ratings based on a load
cycle of 8 hour at 1.0pu and 16 hour at 0.75pu loading. Under this operating scheme,
the cyclic rating factor is 1.09 for cable sizes up to 35mm2, and 1.13 for cable sizes
varying from 70mm² to 630mm² [19].

36
Chapter 1: Introduction

1.2.3 Probabilistic Rating

As discussed, continuous ratings are conservatively calculated based on the worst


case parameter values. However, the cable ratings are influenced by multiple thermal
parameters, including ground temperature, soil thermal resistivity and thermal
capacity. These factors are rarely constant and usually subject to large variations
between the cable routes. Thus, there is a great amount of headroom capacity unused
in the cables using traditional rating methods.

A probabilistic method chooses less conservative soil thermal parameters in order to


calculate the probabilistic rating. It calculates the probability of exceeding the design
temperature using historical soil parameters and a continuous loading at a designed
rating [20]. For example, a single conservative value of soil thermal resistivity
1.2km/W is recommended by industry standards and manufacturers, because this is
the ‘worst case’ for UK conditions. However, a thermal resistivity of 0.9 km/W is
considered to be more representative of ground conditions in most areas. A
probabilistic rating could be calculated to provide a higher current rating based on
0.9 km/W.

Cables are normally underrated. However, there is also a risk of operating above their
maximum allowable operating temperatures under unfavourable weather or thermal
conditions, such as long periods of dry weather and road crossings. Route checks and
the careful selection of appropriate soil thermal parameters is necessary for
probabilistic rating calculations in order to avoid unexpected overloading and
excessive ageing.

1.2.4 Emergency Rating

Emergency rating is the current that the cable is able to conduct for a specified period
of time before the limited maximum temperature is reached. After that, the current
must be lowered to the continuous rating level. While continuous ratings have no
limiting application time period, emergency ratings permit a higher current that
exceeds the continuous rating for a finite period of time when accelerated aging and
life loss in excess of the design criteria may occur.

37
Chapter 1: Introduction

Emergency ratings provide short-term flexibility to network circuits when a


contingency occurs in the network. It allows the cable to be loaded above its
continuous rating, which permits more interrupted demands to be restored during
emergencies.

One of key cable system design factors that could help to increase the short-term
emergency loadability is the thermal inertia of both the cable and the surrounding
environment. It considers that pre-contingency cable temperatures are usually below
the maximum temperature, and that it takes time for the temperature to rise from
the starting temperature to the maximum. This inertia allows the cables to be
temporarily loaded above their continuous rating before the cable conductor
temperature reaches the limit [21].

Another element of the key cable capability which could help to increase their
emergency loadability is the emergency loading temperature. For moderate periods
of temporary emergency overloads, almost all cables have the ability to tolerate an
emergency loading temperature, which is substantially higher than the normal
loading temperature at the expense of an admissible life loss.

Consequently, the cable’s capability to tolerate emergency loading temperatures and


cable resilience to rapid temperature changes provides them with the possibility to
overload the cables for short emergency loading periods. More interrupted
customers can be restored during contingencies and this improves overall network
reliability.

Transmission system operators (TSO) choose to differentiate emergency ratings


between a short duration, i.e. short-term emergency (STE) rating, and a long duration,
i.e. long-term emergency (LTE) rating [22, 23]. The LTE rating defines the maximum
cable loading that can be sustained by a cable for up to 4hrs after a contingency. The
STE rating is the maximum cable loading that can be carried by a cable for up to 15
mins after a contingency. Distribution network operators (DNO) also consider there
to be a 3-day emergency rating, more liberal than the cyclic rating, which may be used

38
Chapter 1: Introduction

in a loaded two-legged ring network. The emergency rating for the 11kV cables is
effectively 110% of the appropriate cyclic ratings [18, 19].

1.2.5 Dynamic Rating

Traditionally, cable ratings are determined using a static rating assuming a


conservative worst-case scenario for the weather and loading conditions. However,
the variation of weather and loading conditions provides opportunities to unlock
network capacities. Dynamic rating, sometime called real-time rating, calculates the
cable rating in real time according to the real-time weather measurement, soil
thermal properties, cable temperatures and cable loading.

The real-time cable temperatures become the key element for dynamic cable rating
calculations to ensure that the temperatures do not exceed the cable’s design limits
at all times. Distributed temperature sensing (DTS) instrumentation can provide
direct and continuous measurement of temperature profile along the cable which can
help to rate the cables up to their maximum ampacity. Therefore, DTS has become
the best tool to realise dynamic rating. However, considering the high cost of fibre
optic sensors and the feasibility of their installation, it is not feasible for all UGCs to
be equipped with DTS, especially concerning the existing (low and medium voltage)
underground distribution cables. Thus, indirect methods have been developed to
estimate the cable temperatures without a DTS system based on a thermal model of
the real-time measurements of the cable’s environmental conditions and real-time
loading [24].

The practicality of this ‘dynamic cable rating’ concept has been demonstrated and
attested within various government projects [24-26]and IEEE journal articles [27-
30]. Some dynamic line rating systems have been applied and installed by utilities
worldwide. For example, the MAXAMP system [31] and EPRI DTCR model [32]. Some
results show that dynamic cable ratings can effectively increase the cable current
carrying capacity by 10-20% without risking damage to the cables [33, 34]. However,
dynamic rating systems are difficult to implement and operate, which limits its wider
business applications.

39
Chapter 1: Introduction

The move towards a greener energy society has driven utilities to increase the
utilisation and efficiency of their infrastructure. However a large proportion of the
power equipment’s capacity is utilised only when there is a combination of rare
events and emergency conditions (failure of components and peak load times). The
wasted margin of circuit capacity is reserved for emergency operations in order to
achieve a satisfied level of network reliability. This has a further consequence with
the increasing utilities’ need to accommodate more distributed generations (DG),
electric vehicles (EV) and wind generation. Reinforcing the network infrastructure
without improving its utilisation will lead to a large amount of non-utilised network
investment. Therefore, how to increase the utilisation of the existing infrastructure
while maintaining relatively high network reliability has become a vital problem to
tackle.

Since there are a large number of cables existing in transmission and distribution
networks, optimising their operation and loadability will enable the utilities to reach
that target. At present, most UGC systems utilise the static rating method, which
brings in two weaknesses. First, static rating assumes there to be a constant loading,
neglecting the fact that the cables are normally not loaded at a constant current for
long periods of time. Second, static rating utilises conservative static weather factors,
such as soil resistivity and soil temperature, neglecting the real-time variation of
environmental conditions.

Some methods have been developed to overcome cable rating constraints, such as
probabilistic rating, emergency rating, and dynamic rating. As discussed in the
previous section, dynamic cable rating (DCR) optimises the utilisation of cables while
ensuring that their thermal limitations are never exceeded. There are only a few risk
related to overloading and ageing for DCR, therefore it can be used in long-term
operations. However, DCR requires complex equipment and time-consuming
constructions which prohibit its implementation in most existing cable circuits.

Emergency ratings provide a way to unlock short-term emergency loadability,


originating from cable thermal inertia, in order to increase the utilisation of cables

40
Chapter 1: Introduction

and to improve overall network reliability. However, it is difficult to determine


whether appropriate emergency and probabilistic ratings can be defined, and to
know whether or not the cable is being either operated at significantly below its
capacity or being overloaded excessively, which can cause premature ageing and
failure.

Currently, some researchers have investigated the temporary overloading capability


of power cables [35, 36]. Only very little research has considered cable emergency
rating in network operations [37, 38]. Almost no research has been done to evaluate
the values and risks of the cable uprating methods on network reliability and cable
ageing. To fill in this research gap, this thesis integrated the cable thermal-ageing
modelling into a network-wide reliability evaluation framework, which allowed for
the evaluation of cable ageing risks using different cable uprating methods.

Power system reliability is a traditional and historic subject, which has been studied
since the 1970s. The development of smart grid technologies requires modern
reliability studies that incorporate the operating aspects of the technologies such as
distributed generation, electric vehicles, HVDC, FACTS, dynamic line rating, storage
and demand responses. The existing research has focused on investigating the impact
of the emerging technologies on the network reliability performance and their
optimised operation schemes. The impact of their utilisation in relation to the
existing network assets is still undefined and not been studied extensively.

For more comprehensive asset management strategies, real-time network


information and specific component design characteristics (materials and operating
temperatures) should be integrated to estimate the risk of individual component
ageing. For this reason, cable thermal modelling and ageing calculations are
integrated into a network-wide evaluation. Some of the relevant past research on
power system reliability, cable thermal modelling and cable ageing has been stated
in the following sections.

41
Chapter 1: Introduction

1.4.1 Power System Reliability

Currently, there have been a lot of reliability studies investigating the impact of
renewable and distributed generation [39-41], electric vehicles [42, 43] and dynamic
line rating [30, 44, 45] on overall power system reliability. However, almost no work
has considered the impact of such technologies on existing network assets, such as
ageing risks, and new operating and maintenance strategies related to the critical
network assets.

There have been a few existing reliability studies incorporating the ageing
component and ageing failure into the network reliability assessment [46, 47]. Both
approaches utilise probabilistic functions to model the unavailability of the
components’ ageing failures. However, the actual ageing of each individual
component in the network cannot be accurately estimated. The existing approaches
do not consider the actual operating aspects, design and environmental properties.
Thus, they are not able to quantify component ageing due to the temporary
overloading operations and the risk of increased flexibility provided by the smart
technologies.

1.4.2 Cable Thermal Modelling

Underground Cables (UGCs) have a significantly delayed temperature elevation after


the increase of the current level due to the thermal inertia of the cable and its
surroundings. This thermal inertia is frequently neglected in steady-state thermal
modelling, resulting in lower cable temperatures, particularly under emergency
loading. Transient thermal modelling considers the thermal inertia of the cable
systems and therefore allows us to more accurately estimate the cable operational
thermal conditions.

Many UGC thermal modelling methods have been developed to simulate and estimate
the cable conductor temperature, based on finite element analyses [28, 48] and
thermo-electric equivalent circuits [37, 38, 49]. The Finite Element Model (FEM) is
deemed to be more accurate since it considers the specific geometric and material
properties. However, it requires a high computational times. The Thermo-Electric
Equivalents (TEE) model is based on current measurements and cable lumped

42
Chapter 1: Introduction

parameters, considering the thermal model of the cable and surrounding soil as an
analog of the electric circuit. The TEE model provides a sufficiently accurate method
to estimate conductor temperature for both steady and transient state analyses
under normal and emergency operations. The accuracy of the TEE method is proved
by the comparison with FEM method in many existing literatures [37, 38, 49].

However, relatively little work has incorporated the TEE thermal modelling of cables
within network-wide studies [11, 38, 50]. However, the existing studies have not
considered the thermal inertia of UGCs in the thermal modelling, which is the most
dominant design factors, particularly when under short-term emergency loading or
intermittent renewable output. More importantly, they do not take the thermal
ageing of the cable into account.

1.4.3 Cable Ageing

Most of cables can tolerate emergency loading temperatures that are substantially
higher than normal loading temperatures for moderate periods [51]. However,
operating at emergency loading temperatures leads to thermal ageing and life loss
within the cables. It is indicated that electro-mechanical failures occur under
emergency loading conditions, caused by the increased current and the
corresponding increase in the heat generation rate and temperature of the cable [52].
The failure statistic indicates that most cable failures are caused by external
excavation activities, but 60% of internal failures in cable systems are caused by the
ageing of the cable insulation materials [53].

The Arrhenius model is the most well-known model used to describe the impact of
temperature on the rate of thermally-initiated chemical reactions. This method is
widely applied related to the ageing of the insulation materials of cables. The
Arrhenius model indicates that cable ageing increases exponentially when the
conductor temperature becomes higher than the normal operating temperature. An
electro-thermal model used to estimate the lifetime of HV cables has been proposed
based on the Arrhenius model, considering the load cycle and voltage variation [21,
54].

43
Chapter 1: Introduction

Most of the previous studies have investigated thermal ageing behaviour and lifetime
estimations for a single cable [21, 55], and only a few studies have incorporated cable
ageing and the Arrhenius model into network reliability analyses [56]. Most of these
studies do not consider the effects of the risk of ageing and the failures caused by the
emergency loading operations of cables into network reliability evaluations.

1.4.4 Summary of Past Research

After reviewing the past research, several research gaps require being addressed:

 Only a limited amount of studies consider detailed plant properties and


modelling in network-wide evaluations. In particular, the existing research
concerning cable thermal and ageing models has focused only on individual
cables, rather than cables in a network.

 The transient thermal modelling of UGCs has not been integrated into network-
wide studies, thus transient temperature changes under emergency or time-
varying loadings have not been accurately captured in network studies before.

 An assessment into the effects of flexible cable rating methods, i.e. emergency,
probabilistic and dynamic ratings on cable thermal behaviour and ageing
performance should be completed. Previous research focused only on their
impact on network aspects, rather than on equipment aspects.

 Tools and metrics are required to quantify the cable thermal performance and
ageing under various network operation schemes with the integration of smart
grid technologies and renewable energy sources.

 A comprehensive asset management strategy that considers the actual network


operating aspects, plant design properties and environmental conditions is
currently lacking. Existing asset management is mainly based on on-site
condition measurements, neglecting the operating history and environmental
conditions.

44
Chapter 1: Introduction

 Limited studies have been conducted to investigate the potential headroom


loading capability of underground cables when transmitting time-varying
renewable outputs and their associated ageing and overloading risks.

The restructuring of the electricity sector forces utilities to operate their existing
power networks with increasing flexibility and efficiency. Flexible cable rating
methods and smart grid technologies provide additional network flexibility and
resilience to current operating uncertainties. However, such approaches may pose
more stress on the network’s aged assets, thus it is important to identify critical
network assets and to develop new operating and maintenance strategies for them.

The main aims of the research are to provide operators with additional tools and
metrics to identify key network cables based on the quantified ageing index in order
to perform optimised network replacement and maintenance strategies. Another aim
is to propose alternative cable operating practices to increase the cable-based
network’s loadability and flexibility in order to improve the overall network
performance but also to avoid the excessive risk of cable ageing. The output of this
research will help to improve the utilisation of existing cable networks, reduce the
investment costs of earlier replacements and reinforcements, and increase the
network flexibility and resilience against long-term uncertainties.

To accomplish these aims, the following objectives have been determined:

1. To summarise the existing approaches for power network reliability


assessments, cable thermal and ageing modelling.

2. To develop the sequential Monte Carlo (SMC) simulation models within Matlab
and Matpower to use to evaluate power transmission and distribution network
reliability.

3. To develop and validate the steady-state and transient-state thermal modelling


of the underground cables based on the thermo-electric equivalent circuit, and
capturing the risk of cable ageing using the Arrhenius equation.

45
Chapter 1: Introduction

4. To propose a methodological framework integrating a SMC-based reliability


evaluation with a cable thermal-ageing model considering key cable design
factors.

5. To recognise the network’s critical cables using a novel ageing index (EECA),
and allowing utilities to make more comprehensively informed decisions on
optimal cable maintenance and replacement.

6. To quantify the additional failure and ageing risk of UGCs at emergencies within
a multiyear network analysis using a double SMC simulation loop.

7. To quantify the network benefits and potential risk of ageing and cascading
failures when extensive emergency ratings (i.e. extended duration and
increased rating) are implemented in the transmission circuits.

8. To propose a flexible current rating (FCR) method that can be implemented


without complex equipment installation to optimise the loading capability of
existing cable tie lines between wind farms and host utilities.

9. To propose emergency cable rating solutions to enable the higher penetration


of electric vehicles (EVs) within distribution all cable networks.

The main outcome of this research was to advance the existing methodology used for
evaluating system reliability by integrating a SMCS with a detailed UGC system
thermal-ageing model that considers cable design factors and environmental
conditions. This integration enables the accurate estimation of the cables’ thermal
behaviour and ageing risk. This also allows the utilities to quantify the benefits drawn
from the different operation schemes and technologies related to network
performance, including adequacy, flexibility and reliability, against the increased risk
of cable ageing and the additional costs of early replacement and frequent
maintenance.

A list of journal and conference publications has been included in Appendix C. The
main contributions of this research are summarised as following bullet points:

46
Chapter 1: Introduction

 The integration of cable thermal-ageing models into a SMC-based network


reliability evaluation. The methodological framework combines the cable TEE
thermal model and Arrhenius ageing model into a sequential Monte Carlo
simulation. Thus, the framework integrates cable design and environmental
conditions into a cable transient thermal-ageing modelling, which enables an
accurate estimation of the cables’ thermal profiles and ageing under time-
varying loadings. The methodology can inform the utilities of the more accurate
estimation of cable ageing and health condition, thus allowing them to provide
suggestions on the optimised cable replacement and maintenance strategies.

 The development of a high-loading failure model for cables operating


above normal continuous rating. This methodology considers the increased
failure and ageing risk of UGCs at emergency loading operations. This approach
allows for the evaluation of the network performances, at the same time,
considers the negative impact of the increased ageing and failure on network.

 An investigation into the effects of an extensive emergency rating of the


transmission circuits on network performance and plant ageing. This
approach considers the extended emergency rating operating duration and
increased emergency rating ampacities to be implemented, at the cost of the
increased risks of ageing and cascading failure. According to the results, utilities
are suggested to change the existing emergency rating operating manual of
transmission lines and cables for optimised network operations.

 Novel flexible current rating approaches have been proposed to unlock


the additional capacity available in wind farm cable circuits for extra wind
integration. The proposed flexible current rating methods can be implemented
without complex equipment in relation to dynamic rating, in order to optimise
the loading capability of existing cable tie lines between wind farms and their
host utilities. This approach allows wind farm operators to increase their
generation capacity in the future without the need to reinforce their existing
cable circuits.

 The use of the emergency loadability of UGCs is proposed to enable a


higher penetration of electric vehicles in distribution networks. The

47
Chapter 1: Introduction

proposed remote switching method utilises automated remote switches to


connect/disconnect aggregated EV charging parks in order to achieve an
optimised network operation with minimum overall costs. Emergency uprating
provides additional loadability to the restructured distribution network at the
cost of increased cable ageing. Both approaches provide DNOs with more
flexibility to control and operate their networks under contingencies, and
maintenance allows for a relatively high reliability and low ageing when a high
penetration of EVs is integrated.

Seven chapters in total are included in this thesis. The following six chapters have
been outlined below.

Chapter 2 - Reliability Evaluation Framework Integrating Cable Design and


Ageing

This chapter introduces a reliability evaluation framework considering cable design


and ageing, which is the basis of the methodology used within this thesis. It presents
the methodological steps of how the cable thermal-ageing modelling was integrated
into a sequential Monte Carlo simulation to use for evaluating the network’s
reliability. The process of state sampling using SMC has also been presented and
described using mathematical equations, flowcharts and demonstration figures. It
further presents the objective function for use related to optimised power flow
calculations. Following this, the thermo-electric equivalent (TEE) circuit model for
cables under different installation conditions has been described in detail, which is
available for steady-state and transient-state calculations. Several existing
transmission and distribution network reliability indices, together with new
proposed cable performance indices, have been introduced.

Chapter 3 - Impact of the Cables’ High-loading Risks on Network Reliability

This chapter introduces an advancement in the methodological framework


(introduced in Chapter 2), which accounts for the increased cable failure risks
because of their emergency operations. A double sequential Monte Carlo (DSMC)
simulation loop is implemented to capture cable ageing risks during emergencies

48
Chapter 1: Introduction

within a multiyear evaluation. The methodology was applied on the IEEE 14-bus test
network with long/short term emergency ratings using steady-state and transient-
state thermal modelling. The methodology proposed within this thesis enables for
the evaluation of increased network flexibility and adequacy in the context of
emergency ratings and emergency loading risks, including high-loading failures and
cable ageing.

Chapter 4 - Impact of Extensive Emergency Ratings on Network Reliability and


Plant Ageing

This chapter evaluates the impact of the extensive emergency ratings of transmission
circuits on network performance and thermal ageing of circuits, which considers
extended durations and increased ampacities for long- and short-term emergency
ratings. It further presents a fault chain search model to capture the cascading
failures of circuits due to severe overload. The methodology was demonstrated on
both 14-bus and 24-bus networks with all of the circuits being overhead lines or
underground cables. Consequently, the results from the case studies indicate the
network benefits and ageing risks when extensive emergency ratings are
implemented. The difference between OHL networks (small thermal inertia) and UGC
networks (large thermal mass) have also been demonstrated.

Chapter 5 - Increasing Wind Power Integration through Flexible Cable Current


Ratings

This chapter proposes alternative flexible thermal rating approaches that can be
implemented on existing tie cables between wind farms and host utilities. Most
importantly, it further quantifies these flexible methods’ capability in wind
integration output and thermal ageing risk against the standard rating practices
(static, probabilistic, seasonal and dynamic cable rating). Such flexible rating
practices are more straightforward and easy-to-implement which enables
implementations for most existing wind farms. The case study results provide
recommendations on changing the existing asset rating policy to consider alternative
rating approaches for sizing the wind farm cable circuits.

49
Chapter 1: Introduction

Chapter 6 – Quantifying the Risk of Emergency Ratings on a Distribution Cable


Network with EVs

This chapter presents a methodology to use to assess the impact of electric vehicle
integration in distribution networks on the thermal behaviour and ageing
performance of distribution cables and their potential to accept an increased load
from EVs. It further presents the modelling for the electric vehicle charging and
restoration processes, and constructs new load profiles by adding EV charging
profiles to regular household loads. This study also proposed a novel method using
remote switching and emergency ratings to reduce the impact of EV charging demand,
reinforcing the reliability of the grid, thus enabling a higher penetration of EVs into
the existing distribution networks. The methodology was applied on a EV-integrated
RBTS bus 4 distribution network implementing different switching schemes and
cable ratings.

Chapter 7 - Conclusion & Future Work

The core conclusion of conducted studies was discussed in this chapter. Additional
suggestions were also given for future work and to indicate potential improvements
drawn from the existing methodologies.

Equation Chapter (Next) Section 1

50
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

2 Reliability Evaluation Framework


Integrating Cable Design and Ageing

This chapter introduces an innovative methodological framework to evaluate


network reliability, which combines a thermal modelling of UGC systems into a
sequential Monte Carlo simulation. This integration captures an estimated
conductor temperature and the ageing risk for each individual cable in the network.
The well-known Arrhenius model was utilised to calculate the thermal ageing of
UGCs. The cable conductor temperature was computed using a thermo-electric
equivalents circuit which including the design properties of cable itself and its
ducting system, as well as the environmental soil properties. In this way, the critical
cables within the evaluated network can be identified according to the equivalent
ageing index.

Consequently, the proposed framework can evaluate the impact of increased


flexibility on system reliability improvements and associated cable ageing risks.
Increased network flexibility can be obtained by implementing flexible cable rating
methods, improved operation strategies and other smart grid technologies. This
methodology, together with the existing method to model OHL ageing, will give the
utilities advisable suggestions on optimal replacement and maintenance strategies
for their circuits. More importantly, this holistic methodology allow the utilities to
evaluate the benefits and risks of such approaches based on the network
performance, individual circuit ageing condition and overall network cost-worth
analysis.

This framework was utilised as a fundamental methodological framework that will


be applied in several specific studies in the next four chapters of this thesis, while
some methodological advancements have been included for future studies in order
to address their specific research problems and to fill in the research gap.

51
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

Utilities are keen to investigate economically competitive solutions that can be


applied to optimise their assets’ efficiency and to increase their networks’ flexibility.
This allows for their existing network to become more resilient to emergency
contingencies and more flexible, maximising the amount of renewable energy in
service. Flexible rating solutions intend to improve the circuits’ loadability by
implementing emergency, probabilistic, dynamic, flexible rating and other smart grid
technologies.

With respect to underground cables, the increase in power transfer adequacies can
be achieved by relaxing the electrical constraints (current rating) or operating
temperature constraints (thermal rating) [57]. Although such flexible methods
increase the power transfer capabilities of the existing network, the additional risk
of thermal ageing will occur for as long as they are not captured within the existing
network evaluation modelling.

Quantifying the ageing risk of those flexible approaches and other smart grid
solutions is essential with aim to select inexpensive, reliable and less risky solutions.
The implementation of more ‘expensive’ solutions can usually increase overall
network performance while maintaining an acceptable level of ageing risk, such as
DTS-DCR monitoring equipment [58, 59]. However, the ‘expensive’ solutions require
a significant amount of initial investment in new equipment or construction, and so
its benefits might not be worth it financially and technically. In order to assess the
benefits of those flexible solutions, a network reliability evaluation is conducted to
quantify the potential ageing risks of cables and their impact on the overall network
performance.

Existing approaches used in the network reliability assessment apply either a


deterministic or probabilistic tool [3, 60]. Deterministic approaches do not consider
the stochastic behaviour of component failures, while a probabilistic evaluation is
much more versatile when modelling random aspects. Among the different
probabilistic methods, a sequential Monte Carlo simulation can actually model time-
varying power sources, time-varying loads and time-varying circuit ratings in a time

52
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

sequence with the information captured in the system state transitions. However,
relatively little work has an integrated cable design and ageing built into the
reliability evaluation framework [11, 61, 62]. Thus, they cannot be used to evaluate
the benefits and risks of flexible rating solutions.

The conductor temperature of UGCs can be modelled and calculated based on several
existing methods. The finite element model considers the properties of cable material
and surrounding environment, thus provides more accurate results. On the other
hand, a lot of computation time is required [28]. The Thermo-Electric Equivalents
(TEE) method simulates the thermal modelling of the cable and its surrounding soil
as an analog of a electric circuit [37, 49]. Then, it estimates the cable conductor
temperature using measured cable current loading and the TEE’s lumped parameters.
when a normal or emergency loading is applied, the TEE method provides an
accurate result under both steady-state and transient-state thermal calculations [37,
38, 63]. However, only few research has integrated the TEE modelling of UGCs into
the network reliability evaluations [11, 38, 50]. More importantly, the majority of
existing work only consider steady-state thermal calculation, neglecting thermal
inertia of cable itself and its surrounding environment, which is one of the most
critical factors

The Arrhenius model is a very famous method used to assess different temperatures’
effect on the rate of thermally initiated chemical reaction [56]. Cable thermal ageing
shows an exponential increase along with its conductor temperature becomes higher
[64]. Although the Arrhenius model has been widely implemented in the ageing
calculation of cable insulation [21, 54, 55], it has not been integrated into the
sequential MCS nor any other network analysis.

This chapter introduces an innovative advancement in the existing network


reliability methodology, which integrates a thermal modelling of UGC systems into a
sequential Monte Carlo simulation. The estimated conductor temperature and the
ageing risk for each individual network cable was calculated based on a thermal-
ageing model of UGCs while combining the Thermo-Electric Equivalents (TEE) model
and the well-known Arrhenius ageing model. It incorporated the cables and their

53
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

ducting system design properties, as well as the surrounding soil properties, in order
to obtain an accurate estimation of the cable conductor operating temperatures.
Consequently, this methodology allowed us to evaluate whether it is worth
increasing the network flexibility to achieve improved network performance, while
additional ageing was also produced.

Section 2.2 describes an overview of the proposed framework, and the integration
between cable thermal-ageing model and SMCS. Section 2.3 describes the modelling
process of state sampling and transition using SMCS. Section 2.4 presents the
equations and constraints within the optimal power flow within the network
optimisation calculation. Section 2.5 and 2.6 demonstrates the thermal-ageing
modelling for UGCs and OHLs, and their integration in the framework. Section 2.7
introduces some existing and novel indices to use to evaluate transmission,
distribution and branch performance.

To evaluate the overall network performance and to capture the ageing risk of the
circuits under the implementation of flexible solutions, the proposed methodological
framework integrated thermal-ageing models of OHLs and UGCs into a SMC-based
reliability evaluation. The proposed framework has been outlined in Figure 2-1.

The time step, Δt, was appointed as the minimum transition time for component state
transition, weather changing and load cycling. The size of the time step will seriously
affect the simulation speed and the output accuracy. Usually, an hourly analysis is
performed for each time step, thus the size of the time step was chosen to be one hour
Δt=1hr. A smaller time step can also be chosen, if required in different studies, such
as Δt=15mins or 30mins.

In the ‘Power System Data Initialisation’ block, four types of data representing the
characteristics of the network and the plant were input. Detailed descriptions for the
input data have been listed below.

54
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

Network PLANT
Reliability Data Operational Data Weather Data Branch Data

Power System Data Initialization

Initial Network Component Status Mapping

Annual Network Optimization Time step


analysis analysis
〈Y=Y+1〉 Thermal-Ageing Modelling of OHLs and UGCs 〈tΔt=tΔt+Δt〉

NO
Annual Computation is completed
Inner SMC-loop
YES
NO
Analysis is completed at specified Cov Outer SMC-loop
YES
Network Reliability Indices and Branch Performance Indices Calculation

Completed Network Reliability Evaluation


Figure 2-1 Flowchart of the network reliability evaluation methodology

 Reliability data presents the failure rates λx and repair rates µx for every
component, x, according to standard annual performance in the history, ∀x∈Nx,
where Nx is the number of network components. Those data can be derived from
annual failure frequency and average repair durations of each component existed
in the utilities’ database.

 Operational data specifies the network operating limits, including the generator
maximum power output, generator operating cost function, transformer
maximum ratings and impedances, synchronous condenser operating limits,
annual load point chronological demand, each circuits’ maximum
normal/emergency ratings, impedances, and their emergency operation manuals.
This information can also be obtained from the utilities.

 For UGCs, the weather data meant the soil data, including time-varying soil
temperature θSoil(tΔt), soil thermal resistivity ρSoil(tΔt), and volume specific heat
CSoil(tΔt). For OHLs, weather data described the hourly wind velocity Vwind(tΔt),
wind incidence angles, ambient temperature θAmb(tΔt) and solar radiation. The

55
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

information was available from the measurement devices, or nearby weather


stations.

 For UGCs, the branch data represents the cable type, size, design, impedances,
maximum normal and emergency temperature, ducting and laying configuration,
such as if it was directly buried in soil or in ducts, trefoil or flat formation, and the
laying depth. For OHLs, branch data indicates the size, maximum strength, and
allowable temperature of the conductor, together with its structure properties.
All of this information was provided by the asset owners and manufacturers. This
additional data blocked together with weather data was utilised in the thermal-
ageing model of UGCs and OHLs in order to calculate the ageing due to emergency
loading events.

As shown in Figure 2-1, the simulation was performed though two dependent
computational iterations. The inner loop was performed in a sequential time vector,
tΔt, segmented in steps, Δt, ∀Δt∈Y, in every simulation year Y, ∀Y∈{1,⋯, NY}. In each
step, the simulation determines whether any changes occurred in both the plant and
network status because of the line rating changes, component failures, weather
changing and demand cycling. The outer loop executes the annual computation of
reliability indices. When a new simulation year was initiated, a brand new operation
and repair transition mapping of each component is produced. The entire simulation
was terminated when the covariance of EENS achieved a maximum boundary of 5%
or when reaching the maximum simulation years, N [3].

In the block ‘Initial Network Component Status Mapping’, the status transition
mapping of each component was built based on an SMC algorithm. A two-state
Markov chain was applied to determine the component status between the loading
(operating) and repairing (failure) states. The detailed description of the two-state
Markov model was given in Section 2.3. This block only captured the random failures
of the components based on their reliability data, while the weather changes, demand
changes and emergency operations may lead to additional failures, such as
immediate cascading failures and prospective ageing failures. This possibility of
additional failure was not considered in the framework, but will be captured within
studies in the following chapters.

56
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

The determined initial component status mapping was input to the ‘Network
Optimisation’ block. Then, optimised generator re-dispatching and load shedding
actions were determined by executing an AC optimal power flow (OPF) in order to
minimise the total operational cost in the network. The objective function and its
constraints for the network optimisation have been detailed in Section 2.4. The AC
OPF was achieved using the Matpower tool and provided detailed network
operational results, such as the current flow Iflow,x(tΔt), the operating voltage Vx(tΔt),of
each component x, the power output Gg, from each generator g, and the actual
supplied load Dlpactual of load point lp, at time tΔt. The network operational results
provided the initial input data for the thermal-ageing calculations and network
reliability indices computation.

The operational results were then fed into the ‘Thermal-ageing Modelling of UGCs and
OHLs’ block to calculated the conductor temperature in each Δt, θC,b(tΔt) of each
branch, b, and the corresponding ageing risk, ∀b∈Nb. The thermal-ageing modelling
has been described in detail in Section 2.5 and 2.6 for UGCs and OHLs respectively.

At the end of each simulation year, the results obtained from the network
optimisation and thermal-ageing calculation were implemented for the indices
calculation. All indices are grouped into two types: branch performance indices and
network performance indices. The mathematical equations have been stated in
Section 2.7.

The block computes a transition status of every network component, x, and thus a
specific system state for every time step exists in a complete simulation year, Y, under
normal operating condition. A two-state Markov chain is commonly applied to
determine the transition between normal-loading and normal-failure state of each
network component, x. The process of state sampling and mapping has been
explained using Figure 2-2. This process helps to build the initial network component
status mapping for NY years at the beginning of the SMC simulation in Figure 2-1.

57
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

As shown in Figure 2-2, for the initialisation, the reliability variables (failure and
repair rates) of each component x have to be specified. Additionally, the number of
network components has to be set, which can vary from 10 to 1000 depending on the
complexity of the network. The size of the time step Δt needs to be defined as per the
requirements of this specific study. For a wind farm example, the time step is selected
to be 15 mins as the wind speed data is collected and recorded in 5-15 mins. However
for a regular bulk T&D network analysis, the time step is usually defined as 1 hour
for a faster speed.

Define the number of network component Nx, the size of time step Δt

Initialize simulation year Y=1, time vector tΔt=0, component number nx=1

Generate a new random number UN

Compute TTFN,x using λN,x and UN for component x

Define ‘x’ as working component for the next TTFN,x period

tΔt=tΔt+TTFN,x

Compute a random TTRN,x using µN,x and a new random


number UN for component x

Define ‘x’ as failed component for the next TTRN,x period

Next year tΔt=tΔt+TTRN,x


(Y=Y+1)
NO
Reach the end of simulation year Y, tΔt < Y
YES
NO Next component
All component computed, nx = Nx
(nx=nx+1)
YES
NO
Required simulation year achieved, Y = NY
YES
Combining the operating cycles of all components to
determine system state in each time step
Figure 2-2 Flow Chart for Initial Network Component Status Mapping

In next step, the failing time and repair time of each component was calculated
assuming an exponential distribution of the state duration. According to the failure

58
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

rate λN,x and the repair rate μN,x used for normal operating conditions, the time-to-fail
(TTFN,x) and time-to-repair (TTRN,x) was calculated in (2.1) and (2.2), using the
component unavailability UN,x which is a random number (0< UN,x <1).

 1 
TTFN , x   
   (2.1)
ln U N , x
 N ,x 
 1 
TTRN , x   
   (2.2)
ln U N , x
 N , x 

Then, based on the TTFN,x and TTRN,x for each component x, the operational state can
be mapped in the simulation year. All of the components’ states begin from the
operating state. After a complete TTFN,x, they went into a repair state for a duration
of TTRN,x, and then went back into the operating state and continued with the same
iteration until the end of the simulation year. The computation then continued with
the next network component until all of the components were mapped. Consequently,
in each time step Δt, the status for each component was obtained.

State transition mapping for component x=1, 2, 3 and 4


Δt Δt Δt Δt Δt Δt
TTFN,1 TTFN,1
Loading
x=1
Failure TTRN,1 TTRN,1

TTFN,2 TTFN,2 TTFN,2


Loading
x=2
TTRN,2 TTRN,2 TTRN,2
Failure

TTFN,3 TTFN,3
Loading
x=3
Failure TTRN,3

Loading
TTFN,4 TTFN,4
x=4
Failure TTRN,4 TTRN,4

Time axis tΔt


tΔt=0 tΔt=Y
Figure 2-3 an example of modelling the component state transition

Figure 2-3 illustrates a state transition mapping with four components that transition
between loading and failure state. It can be observed that for every simulation year,

59
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

each component has a different TTFN,x and TTRN,x due to their own λN,x, μN,x and
different random number UN. The component status mapping at each specific time
step determines the system state in that time. For example, in the marked time step,
component x=2, and 3 are under normal-failure state, thus the network is considered
as under ‘contingency’.

When the network is considered as being under ‘contingency’ (when one or more
components fail), network emergency control actions might be required to protect
any overloaded circuits and maintain the integrity of the network, such as load
shedding, adjusting the generator output, line tripping and emergency overloading.
Additionally, flexible methods might also be useful to provide increased flexibility for
the network in order to reduce load shedding and thus to improve the overall
network performance. Consequently, the initial component status mapping produced
totally random component failures and system states throughout the simulation year.
This provided the basis for the reliability studies, and allows us to evaluate the impact
of the flexible methods under the network contingency conditions.

Apart from the operating state mapping for each component, the load demands and
weather conditions will also be mapped for each time step in the simulation year
using the input operational data and weather data. The combination of system state,
load demands and weather conditions allows for further network optimisation
calculations and thermal-ageing calculations at each time step.

The network optimisation implements an ACOPF to minimise the total generation


and load curtailment costs occurred in the network. The objective function for ACOPF
is stated in (2.3) including some system constraints, such as generator output
constraints Gg, bus voltage constraints Vbus and branch loading constraints Sb.

 Nlp lp lp 
   
Ng
Min   f D  Dactual  Dtotal    f Pg  pg  +f Qg  qg  
lp
(2.3)
 lp g 
subject to: -Smax, b  Sb  Smax, b , G min, g  G g  G max, g , Vmin, bus  Vbus  Vmax, bus

60
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

All of the loads were simulated to be dispatchable, modelled by negative generators


using negative power injections and negative costs (benefits). Their operational cost
were defined to be larger than any network generator. Therefore, the calculation first
g
re-dispatched all of the actual generators g, considering their cost functions, f P for
g
the real power output pg, and f Q for the reactive power output qg. Then, as a second
choice, it re-dispatched the load (negative) generators. This procedure tends to re-
dispatch actual generators first and then load generators, with aim to minimise the
load curtailment and the curtailment costs. The amount of load curtailment at each
lp
load points, lp, was computed as the total required load, Dtotal minus the actual
lp
supplied load, Dactual . The costs of load curtailment were also minimised by
lp
considering the amount of load curtailment and the demand cost function, f D , as
shown in the first part of (2.3).

The thermal-ageing modelling of UGCs converted the current flow, Iflow,i(tΔt), of each
cable i, ∀i∈{1,⋯, NUGC}, at Δt, to the conductor temperature θC,i(tΔt). Then, it calculated
the ageing that occurred within the Δt. Figure 2-4 indicates the process to compute
the conductor temperature in steady-state, SS-θC,i(tΔt), and transient-state, TS-θC,i(tΔt).
The transient TEE time domain, tθ, was used to obtain the θC,i(tΔt) at the end of a given
Δt. The steady-state calculations applied tθ=∞, while the transient-state calculations
considered tθ=Δt. The dashed line indicates the simulation moves to the next time
step Δt (tΔt=tΔt+Δt) after the time step analysis is completed.

Current loading at present


Δt for each cable i, Iflow,i(tΔt) CABLE SYSTEM DATA
Calculate TEE model properties for every x at present Δt loading
〈i.e., WC,i(tΔt), WS,i(tΔt), CC,i, TIns,i, Wd,i, …〉 [IEC 60287]

TEE model is developed for tθ [Figure 2-5 ]


〈i.e., WC,i(tθ), WS,i(tθ), θC,i(tθ), θAr(tθ), …〉

SS thermal calculations TS thermal calculations


〈 θC,i(tθ=∞) 〉 〈 θC,i(tθ=Δt) 〉

SS-θC,i(tΔt) and TS-θC,i(tΔt)


Next Δt 〈tΔt=tΔt+Δt〉
Figure 2-4 Cable thermal calculation flowchart at Δt [62]

61
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

The cable system data, and the current flows, Iflow,i(tΔt), of each cable is utilised for
TEE calculations. The cable system data includes the cable design and material
properties, such as type (e.g., paper or XLPE), size (area in mm2), phases (single or
three cores), internal structure (with or without armour), soil properties
(temperature, thermal resistivity and volume specific heat), laying conditions
(formation, depth…). The metal sheath is used to provide an earth path for fault
currents, circulating currents and dielectric charging currents in the cable.

Thus, the input data provide the basis to establish TEE circuit of each cable. An
example TEE circuit of a single core cable with its ducting and surrounding soil layers
has been established in Figure 2-5 with all of the parameters calculated based on IEC
60287 [15, 16].

C ,i (t ) S ,i (t ) Ar ,i (t ) Air ,i (t ) SL1,i (t ) Soil ,i (tt )
TIns TArm TJ Tair Tduct TSL1 2 TSL1 2
WC Wd Wd WS WAr
2 2
CJ Cduct CSL1 CSL(n )
CC Cd Cd CS CAr
2 2
Conductor Insulation Screen & Armor Jacket Air & Duct Soil Layers 1...n
Figure 2-5 TEE model for a single core cable with n soil layers.

The surrounding soil is sub-divided into multiple layers using an exponential


discretisation approach [49, 65] in order to increase the simulation accuracy. In that
way, a T-equivalent circuit with its thermal resistance Tsoil and thermal capacitance
Csoil is formed for each layer. The exponential discretisation for the surrounding soil
was expressed in (2.4) and (2.5) where bk is the radial position of the layer borders,
rc is the radius of the cable, n is the number of layers, dm is the depth of the soil layer
model, L is the laying depth of the cable, k is 0, 1, …, Nk, and γ is the argument of the
exponential distribution.

e k  1
bk  rc  (d m  rc ) (2.4)
e n  1

62
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

dm  L  L2  rc2 (2.5)

In the equivalent electrical circuit, electrical symbols represent thermal symbols, and
current generators are considered as heat sources. In Figure 2-5, one layer is
allocated to conductor, insulation, metal screen and outer jacket individually
Therefore, Tins, TArm and TJ represent the thermal resistance of the insulation, armour
and jacket respectively. Cc, Cd, Cs, CAr and CJ represent the thermal capacitances of the
conductor, insulation, screen, armour and jacket respectively. Heat sources Wc, Wd,
Ws represent the losses in the conductor, dielectric and screen.

2.5.1 UGC Steady-state Thermal Modelling

The steady-state thermal model was established by neglecting the cable system’s
thermal inertia by the means of utilising the thermal capacitances in the TEE circuit.
Without the thermal capacitances, the steady-state condition can be achieved
immediately with zero time delay, thus the transient time domain, tθ, is considered to
be infinite, tθ=∞, because a steady state is achieved at t⟶∞.

Therefore, the TEE circuit in Figure 2-5 can be simplified into a circuit without
capacitances, as shown in Figure 2-6. The conductor temperature θC,I of cable i is
represented as the voltage at the conductor node, and this can be easily calculated by
summing up the voltage drops across the individual subcomponents. Therefore, a
simple electrical circuit analysis tool could obtain results to the steady-state TEE
model.

C ,i (tt ) S ,i (tt ) Ar ,i (tt ) Air ,i (tt ) Soil ,i (tt )


TIns TArm TJ Tair Tduct TSL1 TSL 2 TSLn
WC Wd Wd W W
S Ar TSoil
2 2

Conductor Insulation Screen & Armor Jacket Air & Duct Soil Layers 1...n
Figure 2-6 Steady-state TEE model for single core cable.

63
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

All of the parameters utilized in TEE model, as well as the multi-layered soil
properties is calculated according to [15, 16, 66]. It is assumed that screen losses, WS,
and armour losses, WAr, are directly related to the current in the conductor, and thus
they are given as a fraction of the conductor losses, where WS=λ1×WC, WAr=λ2×WC. The
steady-state conductor temperature, SS-θC,i(tΔt), of cable i at time tΔt can be expressed
by (2.6), where TSoil is the thermal resistance of the surrounding soil, which equals
the sum of the thermal resistances of each soil layer, TSoil=∑nk=1 TSLn .

SS-C ,i  tt  =(WC +0.5Wd ) TIns + WC +WS +Wd  TArm


+ WC +WS +WAr +Wd  (TJ +TAir +Tduct +TSoil )+ Soil ,i  tt 
(2.6)
=(WC +0.5Wd ) TIns + WC (1+1 )+Wd  TArm
+ WC (1+1 +2 )+Wd  (TJ +TAir +Tduct +TSoil )+ Soil ,i  tt 

2
The conductor losses, WC, can be given by WC= RAC×𝐼𝑓𝑙𝑜𝑤,𝑖 (𝑡𝛥𝑡 ), where RAC is the AC
electrical resistance of the conductor, and Iflow,i(tΔt) is the current flow of cable i at
time tΔt. Therefore, the steady-state loadability of the cable is stated in (2.7) where
the limitation is determined by the maximum conductor temperature θC,Max. This
method has been proved as being sufficiently accurate through years of service
experience, and has been utilised within this thesis for continuous/normal rating
calculations for UGCs.

C ,Max - Soil (tt )  -Wd 0.5TIns +TArm +TJ +TAir +Tduct +TSoil 
I Max ,SS = (2.7)
RACTIns +RAC (1+1 )TArm +RAC (1+1 +2 )(TJ +TAir +Tduct +TSoil )

The calculation of the thermal parameters in the Equation (2.6) and (2.7) varies
significantly with the different cable designs and laying conditions. For an in-depth
description, the reader is recommended to refer to [15, 16]. However, some thermal
parameters for each of the three single-phase cables have been provided here. The
thermal resistance calculation for each internal cable component has been stated in
(2.8) to (2.10), where ρT,Ins, ρT,Arm and ρT,J is the thermal resistivity of the insulation
material, armour and the jacket material; di, dS and dAr is the outer diameter of the
insulation, insulation screen and armour, dc is the diameter of the conductor and De
is the outer cable diameter.

64
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

T ,ins  di 
Tins  ln   (2.8)
2  dc 
T , Arm  d Ar 
TArm  ln   (2.9)
2  dS 
T , J  De 
TJ  ln (2.10)
2  d Ar 

The thermal resistance of the air between cable and duct, Tair, has been calculated in
(2.11), where θm is the mean temperature of the air between cable and duct, U, V and
Y are the constant values depending on the installations and materials. For normal
plastic ducts, U=1.87, V=0.312, Y=0.0037. The thermal resistance of the duct itself,
Tduct, is stated in (2.12), where ρT,duct is the thermal resistivity of the ducting, Do and
Dd is the outside and inside diameter of the duct.

U
Tair  (2.11)
1  0.1(V  Ym ) De
T ,duct  Do 
Tduct  ln   (2.12)
2  Dd 

The calculation of external thermal resistance, Tsoil, varies depending on the


installation methods. Three typical laying conditions for three single phase cables
have been given as examples below. It is worth being mentioned that if the cables are
buried in ducts rather than buried directly, the outer cable diameter, De, in (2.13)-
(2.15) needs to be replaced by the outside diameter of the duct, Do.

For three single phase cables buried in a flat formation (not touching), the equation
has been given in (2.13), where ρT,soil is the thermal resistivity of the surrounding soil,
in Km/W, s1 is the axial separation between the two adjacent cables and L is the laying
depth (distance from the ground surface to cable axis). For three single phase cables
buried in flat formation (touching), the equation has been given in (2.14).

65
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

     2 L 2  

2
  2L  2L  
TSoil  T ,Soil ln     1   ln 1     (2.13)
2   De  De     s1   
   
 4L 
TSoil  T ,Soil [0.475 ln    0.346] (2.14)
 De 

For three single phase cables which are buried in a trefoil formation (touching), the
equation has been given in (2.15), where the thermal resistance of the jacket, TJ, as
calculated by (2.10), is multiplied by a factor of 1.6.

1.5  4L 
TSoil  T ,Soil [ln    0.63] (2.15)
  De 

The calculation for the power losses in the cable components is quite complex. The
reader is asked referred to [15, 16] for more details. Here, Equation (2.16) and (2.17)
provide the fundamental equations for the calculation of conductor losses, WC, and
dielectric losses, Wd, where Iflow is the current flow through the cable conductor, RAC
is the AC resistance of the conductor per meter, E is the phase to ground voltage, ω
equals to 2π×frequency and tanδ is the loss tangent of dielectric, for XLPE,
tanδ=0.005. The calculated conductor losses, WC, and dielectric losses, Wd, are input
into the parameters in TEE circuit, such as the examples in Figure 2 5 and Figure 2 6.

WC  I 2flow RAC (2.16)

Wd  E 2C tan  (2.17)

2.5.2 UGC Transient Thermal Modelling

The transient thermal model was established by considering the cable system’s
thermal inertia by the means of thermal capacitances in the TEE circuit [37, 38]. The
transient TEE model was derived from the differential equations solving the voltages
(i.e. temperatures) in the TEE circuit of Figure 2-5. Thus, a matrix equation (2.18) was
set up with generic definitions as in Table 2-1.

66
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

 θSLn   M11 M12 0 0 0 0 0 0   θSLn   P 


     T-α
  
θSLn-1   M 21 M 22 M 23 0 0 0 0 0  θSLn-1   0 
   
   0 M 32 M 33 M 34 0 0 0 0     
   
θ   0 0 0 0 0  
θ    (2.18)
 SL1       SL1    
θ   0 0 0 0 0   θ air   0 
 air       
 θ Ar   0 0 0 0 0   θ Ar   PT-Ar 
       
 θS   0 0 0 0 0 M n+2,n+1 M n+2,n+2 M n+2,n+3   θS   PT-S 
       
 θ C   0 0 0 0 0 0 M n+3,n+2 M n+3,n+3   θ C   PT-C 

The matrix, M, in (2.18) used in the transient thermal calculation can be computed
using Table 2-1. CSL(k) and T'(k) is the thermal capacitance and resistance in the layer
k of TEE circuit, ∀k∈{-m, -m+1, -m+2,⋯, n}. All soil layers are represented by positive
k, while the cable’s ducting layer is represented when k=0. Negative k indicates
internal cable layers (e.g. jacket, armour, screen and conductor). The component, P,
in (2.18) computes external heat sources PT-k in layer k.

Table 2-1Thermal Matrix M & Layer Thermal Properties


Calculation of the thermal matrix M elements (Mij)
Left-diagonal Mij  1  CSL(n1i )  T(n1i )  , for i=1, 2,⋯, n+m+1, j=i-1
Right-diagonal Mij  1  CSL(n1i )  T(ni )  , for i=1, 2,⋯, n+m+1, j=i+1
Diagonal Mii    Mi (i1)  Mi (i1)  , for i=1, 2,⋯, n+m+1
Calculation of layer k thermal capacitance, CSL(k), and resistance, T'(k)
T(k ) = TSL ( k ) 2+ TSL ( k 1) 2 , k  1, 2, , n -1, n for soil
Layer thermal  =Tduct  TSL1 2
T(0)
resistance T(1) =TJ  Tair 
T'(k) T(2) =TArm  , k  -m, -m+1, , 0 for cable
T(3) =TIns 

CSL( k ) , k =1, 2, , n-1, n for soil
Layer thermal CSL(0) =Cduct 
capacitance CSL(-1)  CJ =CJ  CAr 
CSL(k) CSL(-2) =CS  Cd 2  , k =-m, -m+1, , 0 for cable, duct
CSL(-3) =CC =CC  Cd 2 

Calculation of the external thermal power PT-k in layer k
Layer external PT-α =a (tt ) (CSL(n)  T(n)
 ), PT-Ar =WAr CSL(0)
thermal power PT-S =(WS  Wd 2) CSL(-1) , PT-C =(WC  Wd 2) CSL(-2)

67
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

The thermal capacitance per metre of the different cable components was calculated
based on the cross sectional area of the zone ‘k’, as given in (2.19) where cthermo,k is
the volumetric specific heat of the material in zone ‘k’, rk and rk-1 is the outer and inner
radius of zone ‘k’.

Ck  cthermo,k     rk2  rk21  (2.19)

The electrical circuit analysis techniques in [37, 38] were used to find the differential
solution of the cable for TEE model in order to calculate the transient thermal
response of θ by finding the matrix M’s eigenvalues and eigenvectors. The matrix
equation (2.18) can be expressed in a compacted form, as represented in (2.20). By
denoting the eigenvalues of M by λ1, λ2, …λn+3 and denoting the eigenvectors by v1,
v2, …vn+3, the solution for (2.20) has been given in (2.21).

  t   M   t   PT k (2.20)

  t   c1  v1  e t 
1 
 cn3  vn3  e n  3 t
   (2.21)

θ(∞) is the steady-state temperature by neglecting the capacitances, as in (2.22).

 θSLn       WC  tΔt  +Wd  tΔt  +WS  tΔt  +WAr  tΔt   ×  TSLn 2  +θSoil  tΔt  
   
θ      θ   +  WC  tΔt  +Wd  tΔt  +WS  tΔt  +WAr  tΔt   ×  TSLn-1 2+TSLn 2  

 SLn-1   SLn 
   
   
   
 θSL1      θSL2    +  WC  tΔt  +Wd  tΔt  +WS  tΔt  +WAr  tΔt   ×  TSL1 2+TSL2 2   (2.22)
 = 
 θair      θSL1    +  WC  tΔt  +Wd  tΔt  +WS  tΔt  +WAr  tΔt   ×  TSL1 2+Tduct  
   
 θ    θair    +  WC  tΔt  +Wd  tΔt  +WS  tΔt  +WAr  tΔt   ×  TJ +Tair  
 Ar   
   
 θS      θAr    +  WC  tΔt  +Wd  tΔt  +WS  tΔt   ×TArm 
   
 θC      θS    +  WC  tΔt  + Wd  tΔt  2  ×TIns 

When tθ=0 is applied, the equation (2.21) was transformed into (2.23). The constants
c1, c2, …cn+3 have been calculated by solving (2.23).

68
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

 c1 
c 
v1 v2 vn3    2     0       (2.23)
 
 n 3 
c

By using (2.21) to solve the differential equations in the evaluation of the


temperatures in the power cables, the transient conductor temperature, TS-θC,a(tΔt),
can be calculated by applying tθ=Δt into (2.21).

2.5.3 UGC Thermal Ageing Modelling

Cable failures may occur due to poor workmanship, sudden mechanical shock and
manufacturing defects. This study mainly consider the slow and continuous
degradation or ageing by operational thermal stress. The well-known thermal life
model, the Arrhenius life model, was described in (2.24) to calculate the cable’s
thermal lifetime, L(θC), at θC. where B is equal to Ea/kB, Ea is the activation energy
and kB is the Boltzmann’s constant (in еV/K), θ0 is the ambient reference temperature,
θC is the conductor temperature and L0 is the corresponding cable life at temperature
θ0.

L(C )  L0 exp( BcT ) cT  1 0 1 C (2.24)

In the methodological framework, the calculated temperatures SS-θC,a(tΔt), and TS-


θC,a(tΔt), are input into the ageing calculation conducted through the Arrhenius model
in order to capture the each cable’s life loss within Δt, LFθC,i (tΔt ) . To enable the
comparison of ageing between different cables in different networks, an equivalent
cable ageing (ECA) index was proposed and has been described by (2.26). The index
converts the cables’ ageing occurred at different temperature into an equivalent
ageing at 90 °C, which is the most common continuous loading temperature for XLPE
cables.

t
LFC ,i (tt )   Eai k B 1 o ,i 1 C ,i (tt )  (2.25)
Lo,i  e

69
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

LFC ,i (tt )  Ea  1 1 
ECAi ,t   exp  i    (2.26)
LFi ,90C  k B  363 C ,i (tt )  
   

In order to compare cable-based networks with OHL-based networks, the thermal-


ageing modelling of OHLs was also integrated into the framework based on the
previous research [58, 59, 67] and IEEE standard [68]. The thermal-ageing modelling
of OHLs converted the current flow, Iflow,j(tΔt), for every network line j, ∀j∈NOHL, at Δt,
to its operating temperature θC,j(tΔt), where NOHL is the number of overhead lines.
Both steady-state, SS-θC,j(tΔt), and transient-state, TS-θC,j(tΔt), temperatures of the line
were calculated using the OHL steady-state and transient model. Then, the equation
calculates the equivalent thermal ageing that occurred within the Δt using the index
Equivalent OHL Ageing (ELA).

2.6.1 OHL Steady-state Thermal Model

In IEEE Standard 738 [68], the steady-state temperature of OHL conductor can be
calculated based on its current flow and the ambient temperature. A numerical
iteration method is utilized to solve the heat balance equation in (2.27) considering
the current loading and weather variables, where qS is the solar heat gain, qC is the
convection heat loss, qR is the radiated heat loss, Iflow is the current flow across the
conductor, R(θavg) is the conductor resistance at θavg and θavg is the average
temperature of aluminium strand layers.

qc  qr  qs  I 2flow  R avg  (2.27)

The calculation process within the numerical method has been stated in detail in [68].
It returns to the steady-state conductor temperature SS-θC,j(tΔt) of line j, using steady-
state weather conditions.

2.6.2 OHL Transient Thermal Model

The transient temperature response to changes in the conductor current and time-
varying weather conditions was calculated based on the methods discussed in IEEE

70
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

Standard 738 [68] through a iterative transient calculation. The conductor


temperature change Δθavg, within Δtθ, was calculated using (2.28), where Δtθ is the
iterative time interval of 10 seconds, mCp is the total conductor heat capacity. The
transient conductor temperature TS-θC,j(tΔt) at the end of each given Δt was calculated
using (2.29) as its initial conductor temperature TS-θC,j(tΔt-Δt) at time tΔt-Δt, plus the
total temperature change within Δt.

R avg  I 2flow  qS  qC  qR
avg  t (2.28)
mC p
t
TS-C , j (tt )  TS-C , j (tt -t )    avg (2.29)
t 0

2.6.3 OHL Thermal Ageing Estimation

The evaluated conductor ageing εj,θc was calculated using (2.30) for an OHL, j,
operating at θC,j(tΔt) for a duration tj,θc, while σj,Tc is the installation stress, Kj describes
the conductor type [58, 69]. With aim to make a quantitative comparison between
the ageing occurred in different OHLs, an equivalent ageing, εj,100°C at 100°C is
computed using (2.31). Consequently, the equivalent OHL ageing, ELAj, was
calculated using (2.32), describing the total sum of equivalent ageing occurred in
every elevated temperature events (ETE).

 j ,  f  K j ,c , j , , t 0.16
C C j , 
C
(2.30)

 j ,100C  f  K j ,100C, j ,100C , t 0.16


j ,100C  (2.31)

6.25
  j ,C
ETE 
ELAj     t 0.16  (2.32)

e1   j ,100C
j , C 

The network and branch performance indices were captured using the mathematical
equation provided in this section. Commonly used transmission and distribution
network indices were provided, and some novel branch indices were created to
indicate the loading conditions and ageing performance for each individual circuit.

71
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

2.7.1 Branch Indices

The branch performance indices were computed with (2.33)-(2.38). The ECA and
ELA was captured for each branch x and Δt of every simulation year Y. Expected
Equivalent Cable Ageing (EECA) and Expected Equivalent Line Ageing (EELA) were
the mean expected ECAi and ELAj within N simulation years, respectively. The
Expected Frequency of Cable Failure (EFCF) and Expected Frequency of Line Failure
(EFLF) is the expected average frequency of cable failure, NCFi, and line failures, NLFj.
The Expected Duration of Emergency Loading (EDEL) calculates the average duration
of emergency loading, d_ELb, for each branch b in each year Y. The Expected
Frequency of Emergency Loading (EFEL) computes the annual average frequency of
emergency loading events, N_ELb, for every branch b.

NY
 Y 
EECAi     ECAi ,t  NY (2.33)
Y 1  t 1 
NY
 Y 
EELAj     ELAj ,t  NY (2.34)
Y 1  t 1 
NY
EFCFi   NCFi ,Y NY (2.35)
Y 1

NY
EFLFj   NLFj ,Y NY (2.36)
Y 1

NY
EDELb   d_ELb,Y NY (2.37)
Y 1

NY
EFELb   N_ELb,Y NY (2.38)
Y 1

2.7.2 Transmission Network Indices

The Expected Duration of Load Curtailment (EDLC), Expected Frequency of Load


Curtailment (EFLC), and Expected Energy Not Supplied (EENS) were calculated using
(2.39)-(2.41), where LCDY is the system load curtailment duration in year Y, LCEY is
the number of load curtailment events in year Y, SysENSY is the system energy not
supplied (in MWh) in year Y. The Expected Equivalent Network Ageing (EENA)

72
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

captured the total (annual) network ageing occurred in UGCs or OHLs in year Y, as
calculated in (2.42). The Expected Frequency of Network Failure (EFNF) was
computed in (2.43) in order to capture the annual number of network circuit failures
of the UGCs and OHLs in year Y. The Expected Interruption Cost (EIC) was used to
calculate the average cost of customer interruption using the interruption cost
function, Costlp, based on the interruption duration, dInterrupted,lp and the interrupted
lp lp
demand, Dactual − Dtotal , of each load point, lp [3, 60].

NY
EDLC   LCDY NY (2.39)
Y 1

NY
EFLC   LCEY NY (2.40)
i 1

NY
EENS   SysENSY NY (2.41)
Y 1

NUGC NOHL
EENA   EECA   EELA
a 1
a
b 1
b (2.42)

NUGC NOHL
EFNF   EFCF   EFLF
a 1
Y
b 1
Y (2.43)

 Nlp 
EIC     Costlp  dInterrupted ,lp , Dactual
NY
lp lp
-Dtotal   NY (2.44)
Y 1  lp 1 

2.7.3 Distribution Network Performance Indices

The Customer Average Energy Not Supplied (AENS), System Average Interruption
Frequency Index (SAIFI), System Average Interruption Duration Index (SAIDI) and
the Average Service Availability Index were evaluated based on (2.45)-(2.48), where
NCus is the number of customers served in the distribution system, NCus,Interrupted is the
number of interrupted customers, DCus,Interrupted is the total duration of the customer’s
interruption, and Davailable is the customer hours of available service in year Y.

AENS  EENS NCus (2.45)

73
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

NY
N 
SAIFI    Cus , Interrupted  NY (2.46)
Y 1  NCus 
NY
D 
SAIDI    Cus ,Interrupted  NY (2.47)
Y 1  NCus 
D 
N
ASAI    available  NY
Y

(2.48)
Y 1  8760 

A new network reliability assessment framework has been proposed in this chapter,
which is a combination of several new methodological elements in both plant design
and network operation. It integrates an electro-thermal model of underground cables
within an SMC-based network reliability assessment framework. This methodology
has progressed the current work on power system reliability assessments a step
further to include the design properties of the cable and in order to calculate the
ageing of all of the cables within the network. It enables the evaluation of the network
resilience and operator flexibility when utilising the flexible cable rating methods
(such as probabilistic or emergency rating) that can be generated from the cable’s
thermal inertia and when compared against the potential risk of increased cable
thermal ageing.

The method considers the cable properties, its surrounding design conditions and
network topology to quantify the expected cable ageing risk. A generic structure of
TEE cable modelling has been proposed to allow for its application in multiple types
of cables (oil, EXLPE, oil/paper, etc.) and soil. The methodology allows for the time-
dependency of the TEE, hence they can be used in relation to both the steady state
(SS) and transient state (TS) of the thermal modelling.

It is critical to identify the operating temperature profile of every cable in the network
in order to compute the expected cable ageing. This can then be used to strategically
manage the network operations and to re-define the suitable rating strategies. This
can be used to optimise the overall network performance and to avoid unnecessary
ageing of the cables.

74
Chapter 2: Reliability Evaluation Framework Integrating Cable Design and Ageing

In addition, it provides generic metrics that allow for the method to record useful risk
quantities. The newly proposed indices can assist network operators when forming
their asset management strategies. The proposed ECA index formulates a per-unit
base value for the ageing that occurred on different cables at different current
loadings. Hence, ECA allows us to aggregate the annual ageing and also to compare
the ageing of different cables in the same system. Therefore, the ageing outputs from
the proposed methodology framework can be used to identify the critical cable(s)
based on a relevant (per-unit) ageing quantification of the network cables.

Equation Chapter (Next) Section 1

75
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

3 Impact of Cables’ High-loading Risks on


Network Reliability

Power utilities are challenged to improve their power transfer capabilities and to
increase their resiliency under uncertain circumstances for their existing
networks. This can be resolved by allowing their underground cables to utilise
flexible operating ratings and to allow them to operate at evaluated temperatures
under emergency contingencies. However, current studies does not capture the
flexibility of those schemes and the increased risk of cables when utilising
emergency loading operations.

This chapter presents an innovative methodology in order to evaluate system


reliability accounting for the additional failure risk and the extra ageing when the
cables are operating at emergency loadings. The proposed methodology filled up
the research gap through the integration of a thermal-ageing model of UGCs into a
novel double SMC loop. Three Markov states describing cable emergency operation
were modelled where the state transitions are determined according to two new
proposed risk index, α and β. This proposed methodology can help utilities to
quantify the short and long term risks of emergency rating utilisation on cable
ageing and network performance. It can also quantify the impact of increasing high-
loading failure risks on the network performance within the cables’ life cycle. The
proposed cable performance indices helped to indicate the critically aged cables
and informed the utilities of strategic network maintenance and replacement
practices.

The proposed methodology was applied to a cable-based 14-bus network under


the LTE and STE ratings utilising steady-state and transient-state TEE modelling.
The results indicated that emergency ratings initially produced improved network
performance as a result of the increased flexibility. However, the increased failure
risks reduced the network performance as the ‘year in service’ increased.

76
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

Power utilities try to address challenges related to the security and sustainability of
their networks within a competitive environment. This is often aligned with the
growing need for renewable energy connections, increased demand, improved
energy efficiency and expanding technological dispersion. To effectively conquer the
challenges, power utilities have investigated potential flexible and economic
solutions rather than using new builds, to increase the power flow adequacies.
Practices that are usually implemented and involved with increased risk include (i.e.,
reduced conservatism), probabilistic thermal rating [17, 70, 71] and real-time
measurement for time-varying line ratings [27, 72, 73]. More flexible operation
schemes can also help utilities to achieve that goal for improving their network
adequacy, i.e. implementing emergency or more flexible line ratings under
contingency events or during emergency loading conditions[58, 61, 62].

Although such flexible solutions increases network adequacies economically, these is


achieved by relaxing the thermal constraints. Thus, this often includes an increased
ageing risk of network facilities (transformer, cables, lines) which has not been
considered in most studies and network modelling before.

Underground cables usually own a much larger thermal inertia when compared to
overhead lines [38]. This large thermal inertia origins from the large thermal mass of
the cable insulation materials, and surrounding soil. Therefore, an obvious time delay
of cable’s temperature response can be observed when its current loading is
increased [21]. This characteristic provides the possibility for cables to be loaded up
to twice their normal rating for a short period of time, especially during emergency
operations [38]. IEEE standards permit XLPE cables to be overloaded at an
emergency conductor temperature of 130°C for a maximum duration of 36 hours [51].
Data in [74] also indicated the LTE and STE ratings for both UGCs and OHLs after a
contingency event. Consequently, the emergency ratings of UGCs are well
acknowledged and utilized by power utilities [23, 75, 76]. However, there is a limited
amount of literature that considers the benefits of emergency ratings and weights
those against the risks within a network-wide analysis. Most research assessed only

77
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

the potential capability of time varying thermal rating to improve the power flow
adequacy [28, 30, 44], rather than its risks..

In order to develop models to implement the UGC emergency ratings in a realistic


way, and thus to provide applicable tools to utilities, the modelling should include
three basic elements: (a) a thermal modelling that integrates the design factors of
UGCs; (b) a thermal-ageing modelling which takes account of the ‘overheating’ impact
on thermal degradation and lifetime reduction of UGCs and (c) a failure model that
considers the failure risks due to increased ageing risks [62]. Many thermal models
of UGCs are established using FEM methods [28, 48] or TEE approaches [38, 65] for
computing cables’ core temperature according to a known current loading, cable
design and environmental conditions. Some of the models have been included within
the network-wide evaluations [37]. The cable system’s thermal inertia is neglected in
these existing research, however, the great amount of flexibility in cable-based
network can be increased by considering the inertia. Foremost, current studies take
no consideration of the ageing that is expected in the cable, thus neglecting the
potential costs and risks of utilizing such approaches.

Although some studies have investigated quantifying the effect of high temperature
and high loading on cable thermal ageing and lifetime reduction [21, 55], only limited
work incorporate the cable ageing into the network evaluations [11, 56, 77].
Furthermore, these research have failed to consider the effects of emergency loading
conditions on cable ageing, which results in an increasing risk of failure [21, 52].

This chapter proposes a novel network reliability assessment methodology in order


to quantify the network’s flexibility and cables’ ageing risks when allowing
emergency loading on UGC. This method advances the existing practice by
quantifying the benefits gained from additional flexibility, but also while regarding
the costs from ageing risks and extra failure possibility of UGCs. This methodology
will help utilities to identify cases where increasing the utilisation of UGCs improves
the network performance and reduces the risks by informing the operator’s decisions
on maintenance and asset management.

78
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

Section 3.2 introduces the integration of the UGCs’ thermal-ageing modelling into the
SMC-based network reliability evaluation. Section 3.3 describes the demonstration of
the methodology within a case study of modified 14-bus network, and Section 3.4 and
3.5 discusses the key findings from the results of the case study. Finally, Section 3.6
discusses the key conclusions.

3.2.1 Outline of a Double Sequential Monte Carlo Simulation

To quantify the risk of failures when cables are experienced under emergencies, the
designed network reliability evaluation method combines a thermal-ageing
modelling of UGCs with a SMC-based network analysis. This combination is realised
using double SMC loop, which captures the annual network performance and
individual cable operational thermal profiles as well as their long-term life-cycle
performance.

Within the first SMC loop, two different operation states were employed for cables to
be loaded under normal and emergency loadings to account for the additional cable
failure possibility and increased thermal ageing. The second SMC loop was
implemented to record the cumulative effects of emergency loading on insulation
degradation and its effect on cable failures. The outline flow of this DSMC
methodology is shown in Figure 3-1.

The system analysis was then performed in the two loop mode. The inner 1st SMC-
loop performed an annual network analysis, for each simulation year Y, ∀Y∈{1,⋯, N}
in a sequential time tΔt, divided into steps Δt, ∀Δt∈Y. It considers the additional risks
of cable ageing and potential failures, by updating network component status
changes, if any, due to high-loading operation occurred at Δt. It can be seen in Figure
3-1 that this loop is capable of initially simulating all of the normal failures within the
Initial network component status mapping, and then capturing all potential cascading
and high-loading failures as a result of the emergency loading operation within high-
loading component status mapping.

79
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

NETWORK DATA CABLE SYSTEM DATA


Reliability Operation Topology Ducting & Laying Soil Design
System data initialization

Cable emergency operation history mapping

Initial network component status mapping


Normal failures

1st SMC-loop: Year network evaluation


Life-cycle TTFE<Δt Network Time-step
analysis PF calculation optimization analysis
〈YL=YL+1〉 〈tΔt=tΔt+Δt〉
High-loading component status mapping
Cascading No high-loading High-loading
failures (CF) failures failures (HLF)

2nd SMC-loop: Life-cycle evaluation


Annual Component Temperature mapping – θC,i(tΔt)
analysis
〈Y=Y+1〉 TTFE≥Δt
Annual computation is completed
NO
YES
Accuracy of covariance is achieved
NO YES
Annual network & cable performance indices calculation

Life-Cycle iteration is completed at pre-defined YLife


NO
YES
Completed network reliability and cable ageing evaluation

Figure 3-1 Flowchart of the double SMC methodology [62]

Cascading failures are not captured within the network optimisation, which is an AC
OPF under network operation constraints (introduced in Section 2.4). This considers
the operator’s incompetency to make any corrective response within a very short
period of time when a sudden overload protection works. On the other hand, the
high-loading failures do not happen instantaneously within the same Δt, providing
the operator with more flexibility to respond by enabling for extra post-contingency
loading. But the additional risk under high-loading operation is captured within the
component temperature mapping block. This block utilises the thermal-ageing
modelling discussed in Section 2.5 to calculate the expected cable temperature
θC,i(tΔt), at every Δt, of every cable i, and the associated risk of cable ageing. The
increased risk of failure experienced by the high-loading cables is modelled through
the dashed and dotted lines and the computation of emergency time-to-fail (TTFE)
described in Section 3.2.2. The first SMC loop is completed if a target covariance of
EENS or a specific number of simulation years NY, is reached [60].

80
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

The outer second SMC-loop performed a life-cycle evaluation in life-year steps YL,
∀YL∈{1,⋯, YLife}, considering the calculation of every year’s cable and network
performance indices to construct the updated network’s cable emergency operation
history mapping. This step aggregates all of the previous expected emergency loading
durations and expected cable ageing to calculate a cumulative failure risk of the
cables. Thus, some results from the first loop were recorded and utilised in the
second loop to include previous emergency operation history and compute the
cumulative failure risk at this new life year YL.

3.2.2 1st SMC loop Component Status Mapping Calculation

In the component status mapping of the 1st SMC-loop, the operating state of each
cable at every Δt is determined. The computation was performed through two steps
(Figure 3-1): the initial network component status mapping-Normal failures and the
high-loading component status mapping-Emergency conditions.

1) Initial network component status mapping – Normal failures

At the beginning of each simulation year, the initial network component status
mapping calculated the initial transition times of all of the network components
assuming normal loading. A two-state Markov chain is implemented as indicated in
Figure 3-2 (left).

Equation (3.1) and (3.2) determine the initial annual normal-loading state to normal-
failure state transition times for every normal-loaded component i', ∀i'∈ NUGC, with
NUGC being the total number of cables. The normal failure λN,i', and repair μN,i', rates of
every component i', are utilized to compute the time-to-fail TTFN,i', and time-to-repair
TTRN,i' with a normal unavailability UN,i' [3]. Then, the TTFN,i', and TTRN,i' were used to
map the initial component status of each component i', for complete simulation year,
which has been described in Section 2.3.

TTFN , i '   1 N , i '  ln U N , i ' (3.1)

TTRN , i '   1 N , i '  lnU N , i ' (3.2)

81
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

2) High-loading component status mapping – Emergency conditions

After the status of normal-loaded components i' in year Y have been determined, a
three-state Markov chain is utilized to identify the status transitions of high-loading
components i". The three states in Markov chain includes: Cascading failure state, No
high-loading failure state, and High-loading failure state, as shown in Figure 3-2. This
mapping requires the network loading status at every Δt. Hence, network
optimisation is performed at present Δt, following the process summarised in Section
2.4, to determine the network’s current flows Iflow,i(tΔt) on each cables i at Δt. Worth
to mention that a cable is identified to be in a high-loading state when its current flow
exceeds its normal continuous current, ICon,i.

Normal loading High loading


Normal-loading Network status change High-loading
State State
λN,i' μN,i' Cascading failure YES
Iflow,i(tΔt) >ISTE,i
state (λE,i"=∞)
YES NO
μN,i' YES Iflow,i(tΔt) >ILTE,i
TTFSTE,i" ≤ Δt (λSTE,i")
Normal failure ------- or ------- NO
state TTFLTE,i" ≤ Δt
(λLTE,i")
NO
Normal Operation TTFSTE,i" ≫ Δt YES
Markov States ------- or -------
TTFLTE,i" ≫ Δt (λE,i" ≈0)
Transition paths NO
Iflow,i(tΔt∈TTFSTE,i")>ILTE,i
STE NO No high-loading
---------- or ---------- failure state
LTE Iflow,i(tΔt∈TTFLTE,i")>Icon,i (λE,i" ≈0)
YES
i': normal loading High-loading High-loading Operation
i": high-loading μN,i' failure state Markov States
Figure 3-2 Operation states of a cable i considering its normal (i') and high-
loading (i") conditions [62]

a) Transition path probabilities

A cable may transition between normal-loading state and high-loading state


depending on whether its current loading is above ICon,i or not. A cable’s current flow
may be impacted by various factors, such as adverse weather, network load changes
and the failure of other components. Therefore, these two states are considered to be
a single state without any transition probability as shown in Figure 3-2.

82
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

For a high-loading cables, an emergency time-to-fail (TTFE) is computed based on


(3.3)–(3.8) for the STE loadings (ILTE,i<Iflow,i(tΔt)≤ISTE,i) and LTE loadings
(ICon,i<Iflow,i(tΔt)≤ILTE,i).

TTFLTE , i "  (1 LTE , i " )ln U LTE , i " (3.3)

TTFSTE , i "  (1 STE , i " )lnU STE , i " (3.4)

(( I STE , i Icon , i )1)


STE , i "  N , i "  e (3.5)

(( I LTE , i Icon , i )1)


LTE , i "  N , i "  e (3.6)

U STE , i "  1  U N , i " ( I con ,i / I STE ,i )  (3.7)

U LTE , i "  1  U N , i " ( I con,i / I LTE ,i )  (3.8)

Due to different emergency loading events are independent, the unavailabilities of


USTE,i" and ULTE,i" were calculated by (3.7) and (3.8) individually. These equations were
formed by the functions of current loadings considering the risk of unavailability
increases along with the increase of cable loading. The equation involved with two
new coefficients, α and β, in order to determine the design risk aspect and operation
risk aspect.

Design risk coefficient α was implemented to define the cable’s failure risk related to
its design and condition. It indicates the cable’s design properties (e.g. laying in
unfavourable soil conditions, crossing road) and the repair and maintenance
standard. Thus, α value was determined for each individual cable based on the cable’s
location and operators’ experience. A zero α (α=0) suggests null unfavourable
conditions along the cable and an easily accessible cable for repairs.

An ageing acceptability risk coefficient β was implemented to define the operator risk
acceptability, indicating the operator’s willingness to implement cable emergency
ratings. A zero β (β=0) indicates a brand new cable that has undertaken no high-
loading operation.

The repair duration TTRE, under emergency events is considered to be the same with
the repair time under normal-loading events. Thus, the TTRN,i' and TTRN,i" were

83
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

produced using μN,i' for normal failures, high-loading failures, and cascading failures
of each cable i using (3.2).

b) High-loading states

The cascading failure state indicates the overload protection is triggered


instantaneously due to a overloading of a cable i". Three paths exist to transit to a
cascading failure state as shown in Figure 3-2. One path simulates the instantaneous
thermal protection ‘tripping’ action, when Iflow,i(tΔt) exceeds its STE rating, ISTE,i. Thus,
the cable will definitely transition to a cascading failure state considering λE,i"=∞. The
other two paths simulate the immediate failure (i.e., TTFE<Δt) due to the increased
probability of failure expected when the cable is loaded above ICon. Thus, one path
(continuous line) captures the failure probability for operations within STE loading,
ILTE,i<Iflow,i(tΔt)≤ISTE,i, described by TTFSTE,i"≤Δt. The other one captures the failure
probability for operations within LTE loading, Icon,i<Iflow,i(tΔt)≤ILTE,i", described by
TTFLTE,i"≤Δt. The TTFLTE,i" and TTFSTE,”". The two paths were calculated using (3.3) and
(3.4), respectively.

The high-loading failure state considers a failure resulting from a high-loading


operation of a cable i" where the TTFE is greater than Δt. Thus, this high-loading
failure occurs beyond tΔt in the future. However, the failure occurs only when the
corresponding cable i" remains at high-loading condition during the whole TTFSTE,i"
(or TTFLTE,i"). As shown in Figure 3-1, the TTFSTE,i" (or TTFLTE,i") should be greater than
Δt, and the cable current flow Iflow,i(tΔt) should be larger than ILTE,I, Iflow,i(tΔt)>ILTE,I, for
all Δt∈TTFSTE,i" (or Iflow,i(tΔt)>Icon,i for all Δt∈TTFLTE,i").

The No high-loading failure state happens when a high-loaded cable experiences no


failure. This could occur due to a sufficiently large TTFSTE,i" or TTFLTE,i" duration that
results in no failure event, or due to a break of a high-loading operation within the
TTFSTE,i" or TTFLTE,i" periods. As indicated in Figure 3-1, when a high-loaded cable
transitions to its normal-loading state, prior to its TTFSTE,i" or TTFLTE,i", the expected
high-loading failure is ‘cleared’ and it is classified as a No high-loading failure state.

84
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

Figure 3-3 illustrates how the transition mapping was performed for an example of
three components that transition to normal failure (NLF), high-loading failure (HLF)
and cascading failure (CF) states. The exact current flows in Figure 3 3 (b) determine
the actual current flows of the cables, while the actual simulated current flow
transition mapping is displayed in Figure 3 3 (c). As indicated in Figure 3-3, the failure
of cable 1' cause the cable 2" to be loaded at high-loadings. This calculation resulted
in HLF of the cable 2" after few Δt. The failure of cable 1' did not change Iflow,3 to high-
loading and thus, no change in cable 3 state occurred. However, the failure of cable 2"
resulted in Iflow,3 being at the STE range. This event initiated the TTFSTE,3" computation,
since it is within STE loading during the specific Δt, which resulted in a CF of cable 3".

(a) Exact state transition mapping for component 1, 2, and 3


i'=1' NLF Δt ... Δt ... TTRN,1' Δt ... Δt

i"=2" HLF TTRN,2"


TTFLTE,2"
i"=3" CF TTFSTE,3" TTRN,3"
(b) Exact current-flow transition mapping
I1,flow ISTE,3
ILTE,2
I2,flow
I3,flow
(c) Simulated current-flow transition mapping
I1,flow ISTE,3
ILTE,2
I2,flow
I3,flow
Time axis
Figure 3-3 An example of modelling the operating state and corresponding
current flow transitions [62]

3.2.3 2nd SMC-loop Component Status Mapping Computations

The second SMC loop propose a new coefficient βi,Y to sum up the previous the
L

emergency operation history for each cable. Equation (3.9) captures the increased
failure risk in cable i under frequent high-loading events, and it can be used to reflect
the effects of operational history on the cables’ lifetime. If no emergency loadings are

85
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

allowed on the cables by the utilities, then the EDEL would both be zero, resulting in
a zero βi,Y .
L

YLife YLife

i ,Y  a  DLTE ,i ,(Y 1)  b  DSTE ,i ,(Y 1)  a  EDELLTE ,i ,Y  b  EDELSTE ,i,Y (3.9)
L L L L L
YL 1 YL 1

In (3.9), DLTE,i,(YL-1) and DSTE,i,(YL-1) are the accumulated durations operating within
LTE and STE, of the cable i occurred prior to the year YL. a, b are the parameters that
characterise the cable’s degradation severity due to high-loading operations. These
parameters can be derived from experimentations or empirically estimated
according to operators’ recorded data.

The cable performance indices were computed by (3.10)-(3.15). The EECA was
calculated as the mean equivalent cable ageing (ECA) in the year YL of cable’s life-
cycle. The expected frequency of cable failure (EFCF) was the average frequency of
normal (NCFi,Y) plus emergency (ECFi,Y) cable failures. The expected duration of
emergency loading (EDEL) for both LTE and STE were calculated based on the total
short (d_STEi,Y) and long (d_LTEi,Y) emergency loading durations. The expected
frequency of emergency loading (EFEL) computed the mean value of the emergency
loading events N_emei,Y. EFEL and EDEL assessed the high-loading duration for each
contingency event and thus predicted the effect of cable ageing on network
performance considering the operator’s emergency control operations.

 Y
NY

EECAi ,YL     ECAi ,t  NY (3.10)
Y 1  t 1 
NY
EFCFi ,YL    NCFi ,Y  ECFi ,Y  NY (3.11)
Y 1

NY
EDELLTE ,i ,YL    d _ LTEi ,Y  NY (3.12)
Y 1

NY
EDELSTE ,i ,YL    d _ STEi ,Y  NY (3.13)
Y 1

EDELi ,YL  EDELSTE ,i ,YL  EDELLTE ,i ,YL (3.14)

86
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

NY
EFELi ,YL   N_ELi ,Y NY (3.15)
Y 1

The network performance indices were computed in Section 2.7 by (2.39)-(2.44),


including EENS, EDLC, EFLC, EENA and EFNF. These indices helped to evaluate the
cable ageing condition within the system, and thus were used to derive the best
replacement plan for UGCs by optimising the total network ageing. It can also allow
the utilities to prioritise the new cable installations when there are many cables at
similar ages but under different ageing conditions.

bus 13 bus 14
16

bus 12
15
13
11 14

bus 11 bus 10
10 9 12

Bus 6 bus 9
C C
bus 7 bus 8
bus 5 bus 4

7
3 5 C
6

bus 3

4
2

1
bus 1 bus 2 G Generators
Synchronous
G G C Condensers

Figure 3-4 Single line diagram of IEEE 14-bus network

A modified IEEE 14-bus network was utilised for the case study with an increased
load by 1.6 pu of default demand. The schematic diagram of the IEEE 14-bus network
is shown in Figure 3-4. Only failures occurred on cables are simulated, however other

87
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

network components, such as transformers and generators, are simulated to have no


failures. The IEEE RTS-96 chronological load profile was accepted for the hourly load
profile for the test 14-bus network [74].

3.3.1 Selection of Test Cable Properties

The circuit loading limits and design data is not provided in the original data of 14-
network. Table 3-1 details the four types of single-core XLPE cables with copper
conductors that were utilized for the test network [79, 80]. These were assumed to
lay in flat formation with a 0.1m separation at a depth of 0.8m and 1.5m for 13.8kV
and 69kV, respectively. Normal, the STE and LTE ratings of UGCs are implemented to
provide flexibility at emergencies. Their ILTE and ISTE ratings were 1.103 and 1.143
times of the continuous rating ICon, adopting the same proportions (ILTE/Icon, and
ISTE/Icon) of the 132kV cable used in RTS-96 [74]. The resulted cable thermal ratings
have been stated in Table 3-1. These ratings were calculated based on a 15°C soil
temperature and a 0.9km/W soil thermal resistivity, and a 1.9×106 J/m3K volume
specific heat according to the UK standard, ENA P17 [17].

Table 3-1 Cable Electrical Properties Modelled in an IEEE 14-bus Network


Cable Size R90°C X B ICon ILTE ISTE
Type mm Ω/km Ω/km
2 S/km A [°C] A [°C] A [°C]
A 400 0.060 0.119 6.49×10-5 790[90°] 871[111°] 903[121°]
B 630 0.038 0.109 8.38×10-5 1010[90°] 1114[110°] 1154[119°]
C 400 0.063 0.094 1.41×10-4 845[90°] 932[111°] 966[121°]
D 800 0.035 0.087 1.98×10-4 1200[90°] 1323[110°] 1371[119°]

The original 14-bus system data does not include any circuit lengths. The lengths are
needed to determine their failure and repair rates, when considering cables, as
defined by (3.16) from [81] and (3.17). Equation (3.17) used to calculate the repair
rates was derived from cable repair duration data provided in [81]. Based on the
original impedance of circuits and the 0.7 Ω/mile factor in [82], the length of circuit i,
ℓi, could be derived (in miles). The calculated lengths, reliability data and cable types
applied in the test network are in Table 3-2.

N ,i  0.0062  i  0.226 (occ./ yr ) (3.16)

88
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

1  N ,i  1.5  i  11.5 ( hr / occ.) (3.17)

An hourly soil temperature data from 2015 was utilized. It was achieved from the
ERA dataset provided by ECMWF [83]. For simplicity, a constant soil thermal
resistivity of 0.9km/W and a constant volume specific heat 1.9×106J/m3K were
adopted for the entire simulation year. The Arrhenius model parameter
Eai/kB=12430K was used for XLPE cables in the test network [21].

Table 3-2 Cable reliability properties modelled in the IEEE 14-bus network
Cable From To Voltage Length Outage Outage Cable
No. Bus Bus Level, kV miles Rate duration, Type
1 1 2 69 8.47 Occ./yr
0.279 hr
24.2 B
2 1 2 69 8.47 0.279 [hr/Occ.]
24.2 B
3 1 5 69 15.61 0.323 34.9 A
4 2 3 69 13.84 0.312 32.3 A
5 2 4 69 12.63 0.304 30.4 A
6 2 5 69 12.44 0.303 30.2 A
7 3 4 69 12.49 0.303 30.2 A
8 4 5 69 3.00 0.245 16.0 A
9 6 11 13.8 0.60 0.230 12.4 C
10 6 12 13.8 0.77 0.231 12.7 C
11 6 13 13.8 0.40 0.228 12.1 D
12 9 10 13.8 0.25 0.228 11.9 C
13 9 14 13.8 0.81 0.231 12.7 C
14 10 11 13.8 0.57 0.230 12.4 C
15 12 13 13.8 0.81 0.231 12.7 C
16 13 14 13.8 1.05 0.233 13.1 C

3.3.2 Modelled Scenarios

The study of the flexibility and operation risk due to emergency ratings (ERs) was
performed by employing the proposed methodology under the scenarios in Table 3-3.
The scenarios were developed for examining the effect of cable high-loading failures
on the network and cable performances when different cable design criticality
indexes α and risk indexes β were applied. The base case does not consider the
utilisation of emergency ratings as well as the high-loading failures, while all other
scenarios consider the emergency rating operations and the risk of cable failures
during high-loading operations. To evaluate the benefits of increased cable utilisation

89
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

under emergencies, these scenarios were compared against the base-case scenario
(BC).

Table 3-3 Modelling Scenarios for Emergency UGC Risk of Failure


Scenarios Model Variables Description of risk
Base Case No Emergency Ratings (ERs) Employed
Sc-1 α=0 β=0 ERs + no increased risk
high-load ERs

Sc-2 α=10, 20 ,30 β=0 ERs + unfavourable cable laying risk


failures

Sc-3 α=β=10, 20 ,30 Sc-2 + fixed annual lifecycle ageing


α=10, β≡(3.9) & Sc-2 + operation-based lifecycle
Sc-4
a=0.05, b=0.2 ageing (different ageing in cables)

Sc-1 utilises both LTE and STE ratings, considering that there was no increased risk
of emergency loading (i.e., α=β=0). Sc-2 considers only the additional risk of high-
loading failures, when a more critical cable design and unfavourable conditions (i.e.,
α≠0, β=0) is applied; Sc-3 advances the modelling by including cable risk index β that
utilises the pre-set values of cable criticality and the operator’s risk acceptability
factors. Sc-1, Sc-2 and Sc-3 utilise a constant value α and β for all network cables
within the cable life cycle loop (2nd SMC loop in Figure 3-1), while Sc-4 calculates the
cable risk index β within its life-cycle using its emergency operation history based on
(3.9) and the parameters provided (a=0.05, b=0.2). Sc-4 uses the second SMC loop to
compute the expected ageing in the life cycle. The cable design criticality index α in
Sc-4 remains constant and identical for all cables throughout the cable life cycle,
considering that the cable design and laying conditions do not vary with the increase
of cable age.

The computed overall network performance indices of base case, Sc-1, 2, and 3 have
been shown in Table 3-4. The benefits from the ERs implementation can be
distinguished by examining the results of Sc-1 to Sc-3 and comparing these against
the BC scenario.

Since there is no risk from ERs utilisation, the best network performance is achieved
in Sc-1. The comparison of these two scenarios indicates that ERs utilisation increase

90
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

the network’s loadability and flexibility under contingency operations, and thus
improves the whole reliability performance of the network especially when
considering a relatively low risk of high-loading failures. The BC scenario produces a
high EENS, which is greater than it in Sc-1 and Sc-2, but it has the minimum network
cable ageing due to no emergency loading allowed. The EFNF of base-case and Sc-1
were identical and the minimum across all scenarios, considering that no additional
ERs failures occurred in BC and Sc-1.

Table 3-4 Summary of network performance indices


Scenario EENS EFLC EDLC EFNF EIC EENATS EENASS
coefficient MWh/yr Occ./yr hr/yr Occ./yr M $ hr/yr hr/yr
BC NosER 73.34 2.89 15.18 20.88 1.96 161.97 412.40
Sc-1 α=β=0 30.15 1.12 6.07 20.96 0.77 164.47 488.94
α=10 31.16 1.17 6.32 21.15 0.80 164.72 495.32
Sc-2 α=20 36.47 1.33 7.36 21.67 0.94 165.37 511.18
α=30 57.99 1.87 10.95 22.82 1.54 167.03 550.23
α=β=10 37.48 1.36 7.51 21.75 0.98 165.46 512.42
Sc-3 α=β=20 110.75 3.10 19.34 25.37 3.01 170.46 640.73
α=β=30 429.37 6.67 47.61 30.34 11.92 178.22 824.77

When cable design risk (α) and the operator’s ageing acceptability risk (β) were
implemented, the gains from the ER utilisation in relation to the network reliability
performance declined. The results in Sc-2 and Sc-3 indicate reduced network
performance when the increased failure risk in emergency operations was
considered in the modelling. Due to the effect of ERs, additional high-loading failures
mainly occur on high-loaded cables, seriously worsening the network performance.
However, the network still produces lower EENS and EIC than in the BC scenario, due
to the increased flexibility from the ERs. Meanwhile, extra ageing occurs on the cables
when an increased rate of high-loading failures was considered. This is since a higher
failure rate produces more failures, and thus create more emergency loading events
in the remaining healthy cables. This added ageing may lead to higher maintenances
costs and more unexpected failures.

Network performances start to become worsen considering high values of α and β.


This could be the situation of implementing ERs on a cable with a poor maintenance

91
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

and long operation history. Under the circumstances, the ERs’ benefits related to
flexibility are entirely offset by the negative impact of the additional failures when a
high risk of high-loading failures is considered, resulting in higher interruption cost
(EIC) values than the base-case scenario.

After comparing Sc-2 with Sc-3, it is suggested that the index β plays a more dominant
role than the index α in relation to the occurrences of cable high-loading failures. This
suggests that realistic modelling is established since the ageing occurred in the
cable’s operating history should have a greater negative effect on its failure
possibility than the cable design and laying conditions. These results also indicate
that when the ERs are not frequently utilised in the previous history and no severe
ageing damage has occurred on the cables, then the ERs can increase network
flexibility and improve network reliability performance. However, an over-utilisation
of ERs would increase the cables’ ageing, insulation degradation and increase their
high-loading failure probability, and thus reduce the network resilience to
emergency events.

The comparison of the EENASS from the steady-state method and the EENATS from
transient-state method suggest that the EENASS has been greatly over-estimated.
Actually, power cables are normally loaded below their continuous current rating for
most of time with the exception of short-term high loading events under
contingencies. Ideally, cables have a longer life than designed especially when the
thermal inertia of the surroundings and the transient behaviour of temperature
variation is considered. Although the EENATS does not increase greatly when ERs are
utilised, there is a much higher probability for cable insulation to break down under
high-loadings, thus the risk of high-loading failures needs to be considered related to
the asset management of cables.

This result for EENA also indicates that the thermal inertia cannot be neglected by
utilities if they want to obtain an accurate estimation of cable thermal ageing
conditions. This can also help utilities to achieve maximum benefits and more
realistic cable ageing costs. The variation of EENATS is much smaller than EENASS

92
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

when ERs are implemented. This suggests that the employment of ERs may lead to
less impact on cable ageing than expected in SS method.

Sc-4 was performed within the 2nd SMC loop, considering a cable design lifetime of
30 years. Figure 3-5 and Figure 3-7 compared the EENS and EFNF outputs from Sc-4
and the BC. The recorded indices varied along with the increase of ‘year in service’
and cable age. However, the ageing and network performance in the BC scenario was
constant throughout the cables’ life, considering that the risk of high-loading failures
remained zero and unaffected by the cables’ operation history and age.

250 50
EENS (Sc-4)
EENS (Base case)
200 40
EFNF (Sc-4) Sc-4

EFNF (Occ./Yr)
EFNF (Base case)
EENS (MWh)

150 30
Sc-4
100 Base case 20
Base case
50 10

0 0
0 5
10 15 20 25 30
Cable Life Year YL (Yr)
Figure 3-5 EENS, EFNF of Sc-4 and base case across a cable design life of 30
years

Figure 3-5 suggests an exponential increase of EENF (more cable failures occur)
when the cables become older. This is because the index β become larger as the cables
get older and build up their emergency operation history. Thus the high-loading
failure rate becomes higher and produces a higher possibility of failure at higher
cable ages. The impact of cable age and the index β ‘s growth can also be shown by
the EENS result. The initial EENS in Sc-4 is lower than base case at early stage of cable
life. This indicates higher overall network reliability due to the additional flexibility
by implementing ERs. However, the EENS in Sc-4 increases along with the increase
of cable age and become equivalent to the EENS in BC in year 23. This is also because
of the continuous growth of the index β as the emergency operation history builds up.

93
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

The EENS of Sc-4 becomes higher than BC from age 23 and substantially lower EENS,
EFLC and EDLC outputs of BC are observed than Sc-4. Those results suggest that no
gains can be achieved through the implementation of ERs in an aged cable network
with a long emergency operating history. Hence, utilities should consider those
factors on ERs implementation, cable replacement and maintenance methods.

180 800
TS (Sc-4)
TS (Base case)
175 SS (Sc-4) Sc-4 600

EENASS (hrs/Yr)
EENATS (hrs/Yr)

SS (Base case)

170 Base case 400

Sc-4
165 200

Base case
160 0
0 105 15 20 25 30
Cable Life Year YL (Yr)
Figure 3-6 EENASS and EENATS of Sc-4 and base case from a steady and
transient state thermal model across a cable design life of 30 years

Figure 3-7 shows the EENA of the base case and Sc-4 calculated using transient and
steady state thermal modelling across a cable life period of 30 years. The total
network cable ageing increased exponentially in Sc-4 when the cables’ age become
larger. This is because more failures occur on older cables, resulting in more high-
loading/emergency operation, and thus more ageing. On the other hand, the EENA of
BC kept the lowest value because no ERs were utilised at all. The comparison between
EENASS and EENATS indicates that EENASS was an greatly inflated value where the
steady-state model is implemented. The difference between the over-estimated
EENASS and more accurate EENATS becomes larger along with the increase of cable
age. This is because intermittent emergency loadings under a short duration leads to
wild fluctuations of conductor temperature as estimated from the SS model. This
steady-state method produced four times more ageing than what was more
accurately estimated with the transient-state method. As a result, the EENASS result
could mislead utilities on the nature of the cables’ ageing assessments based on their
operation history.

94
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

Based on the simulation results for individual cables, cables C10, C11 and C13 were
considered to be more “critical”, because these cables have more frequent and longer
ER operation compared to the other ‘low loaded’ cables. Figure 3-7 shows the cables’
emergency operating duration (EDEL) within the cables’ life cycle. As indicated in
Figure 3-7, the EDELs of those critical cables are much longer than others, indicating
that the “critical” cables tolerate a much longer duration of emergency loading than
the other cables. This results in substantially more thermal ageing and a higher
failure risk than in the other cables. Therefore, the output of EDEL provides a good
reference index for utilities to rank their cables, considering the network and cable
design, and the operator’s preferences on ER utilisations. For example, C11 was
identified as the most critical and heavily loaded cable within the test 14-bus network.

35
C1 C10 C11 C13
30
25
EDEL (hrs/Yr)

20
15
10
5
0
0 5 10 15 20 25 30
Cable Life Year YL (Yr)
Figure 3-7 EDEL results of C10, C11, C13 cables throughout their life cycle.

Figure 3-8 shows the frequency of cable failures (EFCF) throughout their complete
cable life. It shows that both the EDEL and EFCF of the cables, C10, C11, C13, increases
exponentially because the frequency of high-loading failures increases along with the
increase of cable age, while other cables, such as C1, do not fail so frequently.

The frequency of cable failures and emergency operation durations provides


network operators and utilities with informed suggestions on emergency rating
utilisation, cable replacement and maintenance. Cable actual performances can be
used as a cable replacement and maintenance criterion, instead of ‘years in service’.
For example, if a maximum boundary of EDEL is set to be 20hrs/yr for all cables, then

95
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

the network operator needs to reduce the emergency rating utilisation and duration
of C10, C11 and C13 after years 22, 28 and 30 in order to avoid excessive ageing and
overloading on the cable. If the maximum annual cable failure rate EFCF is set to be
2.5occ./yr, then suitable replacement and maintenance actions need to be
undertaken to prevent more failures occurring in aged cables, C10, C11 and C13 after
year 24, 27 and 28.

4
C1 C10 C11 C13
3.5
3
2.5
EFCF (Occ./Yr)

2
1.5
1
0.5
0
0 5 10 15 20 25 30
Cable Life Year YL (Yr)
Figure 3-8 EFEL results of C10, C11, C13 cables throughout their life cycle.

More importantly, the cable’s lifetime projection can be indicated by these indices.
Hence, the EDEL and EFCF can predict the optimal maintenance and replacement
year based on the cables’ operating condition and the network performance. By
implementing the methodology and obtaining the output, utilities can receive
suggestions about their cables’ health condition, thus allowing them to design ERs
utilisation.

In addition, utilities are able to correlate these simulation-based indices with


condition monitoring data of their equipment. Furthermore, it can also be indicated
that which cables are under-utilised, allowing operators to implement different
operating strategies for them. Capturing these component indices in an actual
network would be very insightful, improving overall network component
performance.

96
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

From the case study in the 14-bus network, it was suggested that EECA gives a
comparison of estimated ageing for cables in a given network. Therefore, network
operators can classify their cables into three main groups considering their high-
loading risks:

 Accelerated ageing cables indicate a group of cables that experience frequent


overload events leading to the potential reduction of their lifetime and an increase
in the failure rate (e.g., C11). Network operators are suggested to conduct an
extensive inspection & maintenance.

 Infrequent high-loaded cables indicate a group of cables that experience


infrequent emergency loading operations, which may also affect their high-
loading failure rate (e.g., C10, C13). These cables also require inspection,
monitoring and maintenance.

 Under-utilised cables indicates a group of cables that would be expected to be


disturbed mainly due to external condition changes (digging activities or
construction work) and no immediate check is necessary, for example, C1.

A novel methodological advancement based on the SMC-based network reliability


evaluation framework has been proposed in this chapter. It integrates different
existing methods, including a cable thermal-electric equivalent (TEE) model and an
ageing model and failure model within a double SMC simulation, which allows for the
calculation of increased risk due to ageing and overloading. This advancement
addresses the key question of how to quantify the risks related to the current
increasing utilisation and flexibility challenges that utilities face.

The main contribution of this innovative methodology is to capture the additional


risks associated with cables that are operated beyond their limits through three
additional Markov states. This risk has not been considered in any previous work.
These additional states (and the proposed model) will inform utilities about the
benefits and risks of implementing emergency operating states and help them to
make decisions related to optimising their operational flexibility based on the

97
Chapter 3: Impact of Cables’ High-loading Risks on Network Reliability

suitability of their network topology. The equations related to the additional


component’s operational Markov states are novel, and they help to link the effect of
health indices into network assessments, which is a known problem in the industry.

The methodology will help utilities to identify their critically aged cables (from their
actual loadings and not from their ‘years in service’) in a common ageing base. Hence,
they can compare the recorded failure rates and health indices with the ageing
outputs of the method to cluster cables in similar health groups. From the study
presented using the 14-bus RTS critical network, cables can be easily distinguished
based on the proposed EDEL and EFCF indices. In particular, the C11 cable is three
times more critically loaded compared to the second critical cable, C10.

All of the results were highly affected by the proposed cable design, criticality index
α, and the operator risk acceptability index β. A further correlation with the data
derived from the experiments and recorded condition monitoring data is necessary
to be explored in order to obtain the appropriate values for α and β. This is beyond
the scope of this chapter, but the proposed method allows for such correlations to be
implemented. Integrating this methodology with future demand forecasts can greatly
help planners and operators to benefit from most of their assets by increasing both
network reliability and utilisation.

Equation Chapter (Next) Section 1

98
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

4 Impact of Extensive Emergency Ratings on


Network Reliability and Plant Ageing

LTE and STE ratings are designed to allow OHLs and UGCs to sustain moderate
loading congestions following one or more network contingencies. The time
duration criteria was implemented based on industry standards and utilities’
experience, which has not been included in most existing reliability studies for
transmission networks. Furthermore, the increased risks associated with
emergency rating (ER) operations has not been quantified.

This chapter introduces a methodological improvement based on the SMC-based


network reliability evaluation framework that quantifies the effect of emergency
rating operations on network-wide performance and equipment ageing. The
modelling incorporates the utilities’ emergency operating procedure, such as
hierarchical ER operating levels and corresponding ER time duration criteria, into
the network’s reliability modelling. The methodology also integrates a fault chain
search model to capture the circuits’ cascading failures which is generated by
severe overloads. Thermal-ageing modelling of both OHLs and UGCs has been
integrated to calculate the ageing at ER operations. This study performed a cost-
worth evaluation to compare the potential reliability benefits from extended ER
durations and increased ER ampacities against the increased cost of plant ageing
and from cascading failures.

The methodology was applied on the IEEE 14-bus and 24-bus network to test its
scalability. This case study indicated an effective 30% reduction of load curtailment
in the 14-bus network (17% reduction in 24-bus) with the implementation of
extensive emergency ratings compared to the standard practice. The proposed
method allows utilities to quantify the increased flexibility provided by extensive
emergency ratings against the corresponding operational risks and costs, by
applying this methodology to their specific network topology.

99
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

Power utilities seek economically competitive solutions to satisfy the continuously


growing power demand, and to enable a high penetration of renewable energy
integration. This requires system operators to optimally operate their existing
networks in a more resilient and flexible way, particularly in emergencies. To achieve
these goals, utilities employ feasible methods to increase the network’s power
transfer capability under contingencies or during critical loadings events.

The traditional option for transmission networks is to allow the emergency ratings
(ERs) of circuits to tolerate moderate line overloads for a finite period of time at the
cost of accelerated thermal ageing. This provides operating flexibility, allowing
operators to perform corrective actions to mitigate the overloads. Another way is to
utilise time-varying line ratings to increase adequacy, which requires real-time
monitoring instrumentation. A great deal of effort was devoted to explore the concept
and implementation of a Dynamic Thermal Line Rating (DTLR) system and Cable
Monitoring and Rating system (CMARS). However, as of today, these concepts have
failed to materialise because their deployment makes very little business sense and
the systems are impractical to implement and operate.

The explicit ERs operating levels, i.e. long and short-term emergency ratings, and
their duration criteria have been defined in operating orders that need to be followed
by operators. The traditional practice using the ERs duration criteria was established
based on industry standards [84-86] and implemented by TSOs [22, 87] in early
1970s. The operating procedures and duration criteria defined in emergency control
practices [84] and the utilities’ emergency operation manuals [23, 75, 76] have been
stated below.

a. “When a circuit becomes loaded above its STE rating, immediate corrective
actions, including voltage reduction and load shedding, are taken to reduce the
loading to below the STE rating within 5 minutes and furthermore, to below its
LTE rating within 10 minutes.”

100
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

b. “When a line is loaded above its LTE rating but below the STE rating, corrective
actions are taken to return the loading to the LTE rating within 15 minutes.”

c. “After the loading has been reduced below the LTE rating, additional corrective
action is taken to reduce the loading to below its normal continuous (Con) rating
within 30 minutes of the initial overload.”

d. “When a line has been loaded for four continuous hours (or a longer period as
may be established by the rating authority) above its normal continuous rating
but below the LTE rating, corrective action is taken to return the line to its normal
continuous rating.”

e. “When the amount of unscheduled power flow across the overloaded


transmission facility drops to 90 percent of its transfer limit (normal continuous
rating), all requests for assistance should be terminated.”

However, this ERs practice is conservatively and loosely defined with a great
reserved margin to prevent severe thermal damage to transmission facilities. On the
other hand, it take a longer time for the operators to conduct the correct actions
considering the complexity of the current transmission network, as well as the
uncertainty and variability of high penetration renewables. This forces the utilities to
update their ERs policies to a more flexible and extensive ERs operating strategy.

However, the increased risk of thermal ageing of the transmission OHLs and UGCs
during emergency rating operations has only been investigated in limited works [11,
58, 61, 62]. These works also fail to incorporate detailed ERs operating procedures,
such as explicit ERs operating levels and their duration criteria, into their emergency
operation modelling. These research did not consider the occurrence of cascading
failures and their severe impact on network integrity.

In conventional SMC-based network reliability assessments, the occurrence of each


component failure is independent from another considering their individual outage
rate. This failure generation process neglects that a sudden power flow re-
distribution after a major disturbance may lead to a cascading failure in another
circuit (or more than one) in the network [67, 88, 89]. Many reliability studies assume

101
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

that corrective generator re-dispatching and load shedding actions are completed
immediately right after the occurrences of contingencies, neglecting post-
contingency overload operations of the power equipment.

Emergency rating operations of UGCs generate emergency loading temperatures


after a significant time delay due to their thermal inertia, leading to thermal ageing
and the life loss of cables [54]. Exceeding the design temperature on OHLs also results
in an accelerated ageing and loss of tensile strength in the conductors [58]. The UGCs’
temperature variation is not an immediate response to the power flow changes due
to their large thermal inertia. This thermal inertia could allow cable circuits to gain
heat during the emergency rating operations, and to dissipate heat during normal
operation periods. Therefore, this provides UGCs with a greater potential of
overloading at the cost of thermal ageing, compared to OHLs.

To fill in the research gap and to provide utilities with quantified results, this chapter
has developed a methodology to assess the impact of different ERs operating
durations and the ampacities of UGCs and OHLs on network adequacy and flexibility.
The thermal-ageing modelling of OHLs and UGCs was implemented to compute their
thermal ageing as a results of ER operations. The methodology also includes a fault-
chain search model to capture any potential cascading failure caused by severe
overloads of circuits. These two additional models allow for us to calculate the
increased probability of cascading failures and accelerated plant ageing when the
extensive ERs strategy is implemented. This work can also allow utilities to make
corrective decisions to identify optimal ERs ampacities and operating durations for
their UGCs and OHLs. Due to the relatively slow computing time, this methodology is
mainly applied in emergency operation planning, rather than real-time decision-
making.

Section 4.2 introduces the detailed modelling of network reliability evaluation


combining the modelling of cascading failure, ER operation and thermal ageing
computation. Section 4.3 is the implementation of the methodology and Section 4.4-
4.6 discuss some of the main results and findings.

102
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

4.2.1 Outline of the Proposed Methodology

The methodology considers the cascading failure and thermal ageing of circuits
within SMC-based network reliability evaluation framework as described in Chapter
1. In this way, the negative impact of cascading failures and the risk of circuits’ ageing
when ERs are implemented can be accurately captured. The emergency operating
procedures of the transmission OHLs and UGCs was modelled with explicit ERs
operating levels and time duration constraints in the ‘Post-contingency operation of
OHLs and UGCs’ block. The flowchart of the methodology is shown in Figure 4-1.

Network PLANT
Reliability Data Operational Data Weather Data Branch Data

Power System Data Initialization

Initial Network Component Status Mapping

Cascading failure model


Annual Time step
analysis Post-contingency operation of circuits analysis
〈Y=Y+1〉 〈tΔt=tΔt+Δt〉
Thermal-ageing model of OHLs and UGCs
NO
Annual computation is completed Inner SMC-loop
YES
Annual analysis is completed until specified Cov Outer SMC-loop
NO
YES
Network reliability & Branch performance indices calculation

Completed network reliability evaluation


Figure 4-1 Flowchart of the network evaluation integrating cascading failure,
emergency rating operation and circuit ageing

The network data is comprised of the failure λx, and repair rates µx, and the system’s
operating constraints as defined in Section 2.2. Furthermore, additional current
rating constraints for each UGC or OHL were defined, including the normal
continuous (Con) rating Icon, long-term emergency (LTE) rating ILTE, short-term
emergency (STE) rating ISTE, drastic action limit (DAL) IDAL, and their corresponding
duration criteria.

103
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

Similarly to the framework described in Chapter 2, system analyses were conducted


through two iterative SMC loops. The inner SMC-loop was performed for each Δt time
step in every Y, to determine the changes of network operating condition, as a result
of component failures, branch rating variation, weather changes and demand cycling.
The outer loop performed the annual reliability computation and repeated until the
target covariance (Cov) of EENS was achieved or the simulation was completed at the
maximum number of simulation years, N [3].

In the initial network component status mapping block, the operating and repairing
status for each network component x, was mapped for every Δt of year Y, based on a
two-state Markov method, as described in Section 2.3. In the ‘cascading failure model’
block, a fault-chain search method was utilised to capture any potential cascading
failure events of the circuits triggered by a generic failure of another network circuit.
These cascading failures are considered to be an immediate protection response to
severe line overloads, which leaves no time for operators to take any action. The
search loop was iterated until the potential cascading failures had been fully captured
and the repair and maintenance time was updated into the status mapping.

After the severe line overloads were mitigated, the network may still need to operate
at post-contingency conditions for a finite period of time. This period of time was
modelled within the ‘post-contingency operation of circuits’ block, where appropriate
rating was implemented subject to the defined ERs operating levels and duration
criteria, for each network branch b, ∀b∈Nb, with Nb being the total number of network
branches. Then, a network optimization calculation was executed via an AC OPF to
perform the correct generation re-dispatching and load shedding actions with aim to
minimise the total generation and load curtailment costs. This AC OPF based network
optimisation has been described in Section 2.4. After the results from the network
optimisation was obtained, the calculated current flows were then used to update the
emergency loading durations for each branch, b. The whole post-contingency
operation process has been described in Section 4.2.3.

The thermal-ageing model of OHLs and UGCs calculated the circuit’s conductor
temperature θC,b(tΔt), based on its current flow, Iflow,b, the weather data of this Δt and

104
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

the circuit design properties. The potential thermal ageing for each branch, b,
occurred in the ER operations and was also calculated using the method stated in
Sections 2.5 and 2.6.

Some appropriate indices were recorded, for example, EENS, EFLC, EDLC, and EENA
with their calculating equations stated in Section 2.7. The Expected Frequency of
Circuit Cascading Failure (EFCCF) was proposed to capture the expected annual
frequency of cascading failures of UGCs or OHLs, as described in (4.1), where NCCFb,Y
is the number of circuit cascading failures of branch b in year Y.

NY
EFCCFb   NCCFb,Y NY (4.1)
Y 1

A novel network index, expected ageing and interruption cost (EAIC), was used to
capture the combination of the operational costs related to plant ageing and the
interruption cost of load curtailments. The EAIC was calculated in (4.2), where EIC is
the expected interruption cost index, EECAi is the expected equivalent ageing of cable
i, EELAj is the expected equivalent ageing of line j which was calculated in Section 2.7,
CostUGCageing and CostOHLageing were the costs of the equivalent ageing for UGCs and
OHLs respectively in £/hr.

NUGC NOHL
EAIC  EIC   Cost
i 1
UGCageing  EECAi   Cost
j 1
OHLageing  EELAj (4.2)

4.2.2 Cascading Failure Modelling

Cascading failures may occur in transmission circuits because of severe line


overloads which is generated from the sudden power flow re-distribution after a
major disturbance, resulting in its overload protection being triggered within a few
seconds [53].

Figure 4-2 describes the computational steps implemented in the cascading failure
search model in order to simulate the chain of cascading failure events. In this study,
the DAL rating was defined as the current constraint at which immediate overload

105
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

tripping will be triggered to protect equipment from further damage. Exceeding the
DAL rating can trigger the occurrence of a cascading failure, indicating that circuit b
will be instantaneously tripped once its loading is above its DAL rating, IDAL,b.

As shown in Figure 4-2, the initial network component state mapping helps to indicate
if there is any new component failure occurring at the present time tΔt. If there is, an
AC OPF was run based on the pre-contingency network status (excluding the new
contingency) in order to determine the pre-contingency system’s operating state.
After that, an AC power flow was conducted based on the fixed generators’ outputs
and load consumption extracted in the pre-contingency state. This AC PF helped to
determine the re-distributed power flows within the remaining healthy lines when
the new failure occurred. Fixed pre-contingency generation and demand were used
in AC PF because severe overload immediately triggers the circuit’s protection and
leaves no time for the operators to perform corrective generator re-dispatching and
load shedding. Therefore, this study assumed that the generator outputs and load
consumptions were unchangeable during the sudden power flow re-distribution at
the contingency moment.

Initial network component status mapping


NO
Any new network contingency occurs at tΔt
YES
Run AC OPF based on the pre-contingency
network state and constraints

Run AC power flow


Trip branch y regarded Any overload branch
as cascading failures z YES b Iflow,b>IDAL,b
NO
Branch status mapping update considering recovery
time TTRz,b of cascading failure z

Failure chain ends


Figure 4-2 Flowchart of fault chain search model to capture cascading failures

If the calculated flow Iflow,b, of any branch b, exceeds its DAL rating IDAL,b, then the
branch was tripped by the overload protection, and considered as a cascading failure

106
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

of branch b. Then from present tΔt, the repair time TTRz,b of cascading failure z is
considered into the status mapping. After that, another ACPF was conducted to find
any additional cascading failures due to the occurrence of the previous cascading
failure. The loop iterated until no overloaded lines remained and all potential
cascading failures were captured. The repair time of all cascading failures is much
shorter than an actual fault, in this study, the repair duration TTRz,b for the cascading
failure z, of branch b, was considered to be three hours (TTRz,b=3hours) in both OHLs
and UGCs.

4.2.3 Modelling of the Post-contingency Operation of Circuits

Four current ratings are considered for OHLs and UGCs, including; DAL, STE, LTE and
Con, denoted as IDAL, ISTE, ILTE, and Icon. Seasonal ratings were not considered in this
study. Four distinct levels of current ratings have been described in Table 4-1, with a
different sequence of control actions.

Table 4-1 Types and Levels of Emergency Rating


Loading Conventional
ER Level Control Actions Comments
Range Time Available
line tripping
Cascading
Severe > IDAL 1s shed load
failure
generator re-dispatch
generator re-dispatch Neglected in
DAL IDAL - ISTE 5mins
shed load methodology
Moderate

generator re-dispatch
STE ISTE - ILTE 15mins Limited by DSTE
shed load if necessary
4hrs(winter) generator re-dispatch Limited by
LTE ILTE – Icon
12hrs(summer) shed load if necessary DLTE_S or DLTE

Severe overload level indicates any circuit loading above IDAL, immediately triggering
the protection and leading to its cascading failure. STE overload level indicates the
circuit operates between ISTE and ILTE, LTE overload level means between ILTE and Icon.
According to the operating standard, any loading between ISTE and IDAL should be
reduced to below ISTE within 5 minutes. However, because the transition step is set to
be Δt=15min, this 5-min period was neglected and integrated into the STE operation
by assuming the loading was reduced to below ISTE within a negligible time.

107
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

Based on the existing practice, three unique constraints were defined for ER
operating durations.

 DSTE is the maximum allowable time duration for OHLs or UGCs to operate within
the STE level before corrective actions are taken to reduce the loading to ILTE.

 DLTE_S is the maximum allowable time duration for OHLs or UGCs to operate within
the LTE level after the loading was reduced from above to below its ILTE, before
additional corrective actions are taken to further reduce the loading to below Iconl.

 DLTE means the maximum allowable time duration for OHLs or UGCs to operate
within the LTE level before corrective actions are taken to reduce the loading to
below Icon where the line or cable is never loaded above its ILTE.

Updated network status at time tΔt


Line Rating Selection

NO
Operators utilize ILTE & ISTE on branch b
YES
NO
System in emergency operation
YES
NO
STE-durationb ≥ DSTE
YES
NO
LTE-durationb ≥ DLTE_S
YES
YES
Irated,b(tΔt) = Icon STE-durationb ≥ DLTE
NO
Irated,b(tΔt) = ILTE Update Irated,b(tΔt) Irated,b(tΔt) = ISTE
Optimal AC power flow
Emergency Loading Duration Update
YES
ISTE ≥ Iflow,b > ILTE STE-durationb + Δt
NO
YES
ILTE ≥ Iflow,b > Icon LTE-durationb + Δt
NO
NO
Iflow,b ≤ 0.9×Icon No update
YES
STE-durationb = LTE-durationb = 0
Figure 4-3 Flowchart of the ‘post-contingency operation of circuits’ block

Figure 4-3 describes the computational steps of ‘post-contingency operation of circuits’


block in Figure 4-1 which has been divided into two blocks: line rating selection and
emergency loading duration update. In the line rating selection block, when ERs were

108
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

not utilised or no contingency occurs, the rating Irated,b(tΔt) of branch b at time tΔt was
set to a continuous rating Icon. The LTE-durationb and STE-durationb was utilised to
record the total emergency loading duration of branch b, to operate within the LTE
and STE levels, respectively.

Based on the assigned ER for each circuit and updated network status, an AC OPF was
performed with its optimisation calculation has been described in Section 2.4. After
optimal power flow, the LTE-durationb and STE-durationb need to be updated to
include any emergency loading time at this Δt time step.

Based on the network operational results, the current flow Iflow,b of branch b was
calculated. If Iflow,b is between ISTE and ILTE within STE overload level, then the STE
emergency loading duration was updated as STE-durationb=STE-durationb+Δt; If Iflow,b
was between ILTE and Icon within LTE overload level, then the LTE emergency loading
duration was updated as LTE-durationb=LTE-durationb+Δt. This indicates that the
corresponding emergency loading duration was updated by increasing by one time
step Δt. If Iflow,b is between 0.9∙Icon and Icon, then no update was considered for the
emergency loading durations. However if Iflow,b is lower than 0.9∙Icon, then STE-
durationb and LTE-durationb, were re-initialized to zero, indicating that the previous
ER operation finishes and thus all corresponding control actions taken were
terminated. The Iflow,b<0.9∙Icon is selected as the ER operation termination criterion,
not Iflow,b<Icon, because the 0.9∙Icon is selected as the criterion to terminate all request
of assistance in emergency control practices [84].

Figure 4-4 demonstrates an example of component status and current-flow transition


mapping in the time domain for branch b=1, ,2 ,3 and 4. As seen in Figure 4-4, the
normal line failure (NLF) of branch b’=1 leads to a sudden severe overload of branch
b’’=2, which results in a cascading failure (CF) of branch b’’=2 immediately. The NLF
and CF were modelled to occur simultaneously, but lasted for different durations.
Branch b’’=2 was disconnected from the network for a duration of TTRZ,2, and was
back in service afterwards.

109
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

(a) Exact state transition mapping for branch 1, 2


b'=1 NLF TTRN,1

b"=2 CF TTRZ,2
(b) Current-flow transition mapping for branch 1, 2, 3, 4
IDAL
ISTE
DSTE
ILTE D LTE_S
DLTE
Icon
Iflow,1
Iflow,2
Iflow,3
Iflow,4
Time axis
Figure 4-4 An example of modelling the state transition and corresponding
current flows

The normal line failure (NLF) of branch b’=1 also leads to moderate overloads on the
other two branches (i.e. branch 3 and 4) which are allowed to operate at emergency
ratings for certain loading durations. For example, branch 3 is allowed to operate
within the STE level (between ISTE and ILTE) for a duration of DSTE, before actions were
taken to reduce its current flow Iflow,3 to below ILTE. After that, time constraint DLTE_S
was used to further limit branch 3 to operate within its LTE level (between ILTE and
Icon). If the LTE emergency loading duration, LTE-duration3, of branch 3, exceeded
DLTE_S, then its current flow Iflow,3 was reduced to below Icon.

Unlike branch 3, branch 4 only operates within LTE level, but never enters the STE
level. Therefore, an increased time duration of DLTE allows branch 4 to operate within
the LTE level for longer duration than branch 3, before Iflow,4 was reduced to below
Icon. If all three duration constraints, DSTE, DLTE_S and DLTE, are not met, then the branch
rating Irated,b was set to ISTE, in order to enable the maximum flexibility and loadability
under ER operations.

The methodology was demonstrated on the 14-bus network (shown in Figure 3-4)
based on the default demand and generation capacity. It was also demonstrated on

110
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

the modified 24-bus network [58] considering a 1.6 pu demand and generation
capacity increase. The schematic diagram of IEEE RTS-96 24-bus network has been
shown in Figure 4-5. Only branches are allowed to have failures and other network
components are modelled to have a zero failure rate (100% reliable). All network
branches were modelled by either UGCs or OHLs in both the 14 and 24 bus network.

30 32 38
33 31
29 35
28 37
34
36

24
25 26 23
22
21

19
27 18
20

7 14
15 16 17
6
12 13 10
8

9
2
3
4 5
11
1

Figure 4-5 Single line diagram of IEEE RTS-96 24-bus network

4.3.1 Test Network – 14-bus, 24-bus OHL & UGC network

The original 14-bus data does not include the rating and design data of the circuits.
Therefore, they had to be retrieved from the manufacturers’ information, as detailed

111
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

in Table 4-2 [79, 80]. The aluminium conductor Jaguar and the two XLPE cables as
shown in Table 4-2 were used for the 14-bus network study. The cables were directly
buried in a flat formation with a separation of 0.1m at 0.8m and 1.5m depths for
13.8kV and 69kV respectively. The LTE and STE ratings were computed using the
proportions (ILTE/Icon, and ISTE /Icon) of the 132kV cables and the overhead lines in
RTS-96 [74].

The IEEE 24-bus network data included line limits but no branch design data, thus a
suitable branch design was selected as stated in Table 4-2. The 230kV and 138kV
cables were laid in a flat formation with a 0.15m separation and a 1.2m depth. All of
the cable thermal ratings for both 14 and 24-bus networks were computed based on
15°C soil temperature, 0.9km/W soil thermal resistivity, and 1.9×106J/m3K volume
specific heat, which are the default values in ENA P17 [17].

Table 4-2 Branch Electrical Properties Modelled in the 14 and 24-bus


Networks
Test Branch Size R90°C* Icon ILTE ISTE IDAL
Network Type mm2 Ω/km A A A A
Jaguar line 222 0.1655 561 665 702 842
14-bus 69kV XLPE cable 185 0.1273 538 593 614 1076
13.8kV XLPE cable 185 0.1273 574 633 656 1148
230kV Araucaria line 1093 0.00005 1272 1503 1567 1908
138kV Upas line 479 0.00011 732 873 923 1098
24-bus
230kV XLPE cable 1200 0.0618 1130 1246 1291 2260
138kV XLPE cable 400 0.0237 715 788 817 1430
*For OHLs, R90°C become R75°C as their design temperature is set to be 75°C

The Arrhenius model parameter Eai/k=12430 was used to define the XLPE cable type
[54]. The OHL ratings were computed according to the 20°C ambient temperature,
0.5m/s wind speed and nil solar radiation [71]. Considering that UGCs are designed
to operate at 90°C for 30 years, its ageing cost was calculated to be £3.75/hr∙km.
Considering that OHLs are allowed to operate at 100°C for around 2000hrs, its ageing
cost was calculated to be £170/hr∙km [90].

112
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

The length of each circuit, ℓb, were derived based on the original impedance data and
the 0.7 Ω/mile factor [82]. Therefore the failure rate and repair duration of every
branch could be calculated using (4.3)-(4.4) and OHLs using (4.5)-(4.6) where ℓb is in
miles [81].

i  0.0062  i  0.226  occ./ yr  (4.3)

Di  1.5  i  11.5  hr / occ. (4.4)

 j  0.0052  j  0.22  occ./ yr  (4.5)

D j  10  hr / occ. (4.6)

Table 4-3 Branch outage rate and duration in IEEE 14-bus network
UGC outage UGC outage OHL outage OHL outage
Line Length
rate, λi duration, Di rate, λj duration, Dj
No. miles
occ./yr hours/occ. occ./yr hours/occ.
1 8.47 0.279 24.2 0.264 10
2 8.47 0.279 24.2 0.264 10
3 15.61 0.323 34.9 0.301 10
4 13.84 0.312 32.3 0.292 10
5 12.63 0.304 30.4 0.286 10
6 12.44 0.303 30.2 0.285 10
7 12.49 0.303 30.2 0.285 10
8 3.00 0.245 16.0 0.236 10
9 0.60 0.230 12.4 0.223 10
10 0.77 0.231 12.7 0.224 10
11 0.40 0.228 12.1 0.222 10
12 0.25 0.228 11.9 0.221 10
13 0.81 0.231 12.7 0.224 10
14 0.57 0.230 12.4 0.223 10
15 0.81 0.231 12.7 0.224 10
16 1.05 0.233 13.1 0.225 10

The soil temperature measurement profile of 2015 obtained from the ERA-interim
dataset provided by ECMWF was utilized [83]. The weather data for OHLs was
obtained from the Whitworth Meteorological Observatory [91]. The time-varying
ambient and soil temperature data used for the simulation year has been shown in
Figure 4-6.

113
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

35
Ambient Temp Soil Temp
30
25
Temperature (°C)

20
15
10
5
0
-5
0 2000 4000 6000 8000
Simulation Hours [Hrs]
Figure 4-6 Time-varying ambient and soil temperature in a simulation year

4.3.2 Modelled Scenarios

The proposed methodology was applied to both the 14-bus and 24-bus networks in
order to investigate the potential values and increased risks under multiple extensive
ERs scenarios, as described in Table 4-4.

Table 4-4 Modelling Scenarios of the Emergency Rating Operations


Modelling DSTE DLTE_S DLTE ER uprating
Comments
Scenarios hr hr hr factor pu
Base case (BC) 0 0 0 0 No ERs Employed
Sc-1 0.25 0.25 4 U Utilities’ practice
Sc-2 0.5 0.5 8 U 1. Durations are increased
Sc-3 1 1 16 U by 2, 4, 8, 12, 16 times
Sc-4 2 2 32 U 2. ERs are increased by U,
Sc-5 3 3 48 U up to 1.2pu for OHLs and
Sc-6 4 4 64 U 1.5pu for UGCs

These scenarios were compared against the Base-case scenario (BC) to quantify any
improvements from the ERs utilisation. Sc-1 employed both LTE and STE ratings
subjected to conservative operating duration criteria based on the current utilities’
practices. Sc-2 to Sc-6 considered increased allowable ERs durations where three ER
durations, DSTE, DSTE_S, DLTE, were doubled in Sc-2, and increased by 4, 8, 12 and 16
times respectively in Sc-3, 4 ,5 ,6, compared to those in Sc-1. The extended ER
durations help to indicate their potential impact on the network performance and the
thermal ageing of circuits. An ER uprating factor U was applied for the extensive ERs
scenarios (Sc-1 to Sc-6), considering that the ampacities of ILTE and ISTE were

114
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

increased by U times. The ERs, ILTE and ISTE, were allowed to be increased up to 1.5pu
for UGCs, while it was only up to 1.2pu for the OHLs. This was due to their low
tolerance to emergency temperatures. It is worth noting that the Icon and IDAL rating
remained constant and the recovery time of the cascading failures was set to be three
hours, TTRZ,b=3hrs, for both UGCs and OHLs in all of the ER scenarios.

The impact of the extended ER durations and increased ER levels of the OHLs on the
OHL network reliability performance has been in Figure 4-7.

180 14-bus 3300 24-bus


Base case U=1.0 U=1.1 U=1.2
Sc-1
160 Sc-2 3250
Sc-3
EENS (MWh)

Sc-4 Sc-5
Sc-6
140 3200

120 3150

100 3100
0 1 2 3 4 0 1 2 3 4
DSTE (hrs) DSTE (hrs)

95 14-bus 60U=1.1 24-bus


U=1.0 U=1.2
90 55
EDLC (hrs/Yr)

85 50

80 45

75 40

70 35
0 1 2 3 4 0 1 2 3 4
DSTE (hrs) DSTE (hrs)
Figure 4-7 EENS, EDLC output in the time domain of DSTE for the 14-bus (left)
and 24-bus OHL networks (right)

Figure 4-7 illustrates the amount (EENS) and duration (EDLC) of the load curtailment
within the modelling scenarios in the 14-bus (left) and 24-bus (right) networks. The
results have been displayed in the time domain of DSTE, thus the base case scenario

115
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

was shown at DSTE=0hr, and the other ER scenarios are located at their corresponding
DSTE, e.g. Sc-1 is displayed at DSTE=0.25hr. The ER scenarios (Sc-1 to Sc-6) considered
three different ER uprating factors U=1.0 (black), U=1.1 (red) and U=1.2(green), as
marked by three dots in Figure 4-7.

The comparison between Sc-1 and BC suggests a significant improved network


performance in both the 14-bus and 24-bus network when ERs are implemented.
This improvement is because the network loadability and flexibility is greatly
enhanced through the implementation of ERs. ERs provide 20%-25% higher current
rating during emergency operation after contingencies. Therefore, this increased
flexibility loosens the network loading constraints, and thus reduces the load
curtailments, e.g. EENS and EDLC, during emergency operations.

Sc-1 utilises the time criterion of emergency loading durations as defined by the
transmission utilities, thus Sc-1’s result at U=1.0 indicates the network performance
when the existing practice was applied. The comparison between Sc-1 and the other
scenarios (Sc-2 to Sc-6) in the 14-bus network suggests that extended ER operating
durations improve the overall reliability performance. In particular, Sc-6 extends the
ER durations, DSTE, DLTE_S, DLTE to 4hrs, 4hrs, 64hrs, resulting in the lowest EENS and
EDLC. Compared to utilities’ Sc-1, Sc-6 at U=1.0 reduces the EENS by 18.3% from 147
MWh to 120 MWh, and declines the EDLC by around 6% from 85hrs/yr to 80 hrs/yr.

However, as shown in Figure 4-7, the reduction in EENS and EDLC becomes less
effective when we continued to increase the ER durations. For example at U=1.0, the
initial one hour increase from DSTE=0 to 1 hr, led to an EENS reduction of 20 MWh,
while the last one hour that increased from DSTE=3 to 4 hr only resulted in an EENS
reduction of 5 MWh. The benefits gained from the increased ER duration becomes
less, especially considering the increased ageing cost.

As shown in Figure 4-7, uprating ERs from 1.0 to 1.2 pu further reduces EENS, EDLC
and generate a better 14-bus network performance. Improving ER by 10% returns a
constant reduction of around 8 MWh in EENS and 2 hrs/yr in EDLC in Sc-3 to Sc-6.
This is because Increasing ER ampacities further enhances the emergency loadability

116
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

of the circuits within allowable durations, thus provides the network more flexibility
to overload the circuits rather than shed the loads.

The increase of ER ampacity has a higher cost of ageing in order to reduce the same
amount of EENS when comparing to the extension of ER duration. Increased ER
ampacities allow the lines to operate at much higher temperatures than those in their
initial designs. Considering the thermal ageing increases exponentially along with the
temperature increases, more ageing occurs on the lines with higher operating
temperatures. Therefore, ER uprating leads to an exponentially increased conductor
ageing, while the improvement of network performance becomes less and less.

1.2 14-bus 1
24-bus
U=1.0 U=1.1 U=1.2
1.16 0.8
EFCCF (Occ./Yr)

1.12 0.6

1.08 0.4

1.04 0.2

1 0
0 1 2 3 4 0 1
DSTE (hrs) DSTE2 (hrs) 3 4

50
14-bus 24-bus
U=1.0 150
U=1.1 U=1.2
40 120
EENA (hrs/Yr)

30 90

20 60

10 30

0 0
0 1 DSTE2 (hrs) 3 4 0 1 DSTE2 (hrs) 3 4
Figure 4-8 EFCCF and EENA of the modelling scenarios in the time domain of
DSTE for the 14-bus (left) and the 24-bus OHL networks (right)

Figure 4-8 shows the frequency of the circuit cascading failure (EFCCF) and the
network OHL ageing (EENA) in the time domain of DSTE for the 14-bus (left) and 24-
bus (right) networks. The results indicates that the overall network performance is
greatly improved but more cascading failures occurs at the same time when the

117
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

allowable ER durations are extended. For example, Sc-6 produces the lowest EENS
and EDLC and the highest EFCCF and EENA. In the scenarios where there was a 20%
increase in the emergency ratings (U=1.2), the EFCCF edged up from 1.06 occ./yr to
1.12 occ./yr indicating a 6% higher probability of cascading failure occurrence, and
the EENA climbed by 62% from 16.5 hrs/yr to 26.7 hrs/yr.

The increased ageing should also, in reality, have an impact on the conductor’s life
and ground clearances, threatening the OHL’s operating reliability and overall
network reliability. The connection between ageing and the increased failure rate has
not been considered in this methodology, however, the high-loading failures were
included in Chapter 3 for the UGCs.

When ERs are implemented with extended durations and increased ampacities, the
network tend to be more and more heavily loaded and thermally constrained.
Therefore the network is operating closer to its loading constraints, thus more easily
reaches the DAL rating and triggers the cascading tripping under ER operations.
These cascading failures seriously threaten the system reliability, and their negative
impact has been included into the network results. In 14-bus OHL network, the
reliability benefits gained from the extended duration totally outweighed the cost of
slightly increased cascading failures, while the 24-bus network showed totally
different characteristics.

The results in the 24-bus OHL network indicated a reduced EDLC, an increased EFCCF
and EENA when extending the ER durations from Sc-1 to Sc-6. But some ups and
downs exist in the EENS result of 24 bus, instead of a gradual decrease of EENS in 14-
bus network. The reason of this EENS fluctuation is that the number of cascading
failures significantly increased in 24-bus network when the ER durations and levels
are increased. For example, the number of cascading failures increased by 670% from
0.126 occ./yr of base case to 0.971 occ./yr of Sc-6 at U=1.2, whereas for the 14-bus
network, it only increased by 6% more. The cascading tripping mainly occurred on
heavily loaded (critical) lines, thus their occurrences may seriously injury the
network reliability. However, the outage duration of the cascading failures was set to
be three hours, which is much shorter than the outage duration of the other normal

118
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

line failures. Therefore, the increased cascading failures only slowed down the
reduction of outage duration index, EDLC. The comparison between the 14-bus and
24-bus network suggests that increasing the ER durations and levels does not always
bring the network reliability benefits due to the increased risk of cascading failures.
It depends on the network load level, network design and structure, which requires
a cost-worth analysis through a simulation.

Figure 4-9 shows the total cost of the OHL conductor ageing and load curtailment
(EAIC) in the time domain of DSTE for the 14-bus (left) and 24-bus (right) networks.
EAIC provides a good cost-worth reference index to weigh the benefits of the load
curtailment reduction against the increased cost of extra ageing when the extensive
ER durations and levels were utilised.

3
14-bus 64U=1.1 24-bus
U=1.0 U=1.2
2.5 63
2
EAIC (106£/yr)

62
1.5
Minimum Cost 61
1
0.5 60
Minimum Cost
0 59
0 1 2 3 4 0 1 2 3 4
DSTE (hrs) DSTE (hrs)
Figure 4-9 EAIC of the modelling scenarios in the time domain of DSTE for the
14-bus (left) and 24-bus OHL networks (right)

As shown in Figure 4-9, the minimum cost of ageing plus the load curtailment was
achieved at Sc-6 with 20% increased ERs in the 14-bus network. This is mainly
because the decreased EENS leads to a lower cost, in spite of the slightly increased
OHL ageing and cascading failures. Compared to the 24-bus network, the 14-bus
network was relatively more meshed and had a more balanced loading, thus the
extended ER duration & ampacity did not result in more cascading failures.

In the 24-bus network, the minimum cost was reached at Sc-4 with 20% increased
ERs. This indicates that unlike the 14-bus network, further extending the ER duration
does not bring about any positive benefits due to the increased ageing and cascading

119
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

failures. There were 2-3 heavily loaded lines in the 24-bus network, which occurred
due to the cascading failures, blocking the power flow from the top generation centre
to the bottom demand centre. In conclusion, the optimal ER duration and magnitude
with the minimum cost depends on the network structure, load demand distribution,
and power flow distribution, which requires further simulations to determine the
details that apply for each individual network.

Figure 4-10 shows the EENS and EDLC outcome of the base case, as well as Sc-
1(black), Sc-3(blue) and Sc-5(red) with an increased ER uprating factor from
U=1.0pu up to 1.5pu in the 14-bus(left) and 24-bus(right) networks.

160 14-bus Sc-1 1800


Sc-3 Sc-5
24-bus
base case
base case
140
1700
EENS (MWh)

120
1600
100
1500
80

60 1400
1.0 1.1 1.2 1.3 1.4 1.5 1 1.1 1.2 1.3 1.4 1.5
ER uprating factor U (pu) ER uprating factor U (pu)
80 14-bus 25 24-bus
base case Sc-1 Sc-3 Sc-5 base case
75 20
EDLC (hrs/Yr)

70
15
65
10
60
55 5

50 0
1.1 1.2 1.01.3 1.4 1.5 1 1.1 1.2 1.3 1.4 1.5
ER uprating factor U (pu) ER uprating factor U (pu)
Figure 4-10 EENS/EDLC of Sc-1, Sc-3 and Sc-5 across the different ER uprating
factors U in the 14-bus (left) and 24-bus UGC networks (right).

120
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

The results comparison between the OHL network (in Figure 4-7) and the UGC
network (in Figure 4-10) indicates that the UGC network has a better network
performance than the OHL network, although the UGCs have relatively higher outage
rate and longer duration. The reason for this is because the cables have a smaller
electrical resistance, leading to less EENS and load interruption due to the voltage
constraints. The UGC network also had a smaller number of cascading failures
because of their higher DAL rating. More cascading failures were avoided in the UGC
network, leading to better network performance.

The results of the 14-bus network in Figure 4-10 shows an improved network
performance when the ER durations and levels were increased at the cost of
increased cable ageing. The increase of the LTE and STE ratings up to 1.5pu returned
a constant and almost linear drop in the EENS and EDLC, regardless of whether a
short or long ER duration was utilised. EENS in the 14-bus network decreased from
142MWh in the base case and bottomed out at 68.9MWh of Sc-5 when at 50%
increased ERs. The 24-bus results showed a ‘U’ curve related to the increased ER
uprating factor. EENS and EDLC in the 24-bus network were reduced with an
increased ERs, which bottomed at around U=1.2-1.3 and gradually edged up
afterwards. This is due to the negative impact of the increased cascading failures
related to the network reliability performance when the ERs were uprated.

Figure 4-11 shows the EFCF and EENA outcome of the base case, as well as Sc-
1(black), Sc-3(blue) and Sc-5(red) at an increased ER uprating factor from U=1.0pu
up to 1.5pu in the 14-bus(left) and 24-bus(right) networks. Only a slightly increased
number of cascading failures occurred in the UGCs in the 14-bus network, thus
leading to no significant negative impact on the network performances. However, in
the 24-bus network, it illustrated an exponential growth of cascading failures when
the increased ERs were utilised. In particular, Sc-5 reached a maximum frequency of
cascading failures, climbing from 0.082occ./yr to 0.721occ./yr when 50% increased
ERs were applied. Comparing Sc-1 to Sc-3 and Sc-5, EFCF also suggests an exponential
growth when the ER durations were extended. The boosted number of cascading
failures surpassed the benefits gained from the extra ER flexibility, leading to a
degraded network reliability.

121
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

0.194 14-bus Sc-1 0.8


Sc-3 Sc-5
24-bus
0.192
EFCCF (Occ./Yr)

0.6
0.19
0.4
0.188

0.186 0.2

base case base case


0.184 0
1.0 1.1 1.2 1.3 1.4 1.5 1 1.1 1.2 1.3 1.4 1.5
ER uprating factor U (pu) ER uprating factor U (pu)
14 14-bus Sc-1 5
Sc-3 Sc-5
24-bus
12
EENA (103hrs/Yr)

4
10
8 3
6 2
4
1
2
base case base case
0 0
1.1 1.0
1.2 1.3 1.4 1.5 1 1.1 1.2 1.3 1.4 1.5
ER uprating factor U (pu) ER uprating factor U (pu)
Figure 4-11 EFCCF, EENA of Sc-1, Sc-3 and Sc-5 across different ER uprating
factor U in 14-bus (left) and 24-bus UGC network (right).

Figure 4-11 shows the EFCCF/EENA outcome of the base case, as well as Sc-1(black),
Sc-3(blue) and Sc-5(red) at an increased ER uprating factor from U=1.0pu up to 1.5pu
in the 14-bus(left) and 24-bus(right) networks. Only a slightly increased number of
cascading failure occurred in the UGCs in the 14-bus networks, thus leading to no
significant negative impact on the network performance. However, in the 24-bus
network, it illustrated an exponential growth of cascading failures when the
increased ERs were utilised. In particular, Sc-5 reached a maximum frequency of
cascading failures, climbing from 0.082occ./yr to 0.721occ./yr when 50% increased
ERs were applied. Comparing Sc-1 to Sc-3 and Sc-5, EFCCF also suggests an
exponential growth when the ER durations were extended. The boosted number of
cascading failures surpasses the benefits gained from the extra ER flexibility, leading
to a degraded network reliability.

At a low ER level from U=1.0 to 1.3, only a slightly increased thermal ageing of UGCs
occurred compared to the base case. This is because UGCs are more tolerable of the

122
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

short-term overload current compared to the OHLs due to the UGCs’ higher DAL
rating and larger thermal inertia. The cable thermal inertia provided the network
with greater short-term and long-term emergency resilience and flexibility at a
reduced cost of thermal damage and ageing on the cables. However, as the ERs
continued to increase, the EENA of the different scenarios quickly diverged and
increased exponentially. In particular, Sc-5 had a substantially higher EENA than the
Sc-3 and Sc-1 at the high ER factor of U=1.5. High emergency loadings at a long
duration will heat up the cable and its surroundings, and thus produce high
conductor temperatures. Because the thermal ageing of UGCs increase exponentially
at a higher temperature, the EENA of the UGCs will increase exponentially along with
the increase of the ER duration and ampacity.

3
14-bus 35Sc-3
24-bus
Sc-1
Sc-5 base case
base case 34
2.5
EAIC (106£/yr)

2 33
32
1.5
31
1 30
Minimum Cost
0.5 29
0 28 Minimum Cost
1.0 1.1 1.2 1.3 1.4 1.5 1 1.1 1.2 1.3 1.4 1.5
ER uprating factor U (pu) ER uprating factor U (pu)
Figure 4-12 EAIC of Sc-1, Sc-3 and Sc-5 across the different ER uprating factors
U in the 14-bus (left) and 24-bus UGC networks (right).

Considering the exponentially increased cascading failure and thermal ageing of the
UGCs, the benefits gained from the extensive ERs will eventually outweighed by the
extra costs. Figure 4-12 illustrates the total cost of the UGC thermal ageing and load
curtailment (EAIC) in the base case, as well as Sc-1(black), Sc-3(blue) and Sc-5(red)
across U=1.0pu to 1.5pu in the 14-bus(left) and 24-bus(right) networks.

As indicated in Figure 4-12, the minimum cost of ageing and load curtailments was
achieved at Sc-5 between 30%-40% increased ERs in the 14-bus network, and
obtained at Sc-3 or Sc-5 with U=1.23. Compared to the OHL network, the UGC
network saved 0.4∙106£/yr in the 14-bus network and almost 30∙106£/yr in the 24-
bus network, by allowing for a much higher ERs and more extended ER duration. Less

123
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

of an ageing cost in the UGC network were due to the fact that the UGCs have a greater
potential to tolerate emergency loading for long durations due to their large thermal
mass. Less cascading failures in the UGC network were because the high DAL rating
of the UGCs (2×Inormal) provides more flexible overloading room for the UGCs, and a
higher triggering condition for cascading failures.

The traditional practice defining the current ampacity and duration criteria of
emergency ratings was conservatively established based on the standards in 1970s
and the utilities’ experience. However, there has been a recent increasing need for an
efficient use of infrastructure and improved network flexibility, particularly when an
increasing penetration of renewable energy is integrated. This leads many utilities to
consider alternative emergency rating practices for both cables and lines. However,
the potential values and quantified risks of such a deployment have not yet been
quantified, disabling utilities from making decisions about more aggressive rating
policies.

This chapter introduced a methodological improvement based on the SMC-based


reliability evaluation for transmission networks, in order to assess the impact of
emergency rating utilisation on network reliability and plant ageing. This study
proposes an extensive emergency rating practice that considers increased ampacities
and extended durations of emergency ratings, compared to the traditional practises.
This methodology integrated a cascading failure model and thermal-ageing model of
both UGCs and OHLs into a SMC-based network evaluation. This integration allows
utilities to quantify the potential benefits on the network reliability performance
against the increased risks of cascading failure, as well as the thermal ageing of
circuits when extensive emergency ratings are implemented.

Results from the case studies regarding the14-bus and 24-bus networks have shown
that increased ERs ampacity and extended durations enhance network resilience and
flexibility during post-contingency operations. In the 14-bus network, Sc-5 with 20%
increased ERs reduced the load curtailment by over 25%, compared to the traditional
practice. However, extensive ERs utilisation leads to a higher risk of cascading

124
Chapter 4: Impact of Extensive Emergency Ratings on Network Reliability and Plant Ageing

failures and increased plant ageing, particularly in some unmeshed networks (i.e. 24-
bus). The negative impact of the additional cascading failures totally overweighed the
benefits brought in by the ERs when the ERs durations and ampacities were
increased to a high level. For example, the EENS bottomed at Sc-3 with 20% increased
ERs in either the UGC or OHL 24-bus network. The cost analysis also indicated that a
minimum cost was achieved with a specific ERs operating scheme, considering of the
cost of the load curtailment and plant ageing.

By comparing the results in the 14-bus and 24-bus network, it is indicated that the
impact of ERs highly depends on the network load level, network design and
structure, which requires a cost-worth analysis for each specific network through
simulation. Therefore, the proposed methodology can inform the utilities of the
quantified results, and help to determine the optimum time duration criteria and
rating ampacities for OHLs and UGCs in transmission networks. The main aim of our
study was to provide options and recommendations to power utilities on changing
their existing emergency rating policy to an extensive ERs operating strategy for their
transmission circuits.

Equation Chapter (Next) Section 1

125
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

5 Increasing Wind Power Integration through


Flexible Cable Current Ratings

Utilities are challenged to develop economic solutions to enhance the network’s


capacity to prompt an increased penetration of wind energy. Traditional rating
methods for cable circuits connecting wind farms to their host utilities neglect the
stochastic and intermittent nature of wind farm output. The idea of this paper was
driven by the request from wind farm developers that ‘increasing their generation
capacity in the future and noting any spare headroom capacity that might be
available’. Advancing the existing rating practices of wind farm cable circuits
toward a flexible rating approach could provide a competitive solution to this
challenge.

This chapter proposes new flexible current rating methods for the tie circuits of
wind farms to efficiently increase the wind integration into existing networks
without the need for reinforcements or additional instrumentation. The proposed
method advances the current practices by utilising the system’s thermal inertia,
forecasted wind farm output and cable aging risk in order to determine the time-
dependent flexible current ratings. The proposed flexible current rating has been
compared against the existing seasonal, probabilistic, and dynamic ratings on a
33kV wind-farm distribution circuit in Lanarkshire, Scotland. The study considered
a multi-year cost-worth analysis of the wind farm operation with the aim of
identifying an optimum cable operating strategy that minimises wind curtailment
as well as cable maintenance and interruption costs due to the cable aging risks.

This industrial case study indicated an effective 21% increase of wind integration
on the existing cables with the implementation of the proposed flexible cable
current rating method compared to the static thermal rating (SCR). This approach
provides a simple and direct-to-implement uprating solution to increase wind
integration without the instrumentation complexity required by dynamic cable
rating (DCR) practices and with a similar cable aging risk. The proposed methods
thus advance practices by providing flexibility at different levels of operational risk.

126
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

Power utilities pursue ways to enable the integration of increasing renewable energy
sources (RES) in their networks. The number of wind generator connection requests
in existing wind farms is also expected to increase [92]. However, there is little or no
spare network capacity available for additional wind power connections based on
traditional power rating methods [25]. These methods are calculated by using
established technical considerations and operational scenarios, but neglecting the
fact that wind farm outputs are inherently variable and that there will be frequent
occasions when the available assets are not fully utilised [93, 94]. To realise the
network capacity for additional wind power integration, flexible rating approaches
and new control strategies have been implemented in the tie cable circuits between
wind farms and their host utilities by distribution network operators in order to
improve network flexibility and loadability, and to avoid net-work traditional
expensive reinforcements [73, 92].

Conventional practices determine the cable size for wind farm connections based on
the sustained maximum output of the wind farms [95]. This static cable rating (SCR)
assumes a constant loading regarding the rated wind farm output and a conservative
environmental condition at all times. A less conservative probabilistic cable rating
(PCR) method considers the historical weather parameters and a continuous loading
at designed rating to establish the probability of exceeding the design temperature
(excursion time) [58, 62]. Nonetheless, the wind farm output follows a stochastic and
intermittent variation, and maintains its maximum capacity only temporarily. It is
suggested that a wind farm exports above 90% of its full capacity only 21% of the
time and that this usually lasts for 20 consecutive hours or less [96]. The temperature
variations of a cable system are not instantaneous like the changes in a power flow
due to the cable’s large thermal inertia. This thermal inertia could allow the cable
circuits to dissipate the heat gained during high wind loading periods at low wind
periods [93, 94]. Therefore, due to the intermittency of wind farm output and the
cables’ thermal inertia, high cable operating temperatures may not occur as
frequently as with other network cables. The potential loading capability of the cable
circuits is thus substantially higher than that recommended by traditional static

127
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

rating practices. Consequently, dynamic cable rating (DCR) approaches are more
suitable for wind farm circuits.

DCR uses the measured temperatures via optical fibre sensors together with the
cable’s thermal model to calculate the conductor temperature and time-varying
current rating, considering its transient thermal behaviour [30, 44]. Neglecting the
complexity involved in implementing and operating DCR on transmission systems,
this study focused on the possibility of applying such concepts only to the tie line
between a wind farm and its host utility. DCR is an efficient solution when deployed
in a closed-loop controlled system where the wind farm output is curtailed by an
active network management (ANM) system to ensure that the circuit always operates
within permissible temperatures [25, 97]. Scottish Power Energy Networks trialled a
distributed temperature sensing (DTS) technology to monitor real-time
temperatures and to calculate the DCR of three 33kV wind farm circuits within a
network innovation project [98]. However, DCR requires complex equipment and is
time-consuming to construct, which prohibits its implementation by existing wind
farms.

This study has proposed alternative flexible current rating (FCR) methods that can
be implemented without the required instrumentation and complexity of DCR in
order to optimise the loading capability of the existing cable tie lines of wind farms.
It calculated a continuously updated transient rating considering the loading history
and cable thermal inertia using estimates rather than actual measurements regarding
the soil thermal properties. FCR can be advanced by the forecasting of wind farm
output in order to identify a forecasted flexible current rating (FFCR) that allows for
the cable to operate above its permissible temperature limit temporarily. The
benefits of these methods have been demonstrated with a multi-year techno-
economic analysis that considers the wind curtailment as well as the cost of
estimation errors and forecasting risk on cable aging against other traditional rating
methods.

To achieve this, the study aims to contribute with the following methodological and
implementation advancements:

128
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

 Integration of the cable’s transient thermal model and thermal aging model in
circuit loading in order to estimate the cable’s aging under time-varying loading.

 Novel flexible rating approaches that use the cable thermal inertia to unlock
additional capacity available in cable circuits for extra wind integration.

 Evaluation of the benefits and aging risks associated with the different flexible
rating approaches.

 Long-term economic evaluation of the methods considering the wind


intermittency and uncertainty effects on cable loading and aging risk.

5.2.1 Outline of the Proposed Methodology

To quantify the economic benefits and aging risks of the different rating methods in
the wind farm underground cable (UGC) circuits, the proposed methodology
integrated a rating calculation model and a thermal-aging model of UGCs within a
time-sequential analysis. The analyses were performed through iterations in a
sequential time domain tΔt, segmented in steps Δt, for each simulation year Y,
∀Y∈{1,⋯, NY}. The time-step analysis calculated the changes in the UGC rating
Irated(tΔt), loading Iflow(tΔt), conductor temperature θC(tΔt) and potential aging at every
Δt, while the annual analysis calculated the annual wind farm and cable performance.
Figure 5-1 has outlined the main steps.

The wind farm data is comprised of the wind speed Vwind(tΔt), and available wind
power Pwind(tΔt), for every time step Δt within each simulation year Y. The cable system
data defined the design and surrounding environment properties of the wind farm
cable circuit. This included the UGC design properties, ducting and laying
configuration, time-varying soil temperature θSoil(tΔt), soil thermal resistivity ρSoil(tΔt),
and volume specific heat CSoil(tΔt). The rating method for the UGC circuits can be
selected by the network operators, considering the appropriate criteria based on
wind farm properties, cable system design and their aging risk acceptance.

129
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

CABLE SYSTEM DATA


WINDFARM Ducting & Laying Soil Design
DATA
Selection of cable rating
Wind Speed
Flexible Static, Seasonal, Dynamic
Wind Power Rating Probabilistic Rating Rating

Cable rating calculations & AC PF


Irated(tΔt)=FFCR(tΔt), FCR(tΔt), DCR(tΔt)
Actual cable loading - Iflow(tΔt)
Annual Windfarm output Time-step
analysis analysis
(Y=Y+1) Temperature calculation - θC(tΔt) (tΔt=tΔt+Δt)
Cable ageing risks
Annual computation is completed
YES NO
Analysis is completed at simulation year N
NO YES
Annual wind farm & cable performance indices calculation
Long-term cost-worth analysis
Figure 5-1 Method to evaluate different cable utilisation on wind farm circuits

After the input data was set, the computations were initialised so as to determine the
calculated current rating Irated(tΔt), of the underground cable circuit at any specific Δt
of time domain tΔt under the selected method (i.e. DCR, FCR, FFCR; the complete
modelling has been detailed in Section 5.2.2). The actual cable loading Iflow(tΔt) was
obtained through an AC power flow considering the cable rating constraints Irated(tΔt),
and the available wind power output Pwind(tΔt), at tΔt. To capture the risk of thermal
overloading, a combined thermal-aging model of UGCs was implemented to calculate
the operating UGC temperature θC(tΔt) and its associated aging, at every Δt. The
cable’s thermal-aging modelling was based on the thermo-electric equivalent (TEE)
and the Arrhenius models, as detailed in Section 2.5.

The time-step computations were performed sequentially until the analysis was
completed within a given simulated year. A new simulation year was then initiated
with the new input data of wind speed, wind power and soil properties. The annual
analysis was terminated when a sufficiently large number of years NY was modelled.
Annual wind farm and cable performance indices (detailed in Section 0) were then
calculated: eexpected wind power curtailed (EWPC), expected equivalent cable aging

130
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

(EECA) and expected cable utilisation profit (ECUP), and compared to static cable
rating. These indices were used within a multi-year economic analysis of wind farm
operation to identify an optimum cable operating strategy to minimise wind
curtailment, network maintenance and interruption costs due to cable aging.

5.2.2 Cable Rating Computation

1) DTS Modelling

The measuring optical fibber, used in the DTS system, was usually laid in the gap
formed in the trefoil formation of three touching cable ducts [99]. Since optical fibre
only measures the temperature of DTS itself, a thermal model was required to
calculate the cable core temperature from the measured duct surface temperature.
For cable circuits equipped with DTS, the developed TEE circuit in Figure 2-5 was
simplified by removing all soil layers, apart from the one used for the duct, and setting
the measured temperature θDTS(tΔt) as the input conditions.

28
28
28
28 Measured duct temp Computed duct temp Loading
0.7
0.7
0.7
0.7
Measured
Measured
Measuredduct
duct
ducttemp
temp
temp Computed
Computed
Computedduct
duct
ducttemp
temp
temp Loading
Loading
Loading
26 0.6
(°C)

26
26
26 0.6
0.6
0.6
Temperature(°C)
(°C)
(°C)

24
24
24
24 0.5
0.5
0.5
0.5

(kA)
Load(kA)
(kA)
(kA)
Temperature
Temperature
Temperature

22
22
22
22 0.4
0.4
0.4
0.4

Load
Load
Load
20
20
20
20 0.3
0.3
0.3
0.3
18
18
18
18 0.2
0.2
0.2
0.2
16
16
16
16 0.1
0.1
0.1
0.1
14
14
14
14 0000
0000 30
30
30
30 60
60
60
60120
120
120
120 90
90
90
90150
150
150
150 180
180
180
180 210
210
210
210 240
240
240
240
Time
Time
Time (h)
Time(h)
(h)
(h)
Figure 5-2 Measured and calculated duct temperatures based on cable loading

Due to the lack of long-term DTS measurement data, the model can utilise the
measured soil thermal properties and suggested TEE model in order to calculate the
conductor temperature θC(tΔt) at tΔt. This approach was validated with the DTS data
from [99]. A comparison of the measured and calculated duct surface temperatures
was given in Figure 5-2, along with the calculated conductor temperature and
recorded wind farm outputs. The errors between the measured DTS and calculated
temperatures were negligible, indicating the high accuracy of the DTS system
approach.

131
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

2) Dynamic, Flexible and Forecasting Rating Calculation

The DCR(tΔt) and FCR(tΔt) were calculated for every Δt considering the cable system
transient thermal loading on the step in which the current changes occurred. Thus,
IDCR(Δt) and IFCR(Δt) were the maximum permissible currents that result in a maximum
conductor temperature limit, θC,Max, at the end of the Δt period. The required actions
to control the wind farm output were triggered at the start of Δt to limit the cable
loading when the conductor temperature was calculated to exceed this limit. The
FFCR(tΔt) was calculated in a similar manner. However, it is allowed to temporarily
exceed the θC,Max and operate at a higher temporary temperature limit, θC,Temp, as set
by the operator. This way, the IFFCR(Δt) is an increased risk rating method that provides
a higher flexibility for wind integration in the system. The initial input conditions for
DCR, FCR, and FFCR calculation have been stated in Table 5-1. To prevent cable aging,
the thermal limit θC,Max was set to the manufacturer’s maximum continuous
conductor temperature (e.g., 90°C), while the θC,Temp and duration were determined
by the operator preferences, usually at emergency loading.

Table 5-1 Initial Conditions for DCR, FCR and FFCR Calculations
Initial Conditions DCR FCR FFCR
Soil properties Measured values Pre-defined*
Initial thermal state (°C) θC,In(tΔt) θ'C,In(tΔt)
Final state thermal limit (°C) θC,Max θC,Max θC,Temp
Thermal transient duration, tθ (mins) Δt Δt tH-W
*based on measurements used in SCR and utility practices conservatism

DCR was computed using the cable’s previous Δt thermal state as the initial condition,
θC,In(tΔt). It measured the soil temperature, thermal resistivity, and volume specific
heat at the present Δt with a thermal transient calculation time of tθ=Δt. Thus, the
IDCR(tΔt) calculation provided the current that resulted in the maximum final
temperature at the end of Δt, which is equal to θC,Max.

FCR was calculated based on an estimated initial thermal state θ'C,In(tΔt), as no


measurements were available, together with a set of soil properties used in SCR and
PCR practices. Hence, the FCR’s soil properties can be pre-defined by the wind farm
operator, or utility, based on the cable aging risk acceptance levels. The soil condition

132
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

used in the SCR calculation was regarded as the most conservative, while PCR-Pex%
applied more flexibility and allowed for an exceeding of the cable’s design
temperature by a Pex% probability when assuming maximum continuous loading.

FFCR was calculated considering a temporary operation at θC,Temp, when the


estimated θ'C,In(tΔt) was greater than θC,Max. In this case, the transient calculation step
tθ was determined by the forecasted high-wind load period tH-W, which was expected
to load the cable above its θC,Max. Thus, tH-W is a fraction of the complete forecasted
time window of tforecast hours, which is dependent on the forecasted wind power
output. The tH-W was determined by an auto-regressive moving average (ARMA)
model, detailed in Section 5.2.3. The FFCR was applied only when the θ'C,In(tΔt) was
lower than the elevated limit, θC,Temp. When θ'C,In((tΔt) has reached the θC,Temp then a
reduced continuous ‘cooling’ cable rating, equal to SCR or PCR-Pex%, was applied
until the cable temperature was reduced to levels below θC,Max, in order to avoid
excessive thermal loading and cable aging.

NO DTS-ANM system Installed YES


YES NO
FFCR Wind forecasting FCR DCR
Limit=θC,Temp, tθ=tH-W Limit=θC,Max, tθ=Δt Limit=θC,Max, tθ=Δt
NO
θC,Max≤θ'C,In(tΔt)<θC,Temp TEE model
Iterative loop
YES Wind forecasting to Thermal rating calculations
identify tH-W duration IDCR(tΔt)±1, θC,Fi(tΔt)→θC,Max
θ'C,In(tΔt)<θC,Max: no need for FFCR IFCR(tΔt)±1, θ'C,Fi(tΔt)→θC,Max
θ'C,In(tΔt)=θC,Temp: Cooling Cable Rating IFFCR(tΔt)±1, θ'C,Fi(tΔt)→θC,Temp

ISCR(tΔt) or IPCR(tΔt) until θ'C≤θC,Max IFFCR(tΔt) IFCR(tΔt) IDCR(tΔt)


Figure 5-3 Computational steps for DCR, FCR and FFCR calculations at Δt

The computational steps for DCR, FCR and FFCR have been shown in Figure 5-3. After
the initial input conditions are determined for each rating, the TEE model calculates
the final conductor temperature, considering its thermal transient behaviour. DCR
calculates the actual θC(tΔt) using measured soil properties, when he DTS-ANM
system is in place, while FCR and FFCR use an estimated θ'C(tΔt) on the pre-defined
soil properties. A computational loop that increases or reduces the calculated rating

133
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

by 1 Amp was applied to calculate the maximum permissible current in each rating
approach. The IDCR(Δt) and IFCR(Δt) follow similar computational paths (in Figure 5-3)
with the second method providing flexibility to the operator via the estimated
transient thermal loading of the cable. The IFFCR(Δt) includes an additional flexibility
element at times when the wind forecasting indicates high wind power output as
shown in Figure 5-3. However, this is only temporarily implemented with a cooling
period following the overloading. The output of the calculations was Irated(tΔt).
150
Load Current Condcutor temp

120
FCR FFCR
C(tΔt)

150
90 high-wind tH-W low-wind
temp

120 θC,Temp
60
θ'

θC,Max
θ'C(tΔt)

3090
Iflow(tΔtCondcutor

1800060
150030 FCR
1200
)

18000 Case 1
900
1500
Load Current

Case 2
600
1200
Iflow(tΔt)

300
900
0
600
300 Cooling periods Irated=SCR
0
t t+5Δt t+10Δt t+15Δt t+20Δt t+25Δt
Time Axis tΔt
Figure 5-4 Example of FCR and FFCR application under a short-term high wind
power output

Figure 5-4 illustrates an example of a wind farm circuit under FCR and FFCR
implementations, which displays the estimated conductor temperature θ'C(tΔt) and
the current loading Iflow(tΔt) changes in the time domain. Case 1 and 2 represent
FFCRs with same θC,Temp limit with tforecast,1 being smaller than tforecast,2. It can be
observed that the FCR is utilised in low-temperature operations, i.e., when initially
the θ'C(tΔt) is lower than the cable θC,Max, while the FFCR is activated. FCR constraints
the θ'C(tΔt) to θC,Max all of the time, and therefore allows for a lower output from the
wind farm. FFCR allows for increased flexibility during the forecasted high-wind
periods. The longer tforecast,2 case produces smoother gradual changes for θ'C(tΔt) and
Iflow(tΔt) within the FFCR period, while a short tforecast,1 in Case 1 calculates a higher
rating Irated(tΔt) with a rapid increase of θ'C(tΔt). Thus, in Case 1, the cooling cable
rating of SCR was applied to cool down the cable from θC,Temp back to θC,Max. Then, a
new FFCR was computed to elevate it to θC,Temp again within the high-wind period.

134
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

5.2.3 Wind Power Forecasting using ARMA

The ARMA time series model was implemented to forecast the future values of wind
speeds within the short-term time horizon of tforecast hours ahead. The model was
described by (5.1), where the realisation of the time-series wind values V(tΔt+k) at
time tΔt+k depending on a linear combination of past wind values was observed
V(tΔt+k-n) with the addition of a moving average series e(tΔt+k-m). The e(tΔt) series is
the normal innovations distribution model (NID) for the white noise vector with zero
mean and sigma variance, i.e., e(tΔt)∈ NID(0,σa2). The ai(i=1,…, n) and bj(j=1,…, m) are
autoregressive (AR) and moving average (MA) coefficients with n and m defining
their order, while k ∈ (0, tforecast) is the number of advanced forecasting hourly
intervals from time tΔt.

Vtt k =a1×V (tt +k -1)++an ×V (tt +k -n)


(5.1)
+e(tt )-b1×e(tt +k -1)--bm×e(tt +k -m)

Monthly ARMA(n,m) models were established to demonstrate different wind speed


characteristics for each month. Different AR and MA orders and coefficients were
determined for each monthly model through the identification process using the
Matlab toolbox ARMASA [100]. The historical hourly wind speed data was utilised as
the primary data source to train the model. Once the model was built for each month,
the future forecasted values V(tΔt+k) were obtained by replacing the V(tΔt+k-n) values
on the right of the (5.1) by their respective forecasts. Their residuals were
substituted by the expected value of zero [101]. Forecasting was executed at every Δt
to predict the future wind speed in the next tforecast hours, with the hourly wind speed
over the previous 12 hours utilised as the input signal.

The forecasted wind speed V(tΔt+k) was converted to the forecasted wind power
output Pwind(tΔt+k) through the empirical wind turbine’s power curve (5.2).

0 0  V (tt )<Vci

(A+B  V (tt )+C  V (tt ) )  Pr Vci  V (tt )<Vr
2
Pwind (tt )=  (5.2)
 Pr Vr  V (tt )  Vco
0
 Vco < V (tt )

135
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

The power output was defined by turbine cut-in Vci, rated Vr, and cut-out Vco, wind
speeds, while Pr was the rated power of a generator unit. The constants, A, B, C can be
calculated as functions of Vci, Vr and Vco using (5.3) [102].

1   Vci  Vr  
3

A Vci Vci  Vr  - 4VciVr   


(Vci - Vr ) 2   2 Vr  
1   Vci  Vr 
3

B  4 Vci  V 
r   - (3Vci  Vr 
) (5.3)
(Vci - Vr ) 2   2Vr  
1   Vci  Vr  
3

C 2 
2 - 4 
Vci - Vr    2Vr  
5.2.4 Metrics of Long-term Wind Integration Risk-Worth Analysis

To capture the increased wind power accommodated in the network due to different
cable rating approaches, a new index was used to record the expected wind power
curtailed EWPC per annum, which was calculated by (5.4) where WPCΔt was the wind
power curtailment within time step Δt in year Y. However, the risk of increased cable
loading under FCR and FFCR was quantified using the expected equivalent cable
aging index EECA as calculated by (5.5). Thus, the average value of the annual
equivalent cable aging of every elevated temperature event was calculated by
converting it to the equivalent aging at 90°C using (5.5). This temperature was
selected as it is considered as a normal cable operating temperature and allows for
quantifying the aging risk. The annual thermal life loss was obtained by the EECA,
divided by 8760 hours. Consequently, the expected cable aging cost ECAC was
calculated using (5.6) by indicating the annual thermal life loss in hours as a fraction
of the cable’s design Life, in years, and the circuit replacement costs, CRep.

 Y
N

EWPC     WPCt  N (5.4)
Y 1  t 1 
 Y
N

EECA     ECAt  N (5.5)
Y 1  t 1 
ECAC   EECA 8760 Life   CRep (5.6)

136
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

The annual expected cable utilisation profit (ECUP) of any implemented rating
method (RM) was computed using (5.7) with any profit compared to the base SCR
case. The ECUP considered the cable aging costs and generation profits from the
reduced wind curtailment, with the latter accounting for the feed-in-tariff FIT rate. A
rate of 0.83 p/kWh for wind generation was used in the UK [103], which was also
implemented in this study.

ECUP=ECACRM -ECACSCR +  EWPCSCR -EWPCRM ×FIT (5.7)

In the long-term economic analysis, this methodology utilised a probabilistic failure


model to predict the cable failure rate for old worn-out cables. Thus, the failure
probability function f(LT) of the cable at its thermal age LT, was expressed in (5.8)
where α is the scale parameter representing the dispersion in time-to-failure. The
parameter α of 3.8 was obtained for the XLPE cables via the breakdown tests [55].
Based on the statistical theory of component failure, the cable failure rate h(LT) at
thermal age LT, was determined by (5.9). The thermal age LT, of the cable at
operational year Yo, was calculated by (5.10) and this indicates that for every year of
operationm the risk of failure increases when additional thermal aging is formed due
to FCR or FFCR.

f ( LT )  1  exp    LT L0   e BcT 

(5.8)

1 d f ( LT )
h( LT )   (5.9)
1  f ( LT ) d LT
Yo
LT   EECAy / 8760 (5.10)
y 0

Consequently, the economic benefits from the flexible rating methods were evaluated
against the other rating methods as shown in Figure 5-5. The total profit function P(Yo)
compared to SCR at operating year Yo was expressed in (5.11) considering the initial
equipment cost CEqu, cable aging and wind curtailment costs. This was calculated in
ECUP, and was shown to exponentially increase the maintenance costs CMai, as well as
the interruption costs CInt, due to cable failures. Consequently, the profits of the

137
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

different rating methods were quantified in a comparative analysis to identify the


optimal uprating solution in order to increase wind integration.

Yo
P(Yo )    ECUPy +CMai × e LT -1 + CInt  h( LT )   CEqu (5.11)
y 0

Rating Flexible Static, Seasonal, Dynamic


methods Rating Probabilistic Rating Rating
Forecast & estimation error Exceedance risk ANM-DTS system
Actual windfarm output
Equipment cost
Cable Arrhenius model for thermal Wind curtailment cost
ageing and lifetime estimation Cable ageing &
Maintenance costs
Cable probabilistic failure model Interruption cost
Cost-worth Comparison
Figure 5-5 Long-term cost-worth analysis of the different rating methods

5.3.1 Design of Test Network – wind farm cable circuit

The proposed methodology was applied to a case study of an actual wind farm in
Lanarkshire, Scotland with a total capacity of 40 MW. Wind turbines with a rated
capacity of 2 MW, and cut-in, rated, and cut-out speeds of 3, 12, and 25 m/s,
respectively, were used. The historical hourly wind speed data [104] collected at the
Ringway weather station in the UK from 1949 to 2008 was used as the input. This
data allowed for us to calculate the hourly wind farm output from 1949 to 2008. The
different levels of wind farm output as a percentage of the installed capacity were
found shown as being very similar to the values recorded in Lanarkshire [96],
indicating a realistic modelling of wind power output.

The wind farm was connected with a 21.4 km long 33 kV cable circuit to the East
Kilbride South 275/33 kV substation. Single-core un-armoured XLPE cables with 630
mm2 copper conductors were used in the network. The three phase circuit cables
were installed in ducts in a touching trefoil formation, buried at a depth of 0.8 m. The

138
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

33 kV cable circuit was designed with a nominal life of 30 years, and its replacement
(excavate & cable purchase) cost was set to be £191,300/km. The annual
maintenance cost was £1,000/year∙km and the interruption cost (cable repair and
wind generation interruption) was £58,000/occ. [105, 106].

20.6%
20.6% 46.3%
46.3% 0% - 20% Windfarm output 21.0% 46.0%
11.2%
11.2% 20% - 40% Windfarm output 11.0%
7.4%
7.4% 40% - 60% Windfarm output 9.0%
14.4% 60% - 90% Windfarm output 14.0%
14.4%
90% - 100% Windfarm output

Figure0%5-6 Probabilities
- 20% Windfarm output
of the modelled
20% - 40% Windfarm output
(left) and recorded 0%(right)
40% - 60% Windfarm output
wind farm
- 20% Windfarm output 20% - 40% Windfarm output
output being
60% - 90% Windfarm output at 90%
different levels of the maximum wind farm
- 100% Windfarm output 60% - 90%capacity.
Windfarm output 90% - 100% Windfarm output

5.3.2 Rating Scenarios Considered for the Test Network

The hourly time-varying soil temperature, soil thermal resistivity and soil volume
specific heat data, i.e., soil property values, at the Ringway weather station from 1949
and 2008 from the Met Office [104, 107] were used to calculate the cable conductor
temperature. Figure 5-7 shows the soil’s annual winter (Dec-Feb), spring (Mar-May),
summer (Jun-Aug) and autumn (Sept-Nov) temperature frequencies.

1000
1000
1000 Annual
Annual
Annual
Winter
(hrs/yr)

800 Winter
Winter
Frequency(hrs/yr)
(hrs/yr)

800
800 Spring
Spring
Spring
600
600
600 Summer
Summer
Summer
Autumn
Frequency
Frequency

400 Autumn
Autumn
400
400
200
200
200
0
00
0
00 5
55 10
10
10 15
15
15 20
20
20 25
25
25
Soil
Soil Temperature
Temperature(°C)
SoilTemperature (°C)
(°C)
Figure 5-7 Frequency distributions of the soil’s annual, winter, spring,
summer and autumn temperatures.

The cable’s rating, including SCR, PCR and DCR, was calculated based on the
manufacturer’s maximum permissible conductor temperature of 90°C. The SCR was
calculated based on the values in ENA P17 [17]. The use of typical static and seasonal
soil values was shown in Table 5-2 with the latter used to calculate the seasonal cable

139
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

ratings (SeCR). The typical seasonal values of soil temperature, thermal resistivity,
volume specific heat and their corresponding seasonal ratings have been presented
in Table 5-2.

Table 5-2 Soil Properties Used for Static and Seasonal cable rating
Ratings Istatic Iwinter Ispring Isummer Iautumn
Current rating (A) 745 829 775 724 775
Soil temperature (°C) 15 10 15 20 15
Thermal resistivity (Km/W) 1.2 0.9 1.05 1.2 1.05
Volume specific heat (J/m3K) 1.9∙106 2.2∙106 1.9∙106 1.6∙106 1.9∙106

Probabilistic cable rating is less conservative than static rating by including a


probability of exceeding the design temperature (e.g., 90°C). Increasing this
probability (excursion time) increases the cable’s current rating, but also the risk of
accelerated thermal aging. These excursion probabilities Pex, were produced using
the soil property values and the steady state cable thermal model, under the
assumption of there being continuous cable loading at its current rating throughout
the year. Therefore, these excursions neglect the transient temperature changes of
cable conductor temperature and the varying wind farm outputs. From the historical
soil temperature and thermal resistivity data 1949-2008, the excursion probability
curve for the cable circuit was calculated as shown in Figure 5-8 (Left). The SCR, as
well as the PCRs with 20%, 60% and 100% excursions, were highlighted in the
corresponding figure. Figure 5-8 (Right) indicates the thermal overloading risk under
PCR conditions in terms of the annual operating hours at each temperature with the
shaded areas highlighted the excursions, and as expected, this was zero for SCR.

100
100
100 999 SCR PCR-20%
SCR
SCR PCR-20%
PCR-20%
2hrs/yr)

8 PCR-60% PCR-100%
2hrs/yr)
Frequency (102hrs/yr)
(%)

PCR-100% 8 8 PCR-60%
PCR-60% PCR-100%
PCR-100%
(%)
time(%)

PCR-20% PCR-60% PCR-100%


PCR-100% 777
859A
PCR-60% 859A
PCR-20% PCR-60%
10 PCR-20% 859A 666
time

825A
Excursiontime

1010 791A 825A


825A
791A
(10

791A
555
Frequency(10

444
Excursion
Excursion

100%
Frequency

1 3 100%
100%
11 33 60%
SCR 222 60%
60%
SCR
SCR
745A 111
745A
745A 20%
0.1
0.1 000 20%
20%
0.1
740 760 780 800 65
6565 70 75
7575 80
8080 85 90
9090 95
9595100 105
105110
740
740 760 780
760Design 800 820
800
780 Ampacity820(A)840
820 840 860
840 860 7070Conductor 8585
Temperature 100
100 105
(°C) 110
110
Design
DesignAmpacity
Ampacity (A)
(A) Conductor
ConductorTemperature
Temperature(°C) (°C)
Figure 5-8 Excursion times and temperature distribution under SCR and PCR.

140
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

5.3.3 Modelled Rating Scenarios

The proposed methodology was simulated on the wind farm cable circuit to evaluate
the impact of wind spillage reduction and cable aging risks under different rating
scenarios as described in Table 5-3. The scenarios were compared against the SCR
method to quantify any profits from increased cable utilisation. SeCR implemented a
rating on the seasonal basis of the soil’s properties, as stated in Table 5-2. Less
conservative PCR utilises a higher continuous current rating by including the risk of
thermal overloading (excursion time). DCR utilises a time-varying current rating
based on the real-time measurements from DTS, which were calculated considering
the cable’s transient thermal behaviour. FCR is a hybrid solution based on SeCR and
DCR methods. It utilises the pre-determined soil properties from SeCR and the
continuously updated transient rating, due to cable thermal inertia, as implemented
in DCR considering the previous cable loading state. FFCR further advances the
flexibility of the FCR method to accommodate more wind by including the forecasting
of wind farm outputs and allowing the cable to exceed its continuous design
temperature limit of 90°C at θC,Temp.

Table 5-3 Modelling Scenarios of Wind farm Cable Ratings


Rating Methods Forecasting Rating (A) Comments
SCR No 745 Static throughout the year
SeCR No From Table III ISeCR depends on the season
PCR-20% No 791 20% excursion risk
PCR-60% No 825 60% excursion risk
PCR-100% No 859 100% excursion risk
DCR No Varying Transient rating
FCR No Varying Transient rating
FFCR Yes Varying Transient rating

The SeCR and PCR were compared against the existing SCR standard practice due to
the cable thermal modelling similarity. All of these methods do not consider the
thermal inertia of the cable system, since they were calculated based primarily in
steady-state loading. The increase of wind power integration with SeCR and PCR is

141
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

evident in Figure 5-9, which illustrates the amount of expected wind power curtailed
(EWPC), expected cable aging (EECA) and the increase in profits (ECUP) when the
installed wind farm capacity was doubled (i.e., 2pu). By increasing the excursion time
(Pex) of PCR, a significant reduction in wind curtailment was observed, even at 2 pu
wind integration levels. In particular, the SeCR increased wind integration in the
network by about 5GWh and a similar effect was observed with the PRC-20%. From
Figure 5-9 (middle), it was also evident that SeCR soil properties were not
conservatively designed, as utilities might have thought. Their selection included
some risk which, from the results in Figure 5-9, seems to be slightly greater than the
risk observed with PCR-20%.

120
120 SCR
SCR SeCR
SeCR 0.35
0.35 160
160
120
120 SCR
SCR SeCR 0.35PCR-20%
SeCR 0.35 PCR-20%
PCR-20%
PCR-20% PCR-60%
PCR-60%
PCR-60%
PCR-60% 160 PCR-100%
160 PCR-100%
PCR-100%
PCR-100%
0.3
0.3 140
140
140
100
100 0.3
0.3 140
100
100
3hrs/yr)
3hrs/yr)

120
120
(GWh/yr)
(GWh/yr)

3£/yr)
0.25
0.25 3£/yr)
(1033hrs/yr)
(10hrs/yr)

120
120
EWPC(GWh/yr)
(GWh/yr)

3£/yr)
3£/yr)
80
80
80
80 0.25
0.25 100
100
0.2
0.2 100
100
60
60 0.2
0.2 80
80
(10
(10

60
60 80
80
(10
ECUP(10

0.15
0.15
EECA(10
EECA(10

40
40 0.15
0.15 60
60
ECUP

60
60
EWPC

40
40 0.1
0.1
ECUP
ECUP
EWPC
EWPC

40
40
EECA
EECA

0.1
0.1 40
40
20
20
20
20 0.05
0.05 20
20
0.05
0.05 20
20
00 0 0 00
00 00 00
11 1.2 1.4 1.6 1.8 2
1
11
1 1.2
1.2
1.21.4
1.2 1.4
1.41.6
1.4 1.6
1.61.8
1.6 1.8
1.8 2
1.8 22
2 1
11 1.2
1.2
1.21.4
1Wind
1.2 1.4
1.4
Wind1.41.6
1.6
1.61.8
1.6 1.8
1.8 2
1.8 222 11 1.2
1.21.4
1.2 1.41.6
1.4 1.61.8
1.6 1.8 2
1.8 22
WindLevel
Wind Level
Level(pu)
Level (pu)
(pu)
(pu)
Figure 5-9 EWPC, EECA and ECUP outputs of SCR, SeCR and PCR 20%, 60% and
100% methods under a range of wind farm integration capacities.

When the risk of cable loading is further increased with the implementation of PCR-
60%, an additional 6GWh of wind power is integrated in the network. The PCR-100%
increased the wind power by 18GWh per year, which is approximately 17% less
curtailment than the SCR. However, the benefits from SeCR and PCRs come with the
expense of additional cable aging, which increases at higher excursions. Nevertheless,
the aging increase does not have a significant impact on the cable when compared to
the SCR’s expected value. This was indicated by the improved economic profits
achieved at higher Pex values. PCR-100% results in a maximum profit of £150k/yr
when compared to SCR at 2 pu capacity (see Figure 5-9 right).

Figure 5-10 shows the risk of thermal exceedance as a percentage of the operation
hours per year (left) and the cable’s expected equivalent aging in hours (right) for

142
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

excursion times by up to 100%, when assuming a 2 pu wind farm capacity. The figure
shows the mean, maximum and minimum values produced by the analysis of 60
simulation years. The analysis considered the uncertainties of annual weather
changes and wind speed variations. A zero exceedance of the cable’s design
temperature resulted from when excursion values of up to 60% were considered
(Figure 5-10, left). This is because the wind farm output variations and cable system’s
thermal inertia allows for heat dissipation. Consequently, there was a higher carrying
capacity compared to SCR without exceeding the design temperature.

55 Mean
Mean 1.2
1.2
Max
Max Min
Min
55 Mean
Mean 1.2
1.2
Max
Max Min
Min
11
(%)
(%)

3hrs/yr)
3hrs/yr)
44 11
Exceedance(%)
Exceedance(%)

3hrs/yr)
3hrs/yr)
44
0.8
0.8
33 0.8
0.8
Exceedance
Exceedance

33 0.6
0.6
0.6
0.6 (10
(10
22 EECA(10
EECA(10
100%
100%
22 0.4
0.4 100%
100%
100%
100% 0.4
0.4
100%
100%
EECA
EECA
11 0.2
0.2
11 0.2
0.2
00 00
00 00
00 20
20 40
40 60
60 8080 100100 00 20 20 40
40 60
60 80
80 100
100
00 20
20 40
40 60
60 8080 PCR
100
100
PCRExcursion00(%)
Excursion (%)20
20 40
40 60
60 80
80 100
100
PCR
PCRExcursion
Excursion(%)
(%)
Figure 5-10 Risk of thermal exceedance and EECA across different PCR-Pex%s
with 2.0 pu of wind farm capacity considering 60 years of wind and soil data.

However, the cable was operating close to 90°C more frequently, hence the mean
value of EECA was increased to 203 h at 60% excursion when compared to 75 h at
SCR (Figure 5-10, right). High excursion times still lead to a small risk of thermal
overloading in certain windy years with the most severe loading observed for PCR-
100% at a maximum of 1120 h loading at 90°C. However, the mean value of EECA was
317 h/yr. It is important to note that this increased aging is totally tolerable
considering that the cable is manufactured to operate at 90°C for the entire year, thus
a design EECA of 8760h/yr is expected. The PCR thus provides a promising
opportunity for most existing wind farms to connect more wind turbines by replacing
conventional SCR with the PCR method with a minimal risk of cable aging.

The increased wind integration flexibility generated by FCR has been shown in Figure
5-11. This analysis focused on FCR with pre-defined soil properties based on static,

143
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

seasonal, spring and winter conditions at an increased wind capacity range of up to


2 pu. These flexible cable ratings were compared against the DCR method since the
flexibility was gained through the transient state of cable temperature, also
considered in DCR. The additional wind integration was quantified by the EWPC,
while the risk of aging (EECA) was used to quantify the complete ECUP benefits and
risks of the rating methods.

80
80 FCRStatic 33
FCRSeasonal FCRWinter
FCR380 DCR
380
Spring

2.5
2.5 330
330
EWPC (GWh/yr)
(GWh/yr)

(103hrs/yr)
hrs/yr)

60
60

(103£/yr)
£/yr)
22 280
280
230
230

3
40
40 1.5
1.5

ECUP (10
3
EECA (10

180
180
EWPC

ECUP
11
EECA

20
20 130
130
0.5
0.5 8080
00 00 3030
11 1.2
1.21.4
1.41.6
1.61.8
1.8 22 11 1.2
1.21.4
1.41.6
1.61.8
1.8 22 11 1.2
1.21.4
1.41.6
1.61.8
1.8 22
Wind
WindLevel
Level(pu)
(pu)
Figure 5-11 Calculated EWPC, EECA and ECUP for standard SCR, DCR, and FCRs
with various predefined soil properties and increased installed wind capacity.

The results in Figure 5-11 illustrate that FCRs and DCR further reduce wind power
spillages, and result in increased profits when compared to the previously
investigated SeCR and PCR methods. FCRs were also shown to provide a significant
increase in wind integration, but also higher cable thermal aging. In particular,
FCRStatic considered the most conservative soil properties based on ISCR, resulting in
the highest wind curtailment, lowest cable loading (at 90°C) and lowest profit. On the
contrary, FCRWinter, FCRSpring, FCRSeasonal applied less conservative soil conditions
leading to reduced curtailment at an increased expected cable aging cost. However,
the aging was still very low as the cable does not exceed 90°C in most cases and
therefore increased profits (ECUP) were produced.

FCRSeasonal had a similar performance to DCR in regards to both wind integration and
profits, while the DCR loaded the cable more frequently at 90°C. This was due to
ignoring the increased risk of failure that is expected, as a result of increased aging,
which would come with the associated interruption costs of older worn-out cables.
The additional risks and their corresponding costs are expected to be higher for the

144
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

FCR methods due to weather and soil condition uncertainty, and were thus included
in the long-term economic assessment (Section 5.2.4).

To account for the uncertainty of the weather conditions, the analysis of loading the
cable circuit based on FCR methodology considered the weather data over a 60 year
period. Thus, different aging risk levels in every year were produced as the wind
outputs and soil conditions changed. This variation has been shown in Figure 5-12,
with the mean (μ), one and two standard deviation (σ) risk ranges of thermal
exceedance and EECA at double wind capacity and FCRSpring cable adequacy
modelling.

45% μ μ
35
35
3545%
35 μ μ ±±7777σ
σ μ
μ ±± 2σ

40% 6
30
30
3040%
30 66
6
3 3hrs/yr)
3hrs/yr)
(%)
(%)

35%
(103hrs/yr)
hrs/yr)
Exceedance(%)
(%)

25
25
25
35%
2530% 5
55
5 97%
97%
(%)

97%
97%
Exceedance
Exceedance(%)

30% 97%
97% 4
44
Exceedance
Exceedance

20
20 97%
97%
20
2025% 4 78%
78%
Exceedance

(10

25% 3 78%
78%
15
15
33
EECA(10
(10

15
1520% 68%
68%
68%
68% 3
20%
EECA

10
10 2
22
EECA
EECA

1015%
10 15% 2
55
5510% 1
11
1
10%
00
00 5%
5%
0
00
0
0 0
00%
0
0%
5
55
5 10
10
10
10 15
15
15
15 20
20
20
20 0
00
0 5
55
5 10
10
10
10 15
15
15
15 20
20
20
20
0% 20% 40% 60% Pre-defined
Pre-defined
80% Soil
100%Soil
SoilTemperature
Temperature
Temperature(°C)
(°C)
0% 20% 40% 60% Pre-defined
Pre-defined
80% Soil
100% Temperature (°C)
(°C)
Figure 5-12 Mean
FCRµ,Excursion
µ±σ (red)
FCR Excursion
(pu) and µ±2σ (blue) ranges of the cable’s thermal
(pu)
exceedance and EECA, with varying soil temperature FCR and 2 pu wind
capacity.

As shown in Figure 5-12, when the standard spring soil temperature (15°C from
Table 5-2) was used to determine the FCRSpring, less cable aging risk was generated.
When using 10°C temperatures, which is the standard for winter, a small increase in
aging was observed due to the thermal inertia of the cable system. However,
implementing less conservative temperatures below 10°C with the FCRSpring led to a
sharper and constant increase in aging and thermal exceedance risks. This indicates
that the bulk of the flexibility produced from the cable’s thermal inertia was
consumed at 10°C modelling.

Furthermore, FCR cannot guarantee a fixed annual thermal exceedance, as with DCR.
Thus, excessive aging was observed when FCRSpring was employed with very generous
soil temperature reductions to less than 10°C. The red and blue ranges displayed in

145
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

Figure 5-12 represents the variations expected from the year-to-year weather
changes and this can help the operator to determine practices based on a multiyear
performance. For example, setting a maximum cable loading at 90°C of 2500 h/yr
equivalent as an operation target, an estimated soil temperature of 4°C, 7°C and 10°C
could be defined for the FCR based on the µ, µ+σ and µ+2σ aging curves, respectively.
These cases consider the 50th, 84th or 98.5th percentile of the weather uncertainties
in the 60-year period of this study.

On the contrary, DCR’s real-time temperature monitoring capability can minimize


these risks further, but at the expense of additional instrumentation and installation
costs, which could be significant in many cases of existing cable systems.

The FFCR was designed to further increase wind integration in the same circuit under
the assumption that the wind farm operator was willing to temporarily overload the
cable at an emergency rating temperature for a pre-defined period of time based on
the forecasted wind output. Figure 5-13 shows the EWPC, EECA and ECUP outputs
for FFCR with seasonal ratings (i.e., FFCRSeasonal) and a 1.5 pu capacity increase when
forecasted durations for up to 6 hours were implemented. The zero forecast duration
(i.e., tforecast=0) indicated the FCRSeasonal performance, when comparing the FFCR
against the FCR approaches.

15 100°C 14
14 110°C 120°C 320
15
15 100°C
100°C 14 110°C
110°C 120°C 320
120°C 320 130°C
130°C
130°C
12
12
12 310
12 310
310
3hrs/yr)
(GWh/yr)

3£/yr)

12
12
(1033hrs/yr)
hrs/yr)
EWPC(GWh/yr)
(GWh/yr)

10
(1033£/yr)
£/yr)

10
10 300
999 888 300
300
290
ECUP(10
(10

666 290
290
EECA(10
(10

666
280
ECUP
ECUP
EWPC
EWPC

444 280
280
EECA
EECA

333 270
222 270
270
000 000 260
260
260
000 111 222 333 444 555 666 000 111 222 333 444 555 666 000 111 222 333 444 555 666
Forecast
Forecast Duration,
ForecastDuration, t
Duration,tforecast (hrs)
tforecast(hrs)
forecast (hrs)
Figure 5-13 FFCR Seasonal performance assessment at different temporary
temperature limits and forecasted durations under 1.5pu wind capacity.

146
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

The results, in Figure 5-13 (Left), indicate a significant reduction of wind curtailment
when wind forecasting was employed because FFCR further increases the loadability
during windy hours by allowing for the temporary overloading of the cable. The
higher the temporary thermal limit θC,Temp, the lower the wind curtailment. However,
this results in an exponential increase in aging. The ECUP results (Figure 5-13 right)
suggest that FFCR at 130°C had the lowest profit due to excessive aging costs.

In addition, extending the forecast duration further ahead (in time) provides
guidance for future rating calculations related to reducing further wind curtailments,
but the method (i.e., seasonal FFCR) becomes less effective due to the forecasting
errors. For the examined system using the ARMA approach, the maximum profit out
of the various FFCRSeasonal forecasted scenarios (from Figure 5-13 right) was gained
at 110°C’s temporary thermal rating with a 6hr forecast duration. This is defined as
the most optimal forecasting approach (i.e., FFCRSeOptim).

The long-term benefits and risks of the proposed flexible rating methods (PTR, FCR,
and FFCR) were compared against standard (SCR and DCR) practices, considering a
1.5 pu installed wind farm capacity as in Table 5-4.

Table 5-4 Key Performance Indicators of the Different Rating Methods under
1.5pu Wind Capacity
Annual Mean Performance Indicators
Rating ECAmax**
Cable Rating Wind Transfer Capability* EECA
Method (hrs/yr)
(A) (GWh/yr) (hrs/yr)
SCR 745 175.2 51 155
SeCR 776 180.0 94 311
PCR-100% 859 190.1 204 744
DCR 955 213.9 1218 1988
FCRSeasonal 948 212.2 1022 1901
FCRWinter 959 214.4 1357 2579
FFCRSeasonal 971 217.8 3839 6032
*Wind Transfer Capability implies the average annual wind power delivered
**ECAMax indicates the maximum equivalent cable aging risk in a single year

147
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

Mean performance indicators were used to compare the deterministic approaches


(SCR) against more risk aggressive and flexible approaches that have a variable rating
(as they are dependent on wind and soil conditions). The maximum cable aging risk
(ECAMax) was also calculated for the 60-year period as an indication of the upper-risk
limit expected from the rating method employed.

As seen in Table 5-4, it is evident that an increased wind transfer capability results in
increased aging risk due to the higher annual mean loading of the cable. However,
even the maximum values (i.e., 3839 or even 6032) are smaller than the normal
(manufacturer specified) EECA risk of 8760 hrs/yr. When comparing the mean aging
risk and maximum experienced risk of the standard practices (i.e., SCR and DCR) with
the FCRSeasonal approach, the latter appears to be the most optimum approach to
implement in terms of risk, implementation complexity and wind integration.
Conversely, FFCR produces almost three times more aging than the familiar DCR
practices and thus, there is a higher cable failure risk. The FFCRSeOptim as defined
earlier might not be a wise option, except when plans to replace the cable are in
development.

The economic benefits of the proposed flexible rating methods were examined in a
long-term 30-year analysis, against the STR, DCR, and traditional network
reinforcements with an additional identical circuit, as in Figure 5-14. This long-term
analysis also considered the expected failures of the cables due to the aging that
occurred in the annual steps as described by (5.8)-(5.11). This, consequently,
captured the complete cost and benefits of overloading the cable under the FFCR
method.

In Figure 5-14, the profits of the main uprating schemes were compared against the
SCR and traditional reinforcement, as these two define the least and most expensive
options. The modelling considered an increase of 50% of the wind farm’s installed
capacity and a 30-year duration. The red area in Figure 5-14 considered the weather
uncertainties in the FCR and FFCR methods when the µ±σ cable aging was considered
in the analysis.

148
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

9
99 PTR-100%
PTR-100%
PTR-100%
DCR
DCR
DCR µ±σ
µ±σageing
µ±σ ageing
ageing
FCR
FCR Seasonal
FCRSeasonal
Seasonal
6
66 FFCR FFCR FCR
FFCR
6£)

FFCRSeasonal
FFCR FFCR FCR
FCR
(1066£)
o (10 £)
Seasonal
SeOptim
Seasonal
Reinforcement
o))o)(10

Reinforcement
Reinforcement
P(Y

3
ProfitsP(Y
P(Y

33
Profits
Profits

SCR
SCRLevel
SCR Level
Level
0
00 Equipment
EquipmentCost
Equipment Cost
Cost

Reinforcement
ReinforcementCost
Reinforcement Costof
Cost ofofAdditional
AdditionalCable
Additional Cable
Cable
-3
-3-3
0
00 5
55 10
10Operating15
10 15 Y (yr)20
15 20
20 25
25
25 30
30
30
Operating Year
OperatingYear
YearYooYo(yr)
(yr)
Figure 5-14 Profits generated from various uprating schemes against SCR
when a 50% increase in wind power integration is considered across a 30 yr
operation.

DCR appears to provide constant economic gains and started to make profits after
year 5, despite the high initial cost of the ANM-DTS system. The reinforcement
achieved the highest economic gains in long-term operations of more than 25 years,
regardless of the large initial investment to install another identical cable. FFCRSeOptim
achieved better economic benefits than the other rating methods in the first 20 years
with zero initial implementation costs. FFCR profit slowed down after the 20 years
due to the surging maintenance and interruption costs, in the worst µ+σ case scenario.
The increased risk of FFCRSeOptim, mainly due to the forecasting uncertainty, makes it
less beneficial for integrating additional wind when compared to FCR. However, it is
the most flexible and easily implemented method along with the FCR, at least until a
more permanent upgrading option is provided.

This chapter has proposed flexible rating options that provide simple, directly
applicable and attractive methods to uprate existing wind farm cable circuits. The
proposed flexible approach provided a significantly improved loadability compared
to the existing static rating practice by considering the loading history, the
intermittent wind nature, and the cable’s thermal transient response. This allows the

149
Chapter 5: increasing Wind Power Integration through Flexible Cable Current Ratings

wind farm developers to increase their generation capacity by unlocking the spare
headroom capacity available, which could increase wind integration and negate or
postpone the network upgrade and the reinforcement of the existing cable circuits.

The flexible rating options were assessed with a methodology that calculated the
transient cable rating, considering the cable system design parameters and hence its
thermal inertia, as well as the effect of thermal aging on the cable’s life expectancy.
This allowed for a holistic evaluation with regards to both the capacity of wind
integration and the risk of cable aging. In this way, the rating options that do not use
a complex monitoring equipment installation, which is a typical requirement for
dynamic cable rating, could be assessed.

The analysis of the single-tie wind farm cable studies in Scotland has shown that a
less conservative probabilistic cable circuit rating for the wind farms can increase
wind integration by 8.5% when compared to the conventional static ratings without
any significant cable thermal aging risk. However, the best performing cable circuit
rating option was the FCR, allowing for an additional 21% wind power injection in
the system, when compared to SCR. This method does not exceed the aging risk that
is expected under the existing DCR method as implemented in the single-tie wind
farm cable studies. The FFCR method was found to be too risky to implement on the
wind farms, with only a small additional wind integration benefit. However, in cases
where the wind is more periodic and simpler to forecast (i.e., off-shore wind farms),
this option might be worth considering.

The proposed methodology allows the host utilities and wind farm developers to
determine the optimum amount of increased wind generation capacity, and to select
the optimal rating strategy based on short or long-term economic prospects and the
operators’ acceptance on cable aging. Further work is required in order to more
accurately quantify the forecasting error and cable degradation with the operating
temperature.

Equation Chapter (Next) Section 1

150
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

6 Quantifying the Risk of Emergency Ratings


on Distribution Cable Network with EVs

One of the key challenges facing distribution network operators (DNOs) today is
the expected increase in electric vehicles. The increased load from EV charging will
result in distribution assets becoming 'thermally overloaded' due to higher
operating temperatures. In addition to the issue of increased load, we have a
limited understanding of the behaviour and performance of the distribution assets
and their potential to accept the increased load. It has been well acknowledged that
EVs increase the network loading level, leading to a reduced system reliability
performance. These results have not been quantified in a realistic case study,
including actual cable rating and design properties.

To address this gap, this chapter introduces an additional case study in a


distribution all cable network using the existing network reliability evaluation
framework, which quantifies the impact of different EV penetration levels on
distribution network reliability, as well as the thermal performance of distribution
cables. The novel load profiles can be constructed by adding an EV charging profile
to household loads and solving the power flows to assess the network impact. This
study has also investigated the benefits of emergency cable ratings and remote
switching to reduce the peak power demand caused by EV, in order to reinforce the
reliability of the grid and to boost the maximum allowable EV penetration in the
distribution networks.

The methodology was applied using a case study on the modified EV-integrated
RBTS bus 4 distribution. The results showed that the negative impact of EVs on
network performance can be mitigated by the implementation of emergency cable
ratings. The peak demand under contingencies can also be accepted by the
emergency loading of cables. The frequency and duration of EV demand
interruption was also significantly reduced. Thus, a higher EV penetration can be
accommodated.

151
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

The rapid development of electric vehicles (EV) has caused heavy stress to
conventional power grids. The peak power demand caused by EV charging requires
a more flexible and intelligent network, posing a great challenge for future smart
grids. As the rapid development of EVs, DSOs will be responsible for upgrading their
existing network in order to supply the EV charging points. Therefore, utilities are
investigating innovative solutions to reinforce their existing distribution networks to
satisfy the increasing demand of EVs [108].

However, most existing urban distribution networks are designed solely based on a
typical residential load cycle, without consideration for massive EV connections and
random EV charging patterns [18, 109]. Most existing distribution networks are more
than 30 years old in the UK, which is operating close to the thermal limit after years
of demand growth [110]. DNOs initially planned to replace and upgrade most of their
network within the next 20 years, gradually due to budget constraints [92]. However,
the expected EV market share will be 25% by 2020, and certain geographical clusters
will have a higher penetration of EVs [111]. Therefore, DNOs are forced to develop
economic solutions to enhance the network capacity for increased EV integration
over the next 5-10 years.

The impact of EV charging at distribution networks, including component


overloading, feeder congestion and undue faults. Different aspects of the impact, such
as energy loss, voltage profiles, the reduced lifetime of the network components, the
thermal loading of cables and transformers, are considered in some studies [112,
113]. However, limited work has been published to quantify the impacts on the
feeder cables in the presence of EV charging.

Controlled smart charging techniques or Vehicle-to-Grid (V2G) technology aids the


power grid by reducing the peak load and transmission congestion, minimising their
effect on the distribution system assets and reducing investments in the network
reinforcements [114-116]. However, similarly to distributed generation (DG), DSOs
would have no control over the location of future EV charging points or stations, and
no direct control over the period and frequency of EV charging [113]. How to

152
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

schedule the charging for each EV based on their battery state and their availability
for charging is a key question. As a consequence, uncontrolled stochastic EV charging
would be more common for the majority of situations in the early future.

Several items of literature describing new business models for the successful
integration of EVs into the distribution network have been published [113, 117].
Public aggregated charging points and charging stations infrastructures, called
“aggregators”, can be developed by intermediate service providers, while distributed
charging infrastructures will still be installed at each household. Compared to
distributed EVs, EV aggregators can be coordinated by DNOs to provide an ancillary
service at the appropriate large-scale power system level. However, limited literature
has considered the existence of both aggregated and distributed charging in relation
to distribution network operations and planning.

During normal operations, medium voltage radial distribution circuits reserve


around 50% of their total capacity for emergency operations. The reserved capacity
can be utilised to accommodate more EVs. However, in a restructured network
during emergency restoration, distribution lines may operate close to their capacity
limits or even overloaded due to the increased load of EV charging. The utilisation of
the emergency rating on the distribution cables provides more network flexibility,
and reduces the interruption of EV charging demands during emergency restoration.

Automated switches are becoming a key component in electric distribution systems


[118]. These remote switching devices can also be utilised in an EV integrated
distribution network, to perform corrective EV load shedding actions in response to
sensing an overload condition, or by receiving control signals from the network
operator. However, no existing literature has considered employing emergency
uprating or remote switching to solve the problem of feeder congestion due to EV
connections.

This study has utilised the overall framework integrating cable design and ageing to
evaluate the impact of EV charging on increased thermal loading and the thermal
ageing of feeder cable circuits, as well as the overall distribution network reliability.

153
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

The aim of this methodology is to achieve a better prediction of the network’s


reliability and distribution cable performance when a high penetration of EVs is
integrated into the existing distribution network. Another aim of this study was to
quantify the risks and benefits of emergency loading of cables could give in an
distribution cable network with EVs.

6.2.1 Overview of the Proposed Methodology

To quantify the impact of electric vehicle charging on the behaviour and performance
of the distribution assets, as well as the overall distribution network reliability, the
proposed SMC-based network reliability evaluation framework was integrated with
asset failure and repair modelling, EV charging and demand restoration modelling,
and distribution network power restoration modelling. The outline of the proposed
methodology has been shown in Figure 6-1.

Network
EV Demand Health Score
PLANT
Operational Data Reliability Data Weather Data Branch Data

Power System Data Initialization

Initialize asset status mapping, TTF, TTR, fault category

Capture asset failures, and unsupplied customers

Power restoration by switching


Pre-restoration Post-overload Remote
EV-switch mode EV-switch mode EV-switch mode
Annual
analysis Post-contingency operation using emergency rating Time step
〈Y=Y+1〉 analysis
〈tΔt=tΔt+Δt〉
EV charging demand restoration

Thermal-ageing model of UGCs


NO
Annual computation is completed
Inner SMC-loop
YES
Annual analysis is completed until specified Cov Outer SMC-loop
NO
YES
Network reliability & Branch performance indices calculation

Completed network reliability evaluation


Figure 6-1 Methodological flowchart for evaluating the EV impact in the
distribution network

154
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

The methodology initially started by generating a time-to-fail (TTF) and time-to-


repair (TTR) for each asset considering the characteristics of the internal and
external fault categories. The health score for each asset determined the internal
failure rate for the distribution assets using the UK DNO common network asset
indices methodology, which has been well explained in Section 6.2.2.

For the distribution networks integrated with EVs, new load profiles need to be
constructed by adding different EV charging profiles to the original household loads.
EV charging profiles for residential & non-residential EV parks were modelled
according to the EV arrival pattern and charging duration pattern, both of which
follow normal distribution to consider the randomness of EV charging demand. The
detailed modelling of EV charging profiles has been described in Section 6.2.3.

The methodology was used to perform the inner SMC-loop for the time step analysis,
and the outer SMC-loop for the annual analysis. Within each time step analysis, the
methodology captured whether or not there was any asset failure, and any customer
without supply. A network reconfiguration algorithm was then used to restore supply
to as many customers as possible using switching actions. Once the restoration is
established, distribution cables on the neighbouring feeder usually operate close to
their capacity limits and in some occasions could be temporarily overloaded. In order
to avoid overloading, the EV charging demand needs to be shed, thus three
operational modes under different contingencies were designed for aggregated EV
charging parks: Pre-restoration EV-off mode, Post-overload EV-off mode and Remote
switch mode. The detailed description of the three models has been provided in
Section 6.4.

In the restructured distribution network, emergency ratings could be employed to


allow for increased loading during emergency restoration. This study considered the
three different UGC ratings (normal rating Inormal, cyclic rating Icyclic, and emergency
rating Ieme) to indicate the impact of emergency rating employment on network
reliability, EV interruption reduction, and cable ageing.

155
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

If some of the EV charging demand was interrupted during any network


contingencies, the EV battery will continue to be charged until it is full after the
interruption finishes, thus the interrupted EV charging demand needs to be restored.
The detailed modelling of the EV charging demand restoration procedure has been
described in Section 6.2.3.

This study introduced an EDNS (EV Demand Not Satisfied, MWh/yr) index that
represents that the charging demand of the EVs might not be fully satisfied since less
energy can be charged into the batteries during a period of emergency dispatching.
Although all interrupted EV charging demands will be restored after the contingency,
the EDNS index indicates the quality of the EV charging service.

NY
EDNS   SysDNSY NY (6.1)
Y 1

The frequency and duration of the EV charging demand interruption was captured
for every EV charging point using the EFEI (Expected Frequency of EV charging
demand Interruption, occ./yr) and EDEI (Expected Duration of EV charging demand
Interruption, hrs/yr) index. The mathematical equations for EDNS, EFEI and EDEI
have been stated in (6.1), (6.2) and (6.3), where SysDNSY is the EV charging demand
not satisfied during the system contingency in year Y, DEIP,Y is the duration of EV
demand interruption at EV charging point P, in year Y, FEIP,Y is the frequency of the
EV demand interruption at EV charging point P, in year Y.

NY
EDEI P   DEI P ,Y NY (6.2)
Y 1

NY
EFEI P   FEI P ,Y NY (6.3)
i 1

6.2.2 Asset Failure and Repair Modelling

Since most cable failure is caused by digging activities instead of internal failures, this
study considers there to be two types of failures for feeder cables; failures caused by
external human activities and failures caused by internal ageing reasons. Overhead

156
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

lines and transformers consider there to be two types of outage; internal failure and
external maintenance outage.

The internal failure rates λinternal for the distribution assets were computed based on
UK DNO Common Network Asset Indices Methodology. A health score (HS) between
0.5 and 10 is allocated to each network component. The smallest HS=0.5 and the
largest HS=10 represents a component in very good and in a very poor condition
respectively. The health score of a new component initialises at HS=0.5, and increases
along with time due to ageing. A third order polynomial relates failure rates to the
health scores, as shown in (6.4), where HS is a variable equal to the Health Score
unless HS≤4 then HS=4, K and C are the constants that are defined for each type of
asset in [8].

  C  HS 2  C  HS 3 
internal  K  1   C  HS     (6.4)
 2! 3! 

However, components with HS<4.0 are assumed to have a minimum HS=4.0 to


ensure that the failure rates are not initially too low. According to [8], the constants
for different distribution assets have been stated in Table 6-1.

Considering the randomness of the internal failures, an exponential distribution was


used to calculate the time-to-fail, TTFinternal, and time-to-repair, TTRinternal, values,
using the internal failure rate, λinternal, and mean repair duration, Dinternal, with an
unavailability, U, randomly generated from a uniform distribution U(0,1) [3].

TTFinternal  (1 / internal )ln U (6.5)

TTRinternal   Dinternal ln U (6.6)

As provided in Table 6-1, the external failure rates for the UGCs and maintenance
outage rates for OHLs and transformers were set to be constant throughout the asset
lifetime [119]. The mean repair durations for the internal and external failures of the
distribution assets have been provided in [119].

157
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

Table 6-1 Reliability parameters of the distribution assets in the modified


RBTS
Parameters 11kV 33kV 33/11kV
UGCs OHLs transformers
K 0.000658 0.001006 0.000454
C 1.087 1.087 1.087
λinternal [occ./yr] Eq.(6.4) Eq.(6.4) Eq.(6.4)
Dinternal [hrs/occ.] 30 8 15
λexternal [occ./yr] 0.05 0.5 1.0
Dexternal [hrs/occ.] 10 8 120

For cable external failures and maintenance outages, time to failure, TTFexternal, was
generated in (6.7) using an exponential distribution. The time to repair TTRexternal was
calculated in (6.8) for the asset subject to a normal distribution Nor with a standard
deviation SD=25%.

TTFexternal  (1 /  external )ln U (6.7)

TTRexternal  Nor  Dexternal , SD  (6.8)

25% is a reasonable assumption that could be redefined as further data becomes


available. Using normal distribution limits the extent to which repair durations vary
dramatically from the average [120]. This is appropriate, as the repairs of the cable
failures always involve a certain amount of work, such as access to the site,
performing replacements or maintenances, and then testing repairs. However, the
impact of long repairs related to internal failures can only be captured by exponential
or Weibull distributions.

6.2.3 Electric Vehicle Charging and Restoration Modelling

3) EV charging modelling

The EV charging pattern is an important component in EV modelling, and it may


result in a significant impact on the distribution system reliability. Controlled
charging strategies, such as Vehicle-to-Grid (V2G) technology and smart charging
technology, were not considered in this methodology. Under an uncontrolled
charging strategy, an EV will consume charging power after it is parked until its
battery is fully charged.

158
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

In this study, the EV charging pattern was modelled by the EV arrival pattern and
charging duration pattern. The charging duration was related to the power
consumption in the last drive, namely the charging power of EV battery. Each
individual EV was assumed to have a 20kWh capacity and a charging speed of 2kW
(240V 1.2A), using the example of the ‘Chevrolet Volt’ model. It is assumed that EVs
are always charged with a constant power draw that is equal to the charging speed.

10
9 Residential weekday
Number of EVs to arrive

8 Residential weekend
7 Non-residential weekday
6 Non-residential weekend
5
4
3
2
1
0
0.6
Charging Demand MW

0.5
0.4
0.3
0.2
0.1
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24
Hours
Figure 6-2 EV arrival pattern and average charging power demand pattern

Figure 6-2(Upper) illustrates the average number of EVs to arrive, N_EVavg(tΔt), for
every 100 EV charging slots in residential areas (distributed or aggregated parking)
and non-residential areas (commercial or industrial parking lot) during weekdays
and weekends [121]. When an average charging duration of 3 hours was assumed,
the average charging power consumption profile for every 100 EV charging slots has
been calculated and displayed in Figure 6-2(Lower).

Considering the randomness of the EV arrival pattern changes, the actual simulated
number of arrived EVs, N_EVP(tΔt) in parking lot P, at time tΔt, was calculated (6.9)
subject to a normal distribution Nor with a standard deviation SD=20%, where
N_EVSP was the number of EV charging slots in parking space P, and N_EVavg(tΔt) was
the number of EVs to arrive per 100 EV charging slots at time tΔt.

N_EVP  tt   Nor  N_EVSP N_EVavg tt  , SD  (6.9)

159
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

The actual simulated charging power consumption P_EVP(tΔt) at parking lot P, at time
tΔt, is calculated in where Pcharging is the EV charging speed, D_EVP(tΔt) is the charging
duration of EVs parked in parking lot P, at time tΔt.

D_EVP  tt 
P_EVP  tt   Pcharging × h 0
N_EVP  tt -h  (6.10)

4) EV demand restoration modelling

This study considered an uncontrolled charging strategy, which means that an EV will
be continuously charged until a full battery is reached. However, EV charging in EV
parking lots may be interrupted deliberately by the network operator due to circuit
overloading under emergency contingency conditions. The interrupted EV batteries
will continue to be charged until they are full, thus EV charging demand needs to be
restored once the network contingency ends and the service is back online. The EV
demand restoration modelling and constraint criteria have been described by the
following equations (6.11) and (6.12).

The new EV charging demand P_EVP,new(tΔt) of parking lot P at tΔt during the
restoration period is the sum of the original EV charging demand P_EVP (tΔt) and the
restored EV charging demand P_EVP,res(tΔt), as expressed by (6.11).

P_EVP,new  tt   P_EVP  tt   P_EVP,res tt  (6.11)

The restored EV charging demand P_EVP,res(tΔt), can be calculated by (6.12), where


P_EVP,int,total is the total interrupted EV charging demand, N_EVP,int,total is the total
number of interrupted EVs, N_EVP,res is the number of EVs which were fully restored,
t0 is the time when the restoration started. This equation indicates that the restored
EV charging demand is constrained by the EV charging speed, the number of EVs that
still needed to be restored, and the total demand that still needs to be restored.

160
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

 tt

 P _EV P ,int ,total   P_EVP ,res (t )


 t  t0
P_EVP ,res (tt )  min  (6.12)
P  tt

 charging   N _EV P ,int ,total   N_EVP ,res (t ) 
  t  t0 

6.2.4 Network Restoration Modelling

In distribution networks, transformers, lines (OHLs and cables), busbars and


breakers may incur a failure or undergo scheduled maintenance. In response to both,
suitable network restoration actions should be taken to restore service to as many
affected customers in the out-of-service areas as possible using both switching
actions and cable emergency loading. The traditional procedure for power
restoration in a typical two-feeder open-ring network as described in the steps below.
The affected area around the fault point can be divided into repaired, upstream and
downstream areas by the switches, as illustrated in Figure 6-3(Left):

Feeder 1 Feeder 2 Feeder 1 Feeder 2

Restored
Upstream
Area AEV SS AEV SS Cable
Emergency
Uprating

Repaired Fault Fault


Area

Restored
Downstream
Area
AEV SS AEV SS

Tie-Switch Tie-Switch
Figure 6-3 An example of power restoration in a two-legged ring network with
or without EV connection

161
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

1. Protection response: protection devices (a fuse or circuit breaker) nearest to the


fault will operate to trip the failed circuit and clear the fault in seconds.
2. Upstream restoration: after the fault location is identified, sectionalising devices
(e.g., disconnecting switches) are operated to isolate the fault. The supply to
upstream area customers is restored through the main supply by reclosing the
upstream circuit breaker after the fault’s isolation.
3. Downstream restoration: the downstream customers are restored by closing the
tie-switch through the supply from the neighbouring feeder. Isolation and
restoration takes seconds if there are automated switches, or hours if there are
manual switches.
4. Repairing process: after the faulty components have been repaired, the supply
can be restored in the repaired area. The disconnecting switches used for
isolation can be closed and the tie-switch opened, so then the network can return
to its pre-fault state.

After power restoration is established for a distribution network with EV connection


in Figure 6-3(Right), distribution cables in the neighbouring feeder usually operate
close to their capacity limits and in some occasions, could be temporarily overloaded.
If network constraints such as line capacity constraints cannot be met, alternative
actions will be undertaken to satisfy the constraints, such as utilising the emergency
ratings of feeder cables, or abandoning an increased load from aggregated EV (AEV)
charging using switching. This study considered three EV operating modes for
restructuring distribution networks with an EV connection under the following
network contingencies: Pre-restoration EV-switch mode, Post-overload EV-switch
mode and Remote EV-switch mode.

(1) Pre-restoration EV-switch mode

Pre-restoration EV-switch mode switches off all aggregated EV charging parks


connected to the neighbouring feeder before the restoration is established, and keeps
them disconnected until the repair is finished and the network returns to its pre-
contingency structure. For example in (Right), two AEV parks on feeder 2 and one
AEV park in the restored downstream area will be disconnected.

162
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

(2) Post-overload EV-switch mode

Post-overload EV-switch mode switches off all aggregated EV charging parks


connected to the neighbouring feeder once any distribution cable on the
neighbouring feeder is overloaded. It keeps them disconnected until the repair
finishes. In principal, this scheme sheds all charging demand from the AEV parks to
prevent the cables from further overloading, thermal ageing and damage. This mode
requires switching devices for EV parks to be opened in response to the monitoring
signal indicating cable overloading.

(3) Remote EV-switch mode

The Remote EV-switch mode aims to meet as many EV charging demands as possible
at the minimum cost. Its objective function and constraints have been stated in (6.13),
where ocrs,P is a binary operation variable indicating whether the remote switch rs at
EV park P is closed (=0) or open (=1); costEV is the cost of EV interruption in £/MVA;
costPL is the cost of the power losses in £/MVA; Ploss,i(tΔt) is power losses of cable i at
time tΔt; costENS is the cost of non-delivered demand in £/MVA and ENS(tΔt) is the
customer energy not supplied in MVA.

Cost  tt  =  ocrs ,P  cost EV  P _ EVP  tt  +  cost PL  Ploss ,i  tt   costENS  ENS  tt 
PN P iNUGC

subject to: -Smax, i  Si  Smax, i , Vmin, bus  Vbus  Vmax, bus

(6.13)

The only decision variables are operational variables ocrs,P, which allow the remote
switches to be opened or closed in real time by receiving control signals from other
locations. The optimisation allows the network operator to decide whether one or
more EV charging parks should be disconnected to prevent cable overloading, to
satisfy as many EV charging demands as possible, and to minimise the cost of power
losses and non-delivered energy.

163
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

The RBTS bus 4 distribution system is widely utilised in reliability studies for
distribution networks. This methodology utilises the original RTBS bus 4 network as
a realistic example of an existing distribution network in an urban area. The original
test network consists of 7 feeders (F1–F7), 38 load points (LP1– LP38), disconnecting
switches on both sides for each feeder cable and 4 tie-switches. Disconnecting
switches are positioned before and after each load bus on each feeder, to allow for
each load bus to be isolated from a fault. The tie-switches (TS1-TS4) are normally
open, while all other switches are closed when in a state of normal operation. Only
one circuit breaker is equipped at the top of each feeder. The default peak load of the
network is the actual 40MW and 13Mvar, including 4769 customers (9 industrial
small users, 70 commercial customers, and 4690 residential customers). The hourly
chronological load profile as described in [19] is used to describe the load profile for
this network.

Table 6-2 The Cable Electrical Properties Modelled in the RBTS bus 4
network
Summer Winter
Cable Size R90 X A A Feeder Cable
Type mm Ω/km Ω/km
2 Number
Inormal Icyclic Ieme Inormal Icyclic Ieme
A 300 0.062 0.098 683 772 849 706 798 878 1,6,9,14,19,22,25
B 240 0.077 0.101 608 687 756 629 711 782 2,7,10,15,20,23,26
C 185 0.100 0.105 525 593 653 544 615 676 3,8,11,16,21,24,27
D 150 0.124 0.108 467 528 580 484 547 602 4,5,12,13,17,18,28,29

6.3.1 Selection of the Feeder Cables

In order to evaluate the impact of feeder loading constraints on EV connection and to


exclude other influencing factors, the capacity of the overhead lines and transformers
was set high enough to not constrain the network even under the N-1 condition. The
RBTS bus 4 network does not provide thermal limits and circuit design data,
therefore the size and current rating of each feeder cable (numbered with C) was
carefully selected to ensure that all existing demands can still be fully supplied under
the N-1 condition through backfeeding from the normal open points. Four types of

164
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

single-core un-armoured XLPE cables with copper conductors were used in the study.
Table 6-2 shows the selected cable design data, normal and cyclic current rating for
each corresponding feeder cable. The cables were assumed to be laid in a trefoil
formation (touching) and directly buried in the ground at a depth of 0.8m. The XLPE
cables and soil thermal parameters used in this study have been listed in Table 6-3.

The cable’s normal current ratings, Inormal, as shown in Table 6-2 were calculated at a
maximum core temperature of 90°C, 0.9 km/W soil thermal resistivity, and 1.6×106
J/m3K volume specific heat, which are the default values in ENA P17 [17]. The
summer (Nov-Apr) and winter (May-Oct) ratings were calculated using a soil
temperature of 15°C and 10°C respectively, and cyclic ratings, Icyclic, that were 1.13
times the corresponding normal rating based on a cyclic loading factor [19]. For the
11kV cables forming a two-legged ring network, an emergency rating that is more
liberal than the cyclic rating, may be used for the design of such networks. As
indicated in [18], distribution utilities allow for an emergency rating, Ieme, which is
effectively 110% of the appropriate summer and winter cyclic ratings, to be used in
their 11kV UGCs as provided in Table 6-2.

Table 6-3 Thermal Parameters in XLPE Cables and Soil


Material Thermal Resistivity Volume Specific Heat
ρtherm [mK/W] Ctherm [J/(m3K)]
Copper (Conductor) 0 3.45×106
XLPE (Dielectric) 3.5 2.4×106
Copper Wire (Screen) 0 3.45×106
PE (Jacket) 3.5 2.4×106
Soil (Surroundings) 0.9 1.6×106

6.3.2 Modifications of the RBTS bus 4 network

Some modifications have been applied to the RBTS bus 4 network in order to simulate
a realistic scenario that aggregates the EV park lots and the distributed residential EV
charging piles that are both available in the distribution network. The modified RBTS
network has been shown in Figure 6-4.

165
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

As can be seen, distributed residential EVs are connected to each residential load
point. Four aggregated residential EV parks are connected directly to four 11kV buses;
bus 5, 23, 36, and 60. The aggregated residential EV (AREV) park on each feeder
consists of X% of the total residential EV parking slots, while the distributed
residential EV (DREV) charging facilities within each household own the rest.

11kV REV CEV


C9 C10 C11 SS C12 C13 SS
19 21 23 26 28

33kV F3
REV 20 REV 22 REV 24 25 REV 27 REV 29 30
LP11 LP12 LP13 LP14 LP15 LP16 LP17
IEV
SS
C6 13 C7 15 C8 17

F2 14 16 18
TS3
LP8 LP9 LP10
REV CEV
SS C5 SS
C1 1 C2 3 C3 5 C4 7 10

O F1
H REV 2 REV 4 REV 6 REV 8 9 REV 11 12
L LP1 LP2 LP3 LP4 LP5 LP6 LP7
TS2
33kV 11kV REV CEV
O C25 C26 C27 SS C28 C29 SS TS1
56 58 60 63 65
H
L F7
REV 57 REV 59 REV 61 62 REV 64 REV 66 67
LP32 LP33 LP34 LP35 LP36 LP37 LP38
IEV
SS
C22 50 C23 52 C24 54

O F6 LP29
51
LP30
53
LP31
55

H
L IEV
SS
C19 44 C20 46 C21 48
45 47 49
F5 LP26 LP27 LP28 TS4
REV CEV
SS SS
C14 31 C15 33 C16 36 C17 39 C18 41

F4
33kV REV 32 REV 34 35 REV REV 37 38 REV REV 40 42 43
11kV LP18 LP19 LP20 LP21 LP22 LP23 LP24 LP25

OHL Overhead Line Tie-Line CEV Commercial EV Park Lot


Bus Switch (TS)
Switch C1 Underground Tie-Line REV Residential EV Park Lot
Cable
Load Point Transformer SS Smart Switch IEV Industrial EV Park Lot
LP1
Circuit Breaker REV Distributed Residential EV

Figure 6-4 Single line diagram of the modified RBTS bus 4 distribution
network

166
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

This study assumes that two individual households exist in each residential customer
as defined in the original RBTS bus 4 network. Each household is assumed to own an
average of 1.5 vehicles, thus the total number of residential vehicles was 14,070. The
EV penetration level indicates the proportion of the number of EVs compared to the
number of total household vehicles in the area.

As shown in Figure 6-4, four commercial car parks with 1,000 parking slots were
integrated into terminal buses 10, 28, 41, and 65, with commercial customers. Three
industrial car parks with 2,000 parking slots were integrated into terminal buses 15,
46, and 52 on three short feeders with a small number of users. Commercial and
industrial parking lots provide parking spaces and EV charging mainly for workers in
offices and factories. The EV penetration level defines the proportion of EV charging
slots to the total number of parking slots in all commercial and industrial parking lots.

In this study, it is assumed that X%=50%, means that half of the EVs are charged at
AREV parks, while the other half are charged at DREV slots located at residential load
points. The EV penetration level is PEV%, which means that the PEV% of all parking
slots were EV charging slots. Thus, the number of EV charging slots on each bus has
been shown in Table 6-4.

Table 6-4 Number of EV charging slots on each bus at PEV% penetration


level
Type Number of EVs Bus No.
charging slots, N_EVSP
2, 4, 6, 8, 20, 22, 24, 32, 34,
Distributed residential 330×PEV%
35, 37, 57, 59, 61
EV charging piles
300×PEV% 9, 25, 27, 38, 40, 62, 64, 66
1620×PEV% 5
Aggregated residential
1590×PEV% 23
parking lots
1920×PEV% 36, 60
Aggregated commercial
1000×PEV% 10, 28, 41, 65
parking lots
Aggregated industrial
2000×PEV% 15, 46, 52
parking lots

In this study, only the feeder cables (numbered with C), 33kV overhead lines and
33/11kV transformers were modelled to be able to fail while all other network

167
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

components, such as the circuit breaker, switches, lateral branches and 11/0.415kV
transformers, were considered to be 100% reliable.

6.3.3 Modelled Scenarios

The proposed methodology was applied to the modified RBTS bus 4 network in order
to study the impact of increasing EV penetration on distribution network reliability
and cable thermal performance under different operating schemes. The scenarios
have been described in Table 6-5.

All of the scenarios consider a health score HS=10 for all distribution network assets
in order to model an aged distribution network. These scenarios were compared
against the base-case scenario (BC) to quantify any reliability reduction and
increased cable ageing due to the increased loads from the EV integration.

Table 6-5 Modelling Scenarios of EV impact on the distribution network


Scenarios EV penetration level Scenario Description
Base case No EV integrated
Sc-1 PEV% No EV-switch mode
Sc-2 PEV% Pre-restoration EV-switch mode
Sc-3 PEV% Post-overload EV-switch mode
Sc-4 PEV% Remote EV-switch mode

Sc-1 does not apply any of the EV-switch modes, where the EV charging demand is
considered to be the same type of load as normal loads (residential, commercial and
industrial). It is not considered to be a special demand that requires unique EV-switch
actions under emergency operations. Sc-2 is the most conservative scenario using the
pre-restoration EV-switch mode, which switches off the EV parks before the
restoration is established. Sc-3 implements the post-overload EV-switch mode, which
allows for the EVs to continue to be charged until overload occurs, and then the EV
parks are switched off until the contingency ends. Sc-3 applies the remote EV-switch
mode by implementing remote switching, monitoring and control devices, to allow
the EV parks to be switched on or off in real-time. The optimal EV switching solution
was determined to minimise the total cost of EV charging interruption, power losses
and non-delivered demand.

168
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

The impact of the EV integration on the network’s reliability and cable ageing can be
indicated by the comparison between the base case and Sc-1. Table 6-6 illustrates the
amount of expected energy not supplied (EENS), expected annual network losses
(EANL), expected equivalent network ageing (EENA), and the expected EV demand
not supplied (EDNS) when the EV penetration level increased from 0% in the base
case up to 100% in Sc-1. It is worth mentioning that the normal current rating, Inormal,
was utilised in all base cases and Sc-1 scenarios as in Table 6-6.

Table 6-6 Network performance indices of the base case and Sc-1
Scenarios Base case Sc-1
PEV% 0% 20% 40% 60% 80% 100%
EENS, MWh/yr 85.55 101.87 123.47 151.54 186.80 229.70
SAIFI, occ./yr 1.28 1.52 1.93 2.37 2.91 3.45
Indices

SAIDI, hrs/yr 8.21 9.29 11.40 13.87 16.64 19.77


EANL, MWh/yr 1362 1398 1498 1602 1711 1824
EENA, hrs/yr 71.7 76.1 83.9 92.7 102.9 114.6
EDNS, MWh/yr 0.00 0.76 1.57 2.45 3.51 4.86

By increasing the EV penetration PEV%, a continuous increase of load curtailment,


power losses and cable ageing was observed. In particular, the EENS rose from
85.55MWh in the base case with zero EV integration, and peaked at 229.70MWh at
100% EV penetration. The frequency and duration of the demand interruption also
increased by almost three times the maximum values of 3.45occ./yr and 19.77hrs/yr.
This indicates a significantly reduced network reliability performance when more
EVs are connected in the distribution networks. This worsen the EENS because EV
charging leads to an additional demand which adds to the existing distribution
network, resulting in the network failing to meet the N-1 criterion. When there are
no EVs, the load curtailment used does not occur due to a single contingency.
Appropriate switching and restoration actions can fully restore the interrupted
demand, therefore the EENS is relatively low when PEV%=0. With EV integration, the
additional EV charging loading exceeds the load constraint of the feeder cable circuits

169
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

and voltage constraints of the terminal buses, thus more load curtailments occur and
network reliability performance is reduced.

The impact of the EV penetration on the feeder cable circuits can be indicated by the
increase of EANL and EENA. The additional EV charging demand causes an increase
in current loading, and also a rise of power losses within the circuits. Based on the
basic theory of “Losses=I2R”, the EANL was increased by 33.9% from 1362MWh to
1824MWh, suggesting a 15.7% increase in the general loading of the whole
distribution network. This result will become worse, considering the natural annual
increase of the residential demand of electric vehicles.

The network cable ageing was almost doubled from 71.7hr/yr to 114.6hr/yr, when
the EV penetration was increased to 100%. This ageing increase seems to be slow
and tolerable for the cables, considering that the expected annual ageing is 8760hr/yr.
The reason for the ageing output is because no emergency and cyclic loading is
allowed, thus the cables operate at quite low temperatures at which negligible ageing
occurs. Sc-1, with the utilisation of Inormal, tends to sacrifice the demand when
network loading constraints are met, rather than utilise emergency loading to
provide extra flexibility and loadability to the network. Consequently, more load
curtailment is made but less ageing occurs.

Sc-1 does not apply any EV-switch mode, which means that the EV charging demand
is treated the same as any other demand. Along with the increase of EENS, the load
curtailment of EV charging demand, EDNS, was also increased by 4.86MWh. However,
the curtailment and interruption in the EV demand was much less costly compared
to regular residential, commercial and industrial demand, because the interrupted
EV demand can be fully restored after the service is recovered. Therefore, suitable
EV-switch actions should be made in order to achieve better network reliability at
the cost of cheap EV demand.

Various EV-switch modes were considered in this methodology in order to provide


DNOs with an alternative option to address the EV issues during network

170
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

contingencies. Table 6-7 illustrates the amount of expected energy not supplied
(EENS), the expected EV demand not supplied (EDNS), the expected duration of EV
demand interruption (EDEDI) and the expected equivalent network ageing (EENA)
across different EV penetrations PEV% from 0% to 100% when various EV-switch
modes were utilised. Normal current rating, Inormal, was still utilised in all scenarios
within Table 6-7.

Table 6-7 The output of EENS, EDNS, EDEDI and EENA in Sc-1, Sc-2, Sc-3 and
Sc-4 across different EV penetration levels
Scenarios EV penetration PEV%
EENS MWh/yr 0% 20% 40% 60% 80% 100%
Sc-1 85.55 101.87 123.47 151.54 186.8 229.7
Sc-2 85.55 92.16 99.42 107.32 116.02 125.56
Sc-3 85.55 53.11 58.74 64.97 71.95 79.70
Sc-4 85.55 89.79 94.52 99.74 105.62 112.18
EDNS MWh/yr 0% 20% 40% 60% 80% 100%
Sc-1 0 0.76 1.57 2.45 3.51 4.86
Sc-2 0 591.3 1182 1773 2365 2956
Sc-3 0 21.14 57.32 112.2 183.3 268.6
Sc-4 0 15.69 38.53 69.33 108.9 157.8
EDEDI hrs/yr 0% 20% 40% 60% 80% 100%
Sc-1 0 0.79 0.85 0.87 1.00 1.11
Sc-2 0 286.9 288.3 288.3 288.3 288.3
Sc-3 0 10.23 13.80 17.96 22.01 25.77
Sc-4 0 5.96 7.13 8.38 9.72 11.13
EENA hrs/yr 0% 20% 40% 60% 80% 100%
Sc-1 71.74 76.07 83.88 92.73 102.9 114.5
Sc-2 71.74 73.66 78.65 84.37 90.93 98.57
Sc-3 71.74 13022 16906 25071 37490 71280
Sc-4 71.74 76.19 83.92 92.68 102.5 113.7

From the EENS results in Table 6-7, Sc-1 achieved the highest load curtailment out of
the four scenarios, indicating that the worst network performance was obtained
when no EV-switch mode was considered. This is because EV charging demand was
considered to be the same load type as regular demand, thus load curtailment
occurred in both EV and regular demand, and there was no clear priority between

171
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

these two. As a result, Sc-1 tends to curtail regular demand in order to satisfy the
network constraints under contingency conditions, thus Sc-1 achieved the highest
EENS but the lowest EDNS out of the four scenarios.

As suggested from the EDNS and EDEDI results in Table 6-7, Sc-2 obtained
significantly more interruption in relation to the EV charging demand (highest among
scenarios), including the amount (EDNS) and the duration (EDEDI) of the interrupted
EV demand. This was because Sc-2 disconnected all aggregated EV charging lots
before restoration was established, thus the EV charging in the EV park lots was
interrupted during all of the contingencies, resulting in a huge EDNS and EDEDI.
However, the benefits of such an operating mode include that it is simple to operate,
there is no need for a smart device and that it is a conservative way to sacrifice the
interruption of EV charging demand for a lower load curtailment of regular demand
(smaller EENS), compared to Sc-1.

From the EENA results in Table 6-7, it is suggested that Sc-3 received significantly
more cable ageing, which was almost 70 times more than the other three scenarios.
Meanwhile, Sc-3 achieved the best reliability performance by obtaining the lowest
EENS of 79.7MWh. Sc-3 disconnected the aggregated EV charging lots when any
feeder circuit iwas overloaded. Compared to Sc-2, Sc-3 fully utilises the loadability of
the cable circuit to satisfy demands before any overload condition occurs, thus the
lowest load curtailment was achieved in Sc-3. However, such an operating scheme
requires remote switching and associated control devices, in order to operate the
correct switching action in response to the overload signals. Meanwhile, Sc-3 tends
to operate the cable circuits at their maximum allowable rating, thus much more
cable ageing is generated compared to the other scenarios. As a result, the EENS in
Sc-3 was reduced to the lowest, at the cost of significantly increased ageing. This
increased ageing may force utilities to perform more frequent condition checks and
maintenance, and replace heavily-loaded and most-aged cables. Thus, this increased
ageing cost and its related maintenance costs may lead Sc-3 to being considered less
cost-worthy.

172
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

Sc-4 proposes an optimised solution that switches off one or more EV parking lots
during contingencies in order to obtain a minimum operational cost of cable ageing,
load curtailment and power losses within the network loading and voltage
constraints. At the same time, due to the utilisation of remote switching and control
devices, Sc-4 was able to control the EV switches remotely in real time, thus choosing
to interrupt or supply the corresponding charging demand during contingencies. This
is a big technical improvement from Sc-2 and Sc-3, where the EV demand was
interrupted until the contingency finished.

From the results in Table 6-7, Sc-4 achieved more EENS than Sc-3, but less than Sc-1
and Sc-2. The second lowest EDNS and EDEDI was obtained by Sc-4, which was only
higher than Sc-1. Meanwhile, Sc-4 maintains the cable ageing (EENA) to an acceptable
level, similar to Sc-1 and Sc-2. This result shows that Sc-4 achieved a relatively low
load of curtailment and less frequent EV demand interruption at the cost of tolerable
ageing. If the utility applies a maximum EENS of 100MWh as their main criterion to
constraint the connection of more EVs, Sc-4 allows for a 60% penetration of EVs,
which is 40% more when compared to Sc-1. By setting the EV demand interruption
performance as a criterion, a projection of allowable EV penetration can be estimated
based on the EDNS or EDEDI. In this case, Sc-4 generates a smaller amount and lower
frequency of EV demand interruptions than Sc-2 and Sc-3, thus a higher EV
penetration can be enabled using the remote EV-switching mode.

Sc-4 applies a cost optimisation tool to achieve the minimum overall operating costs,
indicating its significant economic advantage against other operating modes.
However, the remote EV-switching mode requires that remote switches are equipped
for each EV park under central control and optimisation, thus the initial equipment
installation cost has not been considered in the total cost computation.

Emergency cable rating allows the feeder cables to operate at a higher current rating
in the restructured distribution network after contingencies. It provides more
network flexibility and loadability to supply the network demand and EV demand

173
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

under emergency operations. Figure 6-5 shows the amount of load curtailment
(EENS), the equivalent network cable ageing (EENA), the system average
interruption duration (SAIDI) and the frequency (SAIFI) of Sc-4 when normal (Inormal),
cyclic (Icyclic) and emergency (Ieme) ratings are utilised respectively.

140 Normal 1000


Cyclic Emergency
120 800
EENS (MWh/Yr)

EENA (hrs/Yr)
100
600
80
400
60
40 200

20 0
0% 20% 40% 60% 80% 100% 0% 20% 40% 60% 80% 100%
EV penetration PEV (%)
15 Normal 2
Cyclic Emergency
12 1.6
SAIDI (hrs/Yr)

SAIFI (occ./Yr)

9 1.2

6 0.8

3 0.4
0 0
0% 20% 40% 60% 80% 100% 0% 20% 40% 60% 80% 100%
EV penetration PEV (%)
Figure 6-5 EENS, EENA, SAIDI and SAIFI output of Sc-4 when utilising normal,
cyclic and emergency ratings

As shown in Figure 6-5, the significant reduction of EENS, SAIDI and SAIFI was
observed when increasing the cable rating from Inormal to Icyclic and Ieme, suggesting an
effective improvement of the overall network reliability performance. The reduction
seems to be very constant across all different EV penetration levels. For example, a
constant reduction of over 40MWh/yr in EENS was observed when the cable rating
increased from Inormal to Ieme. This improved network performance was because the
increased current rating provided more loadability and flexibility during
contingencies, thus more network demand, including regular and EV demand, can be
supplied under looser loading constraints. As a result, the negative impact of the EV

174
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

penetration on the network can be significantly mitigated by the implementation of


emergency uprating.

The benefits of emergency uprating can be further indicated by the amount of


interrupted EV demand (EDNS) and the expected duration of the EV demand
interruption (EDEDI) as in Figure 6-6. Although the EDNS increased exponentially
along with the EV penetration PEV% becoming higher, an effective reduction of EDNS
from 158MWh/yr to 55MWh/yr was achieved at PEV%=100% by uprating Inormal to
Ieme. The EDEDI output showed an almost linear rise when the PEV% increased, and
emergency uprating significantly decreased the increasing rate of EDEDI. In
particular, the EV interruption duration was reduced by 64% from 11hrs/yr to
4hrs/yr when Ieme was utilised.

200 Normal 12
Cyclic Emergency
10
EDNS (MWh/Yr)

150
EDEDI (hrs/Yr)

8
100 6
4
50
2
0 0
0% 20% 40% 60% 80% 100% 0% 20% 40% 60% 80% 100%
EV penetration PEV (%)
Figure 6-6 EDNS and EDEDI output of Sc-4 when utilising normal, cyclic and
emergency ratings

The increased load from EV charging will result in the distribution circuits being
loaded close to their limits. The network performance worsens, mainly due to the
loading constraints of the circuits. From the results, it was indicated that the
distribution cables do have the ability to accept additional EV charging demands
under emergency loading conditions. The ability stems from the UGCs’ inherent
characteristics (i.e. thermal inertia) that allow cycling or emergency currents for a
limited period of time.

The potential risks of ‘thermal overloading’, and thus, ageing can be demonstrated by
the EENS result in Figure 6-5. This shows that the emergency uprating leads to

175
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

increased cable ageing, particularly when the PEV% increases. As discussed before,
the current rating increase leads to an exponential growth of thermal ageing,
therefore the network ageing caused by emergency rating is much higher than that
of normal or cyclic rating. From Figure 6-5, the EENA is 766hrs/yr at PEV%=100%
when Ieme is used, while it is 228hrs/yr for Icyclic and 114hrs/yr for Inormal. The growth
rate of EENA along with the increase of PEV% is also significantly larger for the
emergency rating case.

The optimisation function of Sc-4 takes the cost of cable ageing into consideration,
thus it compares the additional ageing cost with the extra benefits from the improved
network reliability. Under the implementation of Ieme, it is more worthy to overload
the cables properly, considering that the increased ageing is still acceptable. However,
further uprating the cable to a higher current rating will lead to a more serious impact
on cable ageing. This increased ageing may require utilities to perform more frequent
maintenance and early replacement for their cable assets, thus emergency uprating
may become less worthy.

This chapter applied the proposed network reliability evaluation framework in order
to evaluate the impact of electric vehicles integration on a distribution network and
the related assets. Emergency ratings of cables were utilised in order to mitigate the
negative impact. The methods also enable a higher EV penetration in the existing
distribution networks which helps to postpone network upgrades and
reinforcements.

Emergency uprating provides utilities with a simple and direct method to uprate
their distribution cables to emergency rating levels under contingency operations.
This allows operators to unlock the emergency loadability of UGCs temporarily in
order to loosen the loading constraint and to transmit more power according to
demands. However, emergency uprating may lead to both the ‘current’ and ‘thermal’
overloading of cables, and therefore produce increased cable ageing. The potential
network benefits and cable ageing risks can be quantified through the proposed
methodology.

176
Chapter 6: Quantifying the Risk of Emergency Ratings on Distribution Cable Network with EVs

The case study on the modified RBTS bus 4 network has shown that the network
reliability performance significantly worsens when a high penetration of EVs is
considered. When EV penetration PEV%=100%, the load curtailment was increased
by 160% from 85.5 to 229.7MWh/yr. The remote EV-switching method applied cost
optimisation to achieve a balance between load curtailment, cable ageing and EV
demand interruption. Thus, the remote switching method reduces the load
curtailment by half while maintaining cable ageing and EV charging interruption at
an acceptable level. Emergency uprating also provides an effective way to improve
network performance at the cost of increased cable ageing. Compared to normal
rating, emergency rating reduces by half the amount of load curtailment and
interrupted EV charging demand. The emergency uprating method could be too risky
when a higher current rating is considered.

The proposed methodology allows utilities to determine the optimum emergency


rating and operating strategy for their distribution networks with a high penetration
of EV connections. Further work is required to design a more advanced network
reconfiguration tool in order to adapt the increasing penetration of EVs.

Equation Chapter (Next) Section 1

177
Chapter 7: Conclusion & Future Work

7 Conclusion & Future Work

The proposed methodology framework integrating cable design and ageing is


applied for several studies. The quantified results obtained from each study
provide utilities with recommendations on updating their existing cable rating
policies, adopt effective and targeted strategies for asset management, as well as
implementing appropriate smart grid technologies.

Nevertheless there are a number of areas where the methodological framework


can be further applied for further studies within the field of reliability-flexibility
evaluation of cable-based networks. Also the methodology can be further extended
and improved to include additional network and plant modelling, and produce
more universal and useful results. This chapter discusses the conclusion of the
research in this thesis, and provides the suggestions for future work.

This thesis proposed a novel methodological advancement in the existing power


network reliability evaluation methodology by integrating a detailed thermal-ageing
modelling of UGCs into the network assessments. The thermal ageing of the cables
requires thermal stress and measurement of electrical insulation performance based
on cable testing. This has been investigated by other researchers while this thesis
focuses on thermal ageing estimation through simulation under different thermal
rating approaches. The thermal-ageing modelling included design properties of the
cable system (cable, laying formation) and environmental properties (soil temp,
resistivity) in order to capture the ageing of all the cables (i.e. map the ageing) within
the network. This inclusion generates complications on ageing and plant
methodological steps, as well as the formulation of novel indices that might seem
simple yet difficult to realise, which is evidently neglected from current literature
methods.

178
Chapter 7: Conclusion & Future Work

This research was motivated by the rising needs from the power utilities to operate
their power networks with increasing flexibility, utilisation and efficiency.
Considering a large number of cable assets existed in both transmission and
distribution networks, flexible cable rating approaches and smart grid technologies
provide suitable solutions to realise the goal. Thus, it is important to exploit the
benefits and potential risks of those solutions. The methodological advancement
proposed in this thesis addresses a key question on how to quantify the benefits and
risks related to the current increasing utilisation and flexibility challenges that power
utilities face.

A thorough review of past research related to the existing reliability evaluation


approaches, cable thermal and ageing modelling has been presented in Chapter 1.
Despite many methods existed to evaluate the ageing and thermal conditions of UGCs,
almost no research has combined multiple cable models (life/ageing model,
probabilistic failure model) with the time-sequential Monte Carlo simulation in order
to compute the consequences in terms of cable ageing, failure probability, failure rate
and network reliability. In this way, the behaviour of power cables and the network
can be evaluated as a whole, rather than separately as before. According to the
assessment of the reliability and cable ageing condition of a network, the total cost
and cable lifetime can be estimated in order to determine an optimised actions and
procedures of cable replacement and maintenance. The main contribution of this
thesis is the integration of different existing methods (cable thermal-electric
equivalent model, Arrhenius ageing model and high-loading failure model) within the
time-series sequential Monte Carlo simulation, which allows a calculation of risk due
to ageing. This integration has been successfully achieved as it can be seen in Chapter
2 and from the following results section.

The second main contribution of this thesis is to implement three additional high-
loading components’ operational Markov states on cable modelling that allow for
capturing the increased risk of failures when worn-out components are operating in
emergency states. This risk inclusion has not been considered before in any previous
work. A life-cycle simulation loop on network modelling was also included with most
of the equations were formulated to capture the sequential loop of the life-cycle of

179
Chapter 7: Conclusion & Future Work

the cables. No equation in previous studies was reported to capture this ageing
challenge. The combination of these high-loading component states and life-cycle
loop allows capturing the benefits from increased network flexibility when utilizing
emergency ratings but also increased ageing risk and failure risk of cables with
extensive emergency loading history.

The cable TEE modelling is formulated into a generic structure that allows its
application on multiple types of cables (i.e. oil, XLPE, paper) and soils. Also, the TEE
equations allow for the time-dependency of the TEE, and hence they can be used for
both steady-state and transient-state of thermal modelling as implemented in the
case study of Chapter 3. The newly proposed equivalent cable ageing (ECA) index is
critical for mapping network component ageing as ECA allows the methodology to
highlight critical cables based on the equivalent ageing base among the other network
cables. Utilities can compare their recorded failure rates and health indices recorded
by their asset management practice with the ageing outputs of this method to cluster
cables into similar health groups, such as heavily loaded cables with accelerated
ageing, low loaded cables with zero ageing. The 14-bus example in Chapter 3 is meant
to highlight this practical implementation of the presented methodological
advancement, helping the utilities to design an action strategy beforehand and
implement their standby actions under emergency conditions. By recording these
actions to respond to critical events, utilities can further formulate future strategies
with an improved performance which will allow replacing, maintaining an asset
considering future expected demand increase.

This thesis further investigated the possibility of extending the operating duration
and ampacities of emergency ratings in transmission cables and lines. The proposed
methodology is based on the methodological framework but also combining with a
cascading failure model, and an emergency rating model that considers emergency
operating procedures. Two comparative case studies were performed in 14-bus and
24-bus network with all OHLs or UGCs equipped. The results clearly indicate that
UGCs have a greater potential (origins from the large thermal inertia of the cable itself
and surrounding soil) to extend their emergency operating durations and ampacities
than OHLs, with an acceptable amount of increased ageing and cascading failure.

180
Chapter 7: Conclusion & Future Work

However, the increasing number of cascading failure may offset, or even totally
overweight the benefits from increased flexibility when utilizing extensive
emergency ratings. The comparison between 14-bus and 24-bus study clearly
indicates the results highly depends on the network load level, network design and
structure. Thus, utilities are suggested to implement the proposed methodology and
related cost-worth analysis in order to design more targeted emergency rating
operation strategies for each specific network. This developed methodology to
quantify the technical/economic benefits and costs of implementing extensive
emergency ratings is the third contribution of this thesis.

The practical requests from the industry (windfarm developers & utilities) drive us
to propose a novel forecasted flexible rating method for their windfarm cable circuits.
Thus, network operators can achieve a similar operational and control scheme for
windfarm cable circuits without the implementation of real-time temperature
monitoring. The proposed method has better business values as it reduces
investment costs and enables implementations for most existing wind farms. From
case studies, over 20% additional headroom capacity can be unlocked, by changing
the circuit’s current constraint to thermal constraint. At least 20% additional wind
power can be exported using flexible rating of existing circuits when only few wind
generation curtailments occur. Our simulation-based analysis demonstrate the
benefits of different rating approaches deployed in windfarm cable circuits, which
boost the confidence for grid code updates and full business adoption in the future.
SPEN has provided a recommendation on changing the existing asset rating policy to
consider flexible rating approaches for sizing the windfarm cable circuits. This novel
developed flexible rating approach for increasing wind integration is the fourth
contribution of this thesis.

This thesis further investigated the impact of electric vehicles penetration in


distribution network, and potential solutions of utilising emergency ratings to
increase the maximum allowable penetration of EVs into the existing network. An
optimisation model was developed to minimise the cost of cable ageing, power losses,
load curtailment and EV interruption under the network loading and voltage
constraints. The results suggest that the increasing penetration of EVs increases the

181
Chapter 7: Conclusion & Future Work

network loading and thus cables’ ageing, leading to a reduced overall network
performances. However, the negative impact of EVs can be greatly mitigated by the
implementation of emergency ratings of UGCs and EV switches. To this end, the
methodology provide DSOs with a tool to assess the impact of EV integration on the
network and distribution cables, as well as two easy-to-implement solutions for
increasing the maximum allowable EV penetration in their existing network. The fifth
contribution of this thesis is to quantify the benefits and risks of emergency rating
implementation of UGCs in a distribution cable networks with EVs.

In addition to the fulfilled research tasks within this thesis, there are a number of
areas that could be further extended and improved as future work. The future work
can be conduct from either the cable modelling or the network modelling.

In the aspect of the cable modelling, the develop methodology so far has addressed
issues related with thermal rating and ageing under normal laying conditions.
However, the modelling of UGCs considering the cable joints as well as thermally
unfavourable conditions (e.g., mutual heating of cables, road crossing, and extreme
soil conditions) could provide more realistic solutions towards improving flexibility
and resilience of cable networks through the implementation of increased thermal
rating of cables. The failure modelling of cable joints and cables under unfavourable
condition should also be included. One aim of future work is to quantify the risk of
thermal ageing of cables under thermally unfavourable conditions, as well as to
provide tools to the utilities to reduce the ageing risk on those ‘hotspots’ and extend
the life of their UGCs networks without compromising network performance.

In the proposed methodology capturing the risk of high-loading failures (Chapter 3),
the α and β coefficients were proposed to determine the transitions of the new high-
loading states. No accurate coefficients had been established and implemented in the
case studies. The α and β coefficients have to be designed by the utility based on field
failure data and operating practices as well as correlate those with cable type and
manufacturer data. The correlation of the recorded data with the simulated data
could help initially identify the impact of failures on the other cables and then

182
Chapter 7: Conclusion & Future Work

formulate the α and β. Thus, future work is required to formulate further these
coefficients in collaboration with utilities, cable manufacturer and cable test labs.

This thesis proposed the methodology to calculate an equivalent cable ageing (ECA)
index purely based on simulation. Therefore there is a necessity to develop methods
for combining several existing health indices (from on-site measurements, existing
Ofgem asset management methodology, and simulation results) into an overall cable
ageing index. Both thermal and voltage stress on cables can be considered within the
modelling of cable ageing and probability of failure. These indices can be converted
into a realistic and practical probabilities of failures of each cable, and then to be
implemented into a network wide performance evaluation in order to quantify the
risk of failure and compare it to the worth of maintenance.

In the aspect of the network modelling, the proposed methodology should be applied
and testified on real-life large transmission and distribution networks with a mixed
cable type (i.e. cable design, type, laying conditions) in order to achieve a more
realistic assessment and results. More realistic variables in real-life networks, such
as renewables and load cycling, should also be included in the methodology.

The proposed methodology can also be extended to assess the impact of other smart
grid technologies for increasing network flexibility, such as renewable sources,
electric vehicles, dynamic ratings, demand response, storage and FACTs. The risk-
flexibility evaluation can be conducted to quantify their benefits from increased
network flexibility but also increased risk of asset ageing and failure, as well as other
potential network risks, such as cascading failure.

This integration between cables and networks could also be applied on other
network components, such as transformers, overhead lines, and protection system.
Relevant dynamic/emergency rating methods and thermal-ageing modelling could
also be integrated into SMC simulation, to assess their impact on the network
reliability.

Another future work is to propose a novel network reconfiguration and optimisation


methodology in order to balance the network loading on each cable circuit based on

183
Chapter 7: Conclusion & Future Work

cable thermal profiles. One of the main outputs is to identify optimal operating rating
(OR) level for better network utilization considering the thermal and design
properties of each network cable. Another aim is to improve distribution network
reliability performance, as well as to balance thermal conditions of cables and extend
cable life. Case studies could be performed on standard (radial and meshed)
transmission and distribution networks.

This PhD project has shown a way to integrate the cable thermal and ageing
calculation in the planning and operation of power systems, and the author suggests
that this topic in general is worth to be further investigated in the future. It would be
very beneficial for network operators to take power equipment’s electrical and
thermal properties into their consideration. Thus, the integration of plant design and
ageing properties into network assessment will be of significant interest to both
power system researchers and power utilities.

184
Chapter 8: References

8 References
[1] R. Billinton and R. N. Allan, "Power-system reliability in perspective," Electronics
and Power, vol. 30, no. 3, pp. 231-236, 1984.
[2] D. S. Kirschen and D. Jayaweera, "Comparison of risk-based and deterministic
security assessments," IET Generation, Transmission & Distribution, vol. 1, no. 4,
pp. 527-533, 2007.
[3] R. Billinton and W. Li, Reliability Assessment of Electrical Power Systems Using
Monte Carlo Methods. London: Springer, 1994.
[4] "EPRI: Electric Power System Flexibility - Challenges and Opportunities," Electric
Power Research Institute, 2016.
[5] "Position Paper: Making the electricity system more flexible and delivering the
benefits for consumers," Ofgem, 2015.
[6] S. Bahadoorsingh and S. Rowland, "A Framework Linking Knowledge of Insulation
Aging to Asset Management - [Feature Article]," IEEE Electrical Insulation
Magazine, vol. 24, no. 3, pp. 38-46, 2008.
[7] J. Schneider, A. J. Gaul, C. Neumann, J. Hogräfer, W. Wellßow, M. Schwan, et al.,
"Asset management techniques," International Journal of Electrical Power &
Energy Systems, vol. 28, no. 9, pp. 643-654, 2006/11/01/ 2006.
[8] "DNO Common Network Asset indices Methodology," 2017.
[9] J. J. Shea, "Aging power delivery infrastructures [Book Reviews]," IEEE Electrical
Insulation Magazine, vol. 17, no. 6, pp. 63-63, 2001.
[10] L. Xu and R. E. Brown, "Justifying the proactive replacement of cable," in 2011 IEEE
Power and Energy Society General Meeting, 2011, pp. 1-6.
[11] M. Buhari, V. Levi, and S. K. E. Awadallah, "Modelling of Ageing Distribution Cable
for Replacement Planning," IEEE Transactions on Power Systems, vol. 31, no. 5, pp.
3996-4004, 2016.
[12] S. K. E. Awadallah, J. V. Milanović, and P. N. Jarman, "Reliability Based Framework
for Cost-Effective Replacement of Power Transmission Equipment," IEEE
Transactions on Power Systems, vol. 29, no. 5, pp. 2549-2557, 2014.
[13] A. A. Chowdhury, T. C. Mielnik, L. E. Lawton, M. J. Sullivan, A. Katz, and D. O. Koval,
"System Reliability Worth Assessment Using the Customer Survey Approach," IEEE
Transactions on Industry Applications, vol. 45, no. 1, pp. 317-322, 2009.
[14] P. Pollak, "Neher-McGrath Calculations for Insulated Power Cables," IEEE
Transactions on Industry Applications, vol. IA-21, no. 5, pp. 1319-1323, 1985.
[15] IEC Standard 60287-1-1: Current rating equations (100 % load factor) and
calculation of losses, International Electrotecnical Commission Standard 2014.

185
Chapter 8: References

[16] IEC Standard 60287-2-1: Calculation of thermal resistance, International


Electrotecnical Commission Standard 2015.
[17] ENA Engineering Recommendation ER P17: ‘Current Rating Guide for Distribution
Cables’, Standard 2004.
[18] "ESDD-02-007: Equipment Ratings_SPEN," SP Power Systems Limited, 2013.
[19] "Code of Practice 203: Current Ratings - Underground Cables," Electricity North
West, 2009.
[20] H. C. Zhao, J. S. Lyall, and G. Nourbakhsh, "Probabilistic cable rating based on cable
thermal environment studying," in PowerCon 2000. 2000 International
Conference on Power System Technology. Proceedings (Cat. No.00EX409), 2000,
pp. 1071-1076 vol.2.
[21] G. Mazzanti, "Analysis of the Combined Effects of Load Cycling, Thermal
Transients, and Electrothermal Stress on Life Expectancy of High-Voltage AC
Cables," IEEE Transactions on Power Delivery, vol. 22, no. 4, pp. 2000-2009, 2007.
[22] ISO New England Capacity Rating Procedures [Online]. Available:
https://www.iso-ne.com/static-
assets/documents/rules_proceds/isone_plan/pp07/capacity_rating_procedures.
pdf
[23] PJM Manual 03: Transmission Operations [Online]. Available:
http://www.pjm.com/~/media/documents/manuals/m03.ashx
[24] "Lessons Learned Report - Real Time Thermal Rating," Customer-Led Network
Revolution (CLNR) Project, Northern Powergrid, 2014.
[25] SPEnergyNetworks, "Close Down Report: Temperature Monitoring Windfarm
Cable Circuits," 2015.
[26] "NIA_NGET0047: Dynamic Ratings for improved Operational Performance (DROP)
" National Grid, 2017.
[27] Y. Jin, B. Xuefeng, D. Strickland, L. Jenkins, and A. M. Cross, "Dynamic Network
Rating for Low Carbon Distribution Network Operation: A U.K. Application," Smart
Grid, IEEE Transactions on, vol. 6, no. 2, pp. 988-998, 2015.
[28] Y. C. Liang and Y. M. Li, "On-line dynamic cable rating for underground cables
based on DTS and FEM," WSEAS Trans. Cir. and Sys., vol. 7, no. 4, pp. 229-238,
2008.
[29] M. Z. Degefa, M. Humayun, A. Safdarian, M. Koivisto, R. J. Millar, and M. Lehtonen,
"Unlocking distribution network capacity through real-time thermal rating for
high penetration of DGs," Electric Power Systems Research, vol. 117, pp. 36-46,
2014/12/01/ 2014.
[30] A. Safdarian, M. Z. Degefa, M. Fotuhi-Firuzabad, and M. Lehtonen, "Benefits of
Real-Time Monitoring to Distribution Systems: Dynamic Thermal Rating," IEEE
Transactions on Smart Grid, vol. 6, no. 4, pp. 2023-2031, 2015.

186
Chapter 8: References

[31] G. J. Anders, A. Napieralski, M. Zubert, and M. Orlikowski, "Advanced modeling


techniques for dynamic feeder rating systems," IEEE Transactions on Industry
Applications, vol. 39, no. 3, pp. 619-626, 2003.
[32] B. Clairmont, "Real-time monitors and dynamic thermal rating methods," in EPRI
Increased Power Flow Guidebook, ed: Electric Power Research Institute, 2010.
[33] S. P. Walldorf, J. S. Engelhardt, and F. J. Hoppe, "The use of real-time monitoring
and dynamic ratings for power delivery systems and the implications for dielectric
materials," IEEE Electrical Insulation Magazine, vol. 15, no. 5, pp. 28-33, 1999.
[34] R. Huang, "Dynamic Rating for Improved Operational Performance," PhD,
Department of Electronics and Computer Science, University of Southampton,
2015.
[35] M. M. Jensen, "Overload Capacity of Polymer Insulated Medium Voltage Cables,"
Master, Department of Electrical Engineering, Technical Institution of Denmark
(DTU), Lyngby, Denmark, 2011.
[36] W. Z. Black and P. Sang-il, "Emergency Ampacities of Direct Buried Three Phase
Underground Cable Systems," IEEE Transactions on Power Apparatus and Systems,
vol. PAS-102, no. 7, pp. 2124-2132, 1983.
[37] R. S. Olsen, J. Holboll, and U. S. Gudmundsd, "Dynamic temperature estimation
and real time emergency rating of transmission cables," in 2012 IEEE Power and
Energy Society General Meeting, 2012, pp. 1-8.
[38] R. Olsen, J. Holboell, and U. S. Gudmundsd, "Electrothermal Coordination in Cable
Based Transmission Grids," IEEE Transactions on Power Systems, vol. 28, no. 4, pp.
4867-4874, 2013.
[39] A. S. Dobakhshari and M. Fotuhi-Firuzabad, "A Reliability Model of Large Wind
Farms for Power System Adequacy Studies," IEEE Transactions on Energy
Conversion, vol. 24, no. 3, pp. 792-801, 2009.
[40] Y. Ding, C. Singh, L. Goel, J. Østergaard, and P. Wang, "Short-Term and Medium-
Term Reliability Evaluation for Power Systems With High Penetration of Wind
Power," IEEE Transactions on Sustainable Energy, vol. 5, no. 3, pp. 896-906, 2014.
[41] M. Al-Muhaini and G. T. Heydt, "Evaluating Future Power Distribution System
Reliability Including Distributed Generation," IEEE Transactions on Power Delivery,
vol. 28, no. 4, pp. 2264-2272, 2013.
[42] L. Cheng, Y. Chang, J. Lin, and C. Singh, "Power System Reliability Assessment With
Electric Vehicle Integration Using Battery Exchange Mode," IEEE Transactions on
Sustainable Energy, vol. 4, no. 4, pp. 1034-1042, 2013.
[43] K. Hou, X. Xu, H. Jia, X. Yu, T. Jiang, K. Zhang, et al., "A Reliability Assessment
Approach for Integrated Transportation and Electrical Power Systems
Incorporating Electric Vehicles," IEEE Transactions on Smart Grid, vol. 9, no. 1, pp.
88-100, 2018.

187
Chapter 8: References

[44] S. H. Huang, W. J. Lee, and M. T. Kuo, "An Online Dynamic Cable Rating System for
an Industrial Power Plant in the Restructured Electric Market," IEEE Transactions
on Industry Applications, vol. 43, no. 6, pp. 1449-1458, 2007.
[45] J. Teh and I. Cotton, "Reliability Impact of Dynamic Thermal Rating System in Wind
Power Integrated Network," IEEE Transactions on Reliability, vol. 65, no. 2, pp.
1081-1089, 2016.
[46] H. Kim and C. Singh, "Power system reliability modeling with aging using thinning
algorithm," in 2009 IEEE Bucharest PowerTech, 2009, pp. 1-6.
[47] L. Wenyuan, "Incorporating aging failures in power system reliability evaluation,"
IEEE Transactions on Power Systems, vol. 17, no. 3, pp. 918-923, 2002.
[48] D. J. Swaffield, P. L. Lewin, and S. J. Sutton, "Methods for rating directly buried
high voltage cable circuits," IET Generation, Transmission & Distribution, vol. 2, no.
3, pp. 393-401, 2008.
[49] M. Diaz-Aguiló and F. d. León, "Introducing Mutual Heating Effects in the Ladder-
Type Soil Model for the Dynamic Thermal Rating of Underground Cables," IEEE
Transactions on Power Delivery, vol. 30, no. 4, pp. 1958-1964, 2015.
[50] V. Levi, M. Buhari, and A. Kapetanaki, "Cable Replacement Considering Optimal
Wind Integration and Network Reconfiguration," IEEE Transactions on Smart Grid,
pp. 1-1, 2017.
[51] "IEEE Recommended Practice for Protection and Coordination of Industrial and
Commercial Power Systems (IEEE Buff Book)," IEEE Std 242-2001 (Revision of IEEE
Std 242-1986) [IEEE Buff Book], pp. 1-710, 2001.
[52] M. Kellow and H. St-Onge, "Thermo-Mechanical Failure of Distribution Cables
Subjected to Emergency Loading," IEEE Power Engineering Review, vol. PER-2, no.
7, pp. 29-30, 1982.
[53] X. Zhang, E. Gockenbach, V. Wasserberg, and H. Borsi, "Estimation of the Lifetime
of the Electrical Components in Distribution Networks," IEEE Transactions on
Power Delivery, vol. 22, no. 1, pp. 515-522, 2007.
[54] G. Mazzanti, "The combination of electro-thermal stress, load cycling and thermal
transients and its effects on the life of high voltage ac cables," Dielectrics and
Electrical Insulation, IEEE Transactions on, vol. 16, no. 4, pp. 1168-1179, 2009.
[55] Z. Xiang and E. Gockenbach, "Assessment of the actual condition of the electrical
components in medium-voltage networks," IEEE Transactions on Reliability, vol.
55, no. 2, pp. 361-368, 2006.
[56] M. Stotzel, M. Zdrallek, and W. H. Wellssow, "Reliability calculation of MV-
distribution networks with regard to ageing in XLPE-insulated cables," IEE
Proceedings - Generation, Transmission and Distribution, vol. 148, no. 6, pp. 597-
602, 2001.
[57] "Increased Power Flow Guide Book: Increasing Power Flow on Transmissionand
Substation Circuits," EPRI, Palo Alto, CA, USA, 2005.

188
Chapter 8: References

[58] K. Kopsidas, C. Tumelo-Chakonta, and C. Cruzat, "Power Network Reliability


Evaluation Framework Considering OHL Electro-Thermal Design," IEEE
Transactions on Power Systems, vol. 31, no. 3, pp. 2463-2471, 2016.
[59] C. Cruzat, K. Kopsidas, and S. Liu, "Evaluating the impact of information and
communication technologies on network reliability," in 2016 18th Mediterranean
Electrotechnical Conference (MELECON), 2016, pp. 1-7.
[60] R. Billinton and R. N. Allan, Reliability evaluation of power systems, 2nd ed. NEW
YORK AND LONDON: PLENUM PRESS, 1996.
[61] S. Liu and K. Kopsidas, "Reliability evaluation of distribution networks
incorporating cable electro-thermal properties," in 2016 Power Systems
Computation Conference (PSCC), 2016, pp. 1-7.
[62] K. Kopsidas and S. Liu, "Power Network Reliability Framework for Integrating
Cable Design and Ageing," IEEE Transactions on Power Systems, vol. PP, no. 99, pp.
1-1, 2017.
[63] R. Olsen, G. J. Anders, J. Holboell, and U. S. Gudmundsdottir, "Modelling of
Dynamic Transmission Cable Temperature Considering Soil-Specific Heat, Thermal
Resistivity, and Precipitation," IEEE Transactions on Power Delivery, vol. 28, no. 3,
pp. 1909-1917, 2013.
[64] D. A. Zarchi and B. Vahidi, "Optimal placement of underground cables to maximise
total ampacity considering cable lifetime," IET Generation, Transmission &
Distribution, vol. 10, no. 1, pp. 263-269, 2016.
[65] M. Diaz-Aguil, F. D. Le, S. Jazebi, and M. Terracciano, "Ladder-Type Soil Model for
Dynamic Thermal Rating of Underground Power Cables," IEEE Power and Energy
Technology Systems Journal, vol. 1, pp. 21-30, 2014.
[66] IEC Standard 60853-2: Calculation of the cyclic and emergency current rating of
cables - Part 2: Cyclic rating of cables greater than 18/30 (36) kV and emergency
ratings for cables of all voltages, International Electrotecnical Commission
Standard 1989.
[67] K. Kopsidas, "Impact of thermal uprating and emergency loading of OHL networks
on interconnection flexibility," in 2016 18th Mediterranean Electrotechnical
Conference (MELECON), 2016, pp. 1-6.
[68] "IEEE Guide for Determining the Effects of High-Temperature Operation on
Conductors, Connectors, and Accessories," IEEE Std 1283-2013 (Revision of IEEE
Std 1283-2004), pp. 1-47, 2013.
[69] K. Kopsidas, S. M. Rowland, and B. Boumecid, "A Holistic Method for Conductor
Ampacity and Sag Computation on an OHL Structure," IEEE Transactions on Power
Delivery, vol. 27, no. 3, pp. 1047-1054, 2012.
[70] C. F. Price and R. R. Gibbon, "Statistical approach to thermal rating of overhead
lines for power transmission and distribution," IEE Proceedings C - Generation,
Transmission and Distribution, vol. 130, no. 5, pp. 245-256, 1983.

189
Chapter 8: References

[71] ENA Engineering Recommendation ER P27: Current Rating Guide for High Voltage
Overhead Lines Operating in the UK Distribution System, Standard 2007.
[72] "Guide for Application of Direct Real-Time Monitoring Systems - Brochure No.
498," Cigre - International Council on Large Electric Systems - Working Group
B2.36, 2012.
[73] K. Kopsidas, A. Kapetanaki, V. Levi, and J. Milanovic, "Optimal Demand Response
Scheduling with Real Time Thermal Ratings of Overhead Lines for Improved
Network Reliability," IEEE Transactions on Smart Grid, vol. PP, no. 99, pp. 1-1, 2017.
[74] C. Grigg, P. Wong, P. Albrecht, R. Allan, M. Bhavaraju, R. Billinton, et al., "The IEEE
Reliability Test System-1996. A report prepared by the Reliability Test System Task
Force of the Application of Probability Methods Subcommittee," IEEE
Transactions on Power Systems, vol. 14, no. 3, pp. 1010-1020, 1999.
[75] ISO New York Emergency Operations Manual [Online]. Available:
http://www.nyiso.com/public/webdocs/markets_operations/documents/Manu
als_and_Guides/Manuals/Operations/em_op_mnl.pdf
[76] ISO New England Operating Procedure No. 19 - Transmission Operations [Online].
Available: http://www.iso-ne.com/static-
assets/documents/rules_proceds/operating/isone/op19/op19_rto_final.pdf
[77] M. Buhari, K. Kopsidas, C. Tumelo-Chakonta, and A. Kapetanaki, "Risk assessment
of smart energy transfer in distribution networks," in IEEE PES Innovative Smart
Grid Technologies, Europe, 2014, pp. 1-6.
[78] M. S. R. B. M. G. d. Santos, "Bibliography on Power Systems Probabilistic Security
Analysis 1968-2008," International Journal of Emerging Electric Power Systems,
vol. 10, no. 3, pp. 1-48, 2009.
[79] Medium Voltage Cables Brochure - Oman Cables [Online]. Available:
http://www.omancables.com/brochure/Medium_Voltage_Cables.pdf
[80] High and Extra High Voltage Cable System - Demirer Kablo [Online]. Available:
http://www.demirerkablo.com/media/33103/catalogue.pdf
[81] P. M. Subcommittee, "IEEE Reliability Test System," IEEE Transactions on Power
Apparatus and Systems, vol. PAS-98, no. 6, pp. 2047-2054, 1979.
[82] L. L. Grigsby, Power Systems, Third Edition: CRC Press, 2016.
[83] ECMWF. (2014). ERA Interim dataset: soil temperature. Available:
http://apps.ecmwf.int/datasets/data/interim-full-daily
[84] "Emergency Control Practices," IEEE Transactions on Power Apparatus and
Systems, vol. PAS-104, no. 9, pp. 2336-2341, 1985.
[85] NERC Standard FAC-008-1-Facility Rating Methodology, Standard 2006.
[86] "IEEE Standard for Calculating the Current-Temperature Relationship of Bare
Overhead Conductors," IEEE Std 738-2012 (Revision of IEEE Std 738-2006 -
Incorporates IEEE Std 738-2012 Cor 1-2013), pp. 1-72, 2013.

190
Chapter 8: References

[87] ISO New England Planning Procedure No. 7, Procedures for Determining and
Implementing Transmission Facility Ratings in New England [Online]. Available:
http://www.iso-ne.com/static-
assets/documents/rules_proceds/isone_plan/pp07/pp7_final.pdf
[88] S. Liu, C. Cruzat, and K. Kopsidas, "Impact of transmission line overloads on
network reliability and conductor ageing," in 2017 IEEE Manchester PowerTech,
2017, pp. 1-6.
[89] M. Abogaleela, K. Kopsidas, C. Cruzat, and S. Liu, "Reliability evaluation framework
considering OHL emergency loading and demand response," in 2017 IEEE PES
Innovative Smart Grid Technologies Conference Europe (ISGT-Europe), 2017, pp.
1-6.
[90] I. Jones, "EN020014-001084: Analysis of Costings of Overhead Line v Underground
Cable," 2014.
[91] University of Manchester: The Whitworth Meteorological Observatory [Online].
Available: http://www.cas.manchester.ac.uk/restools/whitworth
[92] "Electricity Ten Year Statement (ETYS)," National Grid, 2016.
[93] S. Catmull, R. D. Chippendale, J. A. Pilgrim, G. Hutton, and P. Cangy, "Cyclic Load
Profiles for Offshore Wind Farm Cable Rating," IEEE Transactions on Power
Delivery, vol. 31, no. 3, pp. 1242-1250, 2016.
[94] R. Chippendale, J. Pilgrim, A. Kazerooni, and D. Ruthven. (2017). Cyclic rating of
wind farm cable connections. CIRED - Open Access Proceedings Journal 2017(1),
91-95. Available: http://digital-library.theiet.org/content/journals/10.1049/oap-
cired.2017.0922
[95] P. Chen and T. Thiringer, "Time-Series Based Cable Selection for a Medium Voltage
Wind Energy Network," IEEE Transactions on Sustainable Energy, vol. 3, no. 3, pp.
465-473, 2012.
[96] SPEnergyNetworks, "NIA Distribution Annual Report 2015/2016," 2016.
[97] L. F. Ochoa, C. J. Dent, and G. P. Harrison, "Distribution Network Capacity
Assessment: Variable DG and Active Networks," IEEE Transactions on Power
Systems, vol. 25, no. 1, pp. 87-95, 2010.
[98] (2017). NIA_SPEN0003: Enhanced Real-Time Cable Temperature Monitoring.
Available: http://www.smarternetworks.org/Project.aspx?ProjectID=1684
[99] A. Phillips, "Evaluation of Instrumentation and Dynamic Thermal Ratings for
Overhead Lines," EPRI, 2013.
[100] P. M. T. Broersen, "Automatic spectral analysis with time series models," IEEE
Transactions on Instrumentation and Measurement, vol. 51, no. 2, pp. 211-216,
2002.
[101] J. L. Torres, A. García, M. De Blas, and A. De Francisco, "Forecast of hourly average
wind speed with ARMA models in Navarre (Spain)," Solar Energy, vol. 79, no. 1,
pp. 65-77, 2005/07/01/ 2005.

191
Chapter 8: References

[102] R. Karki, H. Po, and R. Billinton, "A simplified wind power generation model for
reliability evaluation," IEEE Transactions on Energy Conversion, vol. 21, no. 2, pp.
533-540, 2006.
[103] "Feed-in Tariff: Generation & Export Payment Rate," Ofgem, Ed., 7th April 2017
ed, 2017.
[104] Met Office MIDAS: UK Hourly Weather Observation Data [Online]. Available:
http://data.ceda.ac.uk/badc/ukmo-midas/data/WH/
[105] "Review of WPD Unit Costs," Parsons Brinckerhoff, 2013.
[106] "WPD Lifetime Costs Report: Brechfa Forest Connection Project," Western Power
Distribution, 2015.
[107] A. Coelho, C. L. C. de Castro, M. G. da Silva, and A. B. Rodrigues, "Inclusion of
voltage drop and feeder loading constraints in the evaluation of reliability indices
for radial distribution networks," IEE Proceedings-Generation, Transmission and
Distribution, vol. 153, no. 6, pp. 661-669, 2006.
[108] "Impact of Electric Vehicle and Heat Pump loads on network demand profiles," UK
Power Networks, 2014.
[109] R. C. Green, L. Wang, and M. Alam, "The impact of plug-in hybrid electric vehicles
on distribution networks: A review and outlook," Renewable and Sustainable
Energy Reviews, vol. 15, no. 1, pp. 544-553, 2011/01/01/ 2011.
[110] "Closedown Report: Distribution Asset Thermal Modelling 'ENWL002'," Electricity
North West, 2016.
[111] S. W. Hadley and A. A. Tsvetkova, "Potential Impacts of Plug-in Hybrid Electric
Vehicles on Regional Power Generation," The Electricity Journal, vol. 22, no. 10,
pp. 56-68, 2009/12/01/ 2009.
[112] R. C. Green, L. Wang, and M. Alam, "The impact of plug-in hybrid electric vehicles
on distribution networks: a review and outlook," in IEEE PES General Meeting,
2010, pp. 1-8.
[113] L. P. Fernandez, T. G. S. Roman, R. Cossent, C. M. Domingo, and P. Frias,
"Assessment of the Impact of Plug-in Electric Vehicles on Distribution Networks,"
IEEE Transactions on Power Systems, vol. 26, no. 1, pp. 206-213, 2011.
[114] V. Aravinthan and W. Jewell, "Controlled Electric Vehicle Charging for Mitigating
Impacts on Distribution Assets," IEEE Transactions on Smart Grid, vol. 6, no. 2, pp.
999-1009, 2015.
[115] R. Mehta, D. Srinivasan, A. M. Khambadkone, J. Yang, and A. Trivedi, "Smart
Charging Strategies for Optimal Integration of Plug-In Electric Vehicles Within
Existing Distribution System Infrastructure," IEEE Transactions on Smart Grid, vol.
9, no. 1, pp. 299-312, 2018.
[116] R. A. Verzijlbergh, M. O. W. Grond, Z. Lukszo, J. G. Slootweg, and M. D. Ilic,
"Network Impacts and Cost Savings of Controlled EV Charging," IEEE Transactions
on Smart Grid, vol. 3, no. 3, pp. 1203-1212, 2012.

192
Chapter 8: References

[117] W. S. P. Denholm, "An Evaluation of Utility System Impacts and Benefits of


Optimally Dispatched Plug-In Hybrid Electric Vehicles," National Renewable
Energy Laboratory, 2006.
[118] "Reliability Improments from the Application of Distribution Automation
Technologies - Initial Results," U.S. Department of Energy, 2012.
[119] R. N. Allan, R. Billinton, I. Sjarief, L. Goel, and K. S. So, "A reliability test system for
educational purposes-basic distribution system data and results," IEEE
Transactions on Power Systems, vol. 6, no. 2, pp. 813-820, 1991.
[120] D. Clements and P. Mancarella, "Systemic modelling and integrated assessment
of asset management strategies and staff constraints on distribution network
reliability," Electric Power Systems Research, vol. 155, pp. 164-171, 2018/02/01/
2018.
[121] D. Wu, D. C. Aliprantis, and K. Gkritza, "Electric Energy and Power Consumption
by Light-Duty Plug-In Electric Vehicles," IEEE Transactions on Power Systems, vol.
26, no. 2, pp. 738-746, 2011.

193
Appendix A: Network Data

Appendix A: Network Data


This appendix provide the network data required to perform reliability studies on
the test systems used within this thesis. In all cases, a system base power of 100 MVA
is used.

A.1. IEEE 14-bus Test System


A single line diagram of IEEE-14 bus test system is shown in Figure 3-4. It consists of
five synchronous machines with IEEE type-I exciters, three of which are synchronous
compensators used only for reactive power support. There are eleven loads in the
system totaling 259 MW, 73.5 Mvar and 3445 customers. The original 14-bus test
case does NOT have line limits. Compared to 1990's power systems, it has low base
voltages and an overabundance of voltage control capability. The reliability data and
branch data for the modified 14-bus network is presented in Table 4-2 and Table 4-3.

A.1.1 Bus Data


The bus data for 14-bus network is shown in Table A-1, where bus 1 is the slack bus.

Table A-1 Bus Data of IEEE 14-bus Test System


Bus
type Pd Qd Gs Bs Vm Va baseKV area Vmax Vmin
Num
1 3 0 0 0 0 1.06 0 69 1 1.06 0.94
2 2 21.7 12.7 0 0 1.045 -4.98 69 1 1.06 0.94
3 2 94.2 19 0 0 1.01 -12.72 69 1 1.06 0.94
4 1 47.8 -3.9 0 0 1.019 -10.33 69 1 1.06 0.94
5 1 7.6 1.6 0 0 1.02 -8.78 69 1 1.06 0.94
6 2 11.2 7.5 0 0 1.07 -14.22 13.8 1 1.06 0.94
7 1 0 0 0 0 1.062 -13.37 13.8 1 1.06 0.94
8 2 0 0 0 0 1.09 -13.36 18 1 1.06 0.94
9 1 29.5 16.6 0 19 1.056 -14.94 13.8 1 1.06 0.94
10 1 9 5.8 0 0 1.051 -15.1 13.8 1 1.06 0.94
11 1 3.5 1.8 0 0 1.057 -14.79 13.8 1 1.06 0.94
12 1 6.1 1.6 0 0 1.055 -15.07 13.8 1 1.06 0.94
13 1 13.5 5.8 0 0 1.05 -15.16 13.8 1 1.06 0.94
14 1 14.9 5 0 0 1.036 -16.04 13.8 1 1.06 0.94

194
Appendix A: Network Data

A.1.2 Branch Data


The original branch impedance data for 14-bus network is shown in Table A-2.

Table A-2 Branch Data of IEEE 14-bus Test System


Branch
fbus tbus R (p.u.) X (p.u.) B (p.u.) ratio
Num
1 1 2 0.01938 0.05917 0.0528 0
2 1 2 0.01938 0.05917 0.0528 0
3 1 5 0.05403 0.22304 0.0492 0
4 2 3 0.04699 0.19797 0.0438 0
5 2 4 0.05811 0.17632 0.034 0
6 2 5 0.05695 0.17388 0.0346 0
7 3 4 0.06701 0.17103 0.0128 0
8 4 5 0.01335 0.04211 0 0
9 4 7 0 0.20912 0 0.978
10 4 9 0 0.55618 0 0.969
11 5 6 0 0.25202 0 0.932
12 6 11 0.09498 0.1989 0 0
13 6 12 0.12291 0.25581 0 0
14 6 13 0.06615 0.13027 0 0
15 7 8 0 0.17615 0 0
16 7 9 0 0.11001 0 0
17 9 10 0.03181 0.0845 0 0
18 9 14 0.12711 0.27038 0 0
19 10 11 0.08205 0.19207 0 0
20 12 13 0.22092 0.19988 0 0
21 13 14 0.17093 0.34802 0 0

A.1.3 Generator Data


The generator output and cost data for 14-bus network is shown in Table A-3.

Table A-3 Generator Data of IEEE 14-bus Test System


Gen Cost
bus Pmax Pmin Qmax Qmin Vg
Num startup c2 c1 c0
1 1 332.4 0 10 0 1.06 0 0.04302926 20 0
2 2 140 0 50 -40 1.045 0 0.25 20 0
3 3 100 0 40 0 1.01 0 0.01 40 0
4 6 100 0 24 -6 1.07 0 0.01 40 0
5 8 100 0 24 -6 1.09 0 0.01 40 0

195
Appendix A: Network Data

A.2. IEEE 24-bus Reliability Test System (RTS-96)


A single line diagram of IEEE-24 bus reliability test system is shown in Figure 4-5. It
consists of 22 generators, and 38 branches (originally including 31 lines, 2 cables and
5 transformers). There are eighteen loads in the system totaling 2742 MW, 558 Mvar
and 3955 customers. The original 24-bus test system does have line limits, and
corresponding line and cable sizes are selected to match the original line ratings,
which is displayed in Table 4-2.

A.2.1 Bus Data


The bus data for 24-bus network is shown in Table A-4, where bus 13 is the slack bus.

Table A-4 Bus Data of IEEE 24-bus Reliability Test System


Bus
type Pd Qd Gs Bs Vm Va baseKV area Vmax Vmin
Num
1 2 108 22 0 0 1 0 138 1 1.05 0.95
2 2 97 20 0 0 1 0 138 1 1.05 0.95
3 1 180 37 0 0 1 0 138 1 1.05 0.95
4 1 74 15 0 0 1 0 138 1 1.05 0.95
5 1 71 14 0 0 1 0 138 1 1.05 0.95
6 1 136 28 0 -100 1 0 138 2 1.05 0.9
7 2 125 25 0 0 1 0 138 2 1.05 0.95
8 1 171 35 0 0 1 0 138 2 1.05 0.95
9 1 175 36 0 0 1 0 138 1 1.05 0.95
10 1 195 40 0 0 1 0 138 2 1.05 0.95
11 1 0 0 0 0 1 0 230 3 1.05 0.95
12 1 0 0 0 0 1 0 230 3 1.05 0.95
13 3 265 54 0 0 1 0 230 3 1.05 0.95
14 2 194 39 0 0 1 0 230 3 1.05 0.95
15 2 317 64 0 0 1 0 230 4 1.05 0.95
16 2 100 20 0 0 1 0 230 4 1.05 0.95
17 1 0 0 0 0 1 0 230 4 1.05 0.95
18 2 333 68 0 0 1 0 230 4 1.05 0.95
19 1 181 37 0 0 1 0 230 3 1.05 0.95
20 1 128 26 0 0 1 0 230 3 1.05 0.95
21 2 0 0 0 0 1 0 230 4 1.05 0.95
22 2 0 0 0 0 1 0 230 4 1.05 0.95
23 2 0 0 0 0 1 0 230 3 1.05 0.95
24 1 0 0 0 0 1 0 230 4 1.05 0.95

196
Appendix A: Network Data

A.2.2 Branch Data


The original branch impedance data for 24-bus network is shown in Table A-5.

Table A-5 Branch Data of IEEE 24-bus Reliability Test System


Branch R X B Normal LTE STE
fbus tbus ratio
Num (p.u.) (p.u.) (p.u.) MVA MVA MVA
1 1 2 0.0026 0.0139 0.4611 175 193 200 0
2 1 3 0.0546 0.2112 0.0572 175 208 220 0
3 1 5 0.0218 0.0845 0.0229 175 208 220 0
4 2 4 0.0328 0.1267 0.0343 175 208 220 0
5 2 6 0.0497 0.192 0.052 175 208 220 0
6 3 9 0.0308 0.119 0.0322 175 208 220 0
7 3 24 0.0023 0.0839 0 400 510 600 1.03
8 4 9 0.0268 0.1037 0.0281 175 208 220 0
9 5 10 0.0228 0.0883 0.0239 175 208 220 0
10 6 10 0.0139 0.0605 2.459 175 193 200 0
11 7 8 0.0159 0.0614 0.0166 175 208 220 0
12 8 9 0.0427 0.1651 0.0447 175 208 220 0
13 8 10 0.0427 0.1651 0.0447 175 208 220 0
14 9 11 0.0023 0.0839 0 400 510 600 1.03
15 9 12 0.0023 0.0839 0 400 510 600 1.03
16 10 11 0.0023 0.0839 0 400 510 600 1.02
17 10 12 0.0023 0.0839 0 400 510 600 1.02
18 11 13 0.0061 0.0476 0.0999 500 600 625 0
19 11 14 0.0054 0.0418 0.0879 500 600 625 0
20 12 13 0.0061 0.0476 0.0999 500 600 625 0
21 12 23 0.0124 0.0966 0.203 500 600 625 0
22 13 23 0.0111 0.0865 0.1818 500 600 625 0
23 14 16 0.005 0.0389 0.0818 500 600 625 0
24 15 16 0.0022 0.0173 0.0364 500 600 625 0
25 15 21 0.0063 0.049 0.103 500 600 625 0
26 15 21 0.0063 0.049 0.103 500 600 625 0
27 15 24 0.0067 0.0519 0.1091 500 600 625 0
28 16 17 0.0033 0.0259 0.0545 500 600 625 0
29 16 19 0.003 0.0231 0.0485 500 600 625 0
30 17 18 0.0018 0.0144 0.0303 500 600 625 0
31 17 22 0.0135 0.1053 0.2212 500 600 625 0
32 18 21 0.0033 0.0259 0.0545 500 600 625 0
33 18 21 0.0033 0.0259 0.0545 500 600 625 0
34 19 20 0.0051 0.0396 0.0833 500 600 625 0
35 19 20 0.0051 0.0396 0.0833 500 600 625 0
36 20 23 0.0028 0.0216 0.0455 500 600 625 0
37 20 23 0.0028 0.0216 0.0455 500 600 625 0
38 21 22 0.0087 0.0678 0.1424 500 600 625 0

197
Appendix A: Network Data

A.2.3 Generator Data


The generator output and cost data for 24-bus network is shown in Table A-6.

Table A-6 Generator Data of IEEE 14-bus Test System


Gen Cost
bus Pmax Pmin Qmax Qmin Vg
Num Startup c2 c1 c0
1 1 20 16 10 0 1.035 1500 0 130 400.68
2 1 20 16 10 0 1.035 1500 0 130 400.68
3 1 76 15.2 30 -25 1.035 1500 0.014 16.08 212.31
4 1 76 15.2 30 -25 1.035 1500 0.014 16.08 212.31
5 2 20 16 10 0 1.035 1500 0 130 400.68
6 2 20 16 10 0 1.035 1500 0 130 400.68
7 2 76 15.2 30 -25 1.035 1500 0.014 16.08 212.31
8 2 76 15.2 30 -25 1.035 1500 0.014 16.08 212.31
9 7 100 25 60 0 1.025 1500 0.014 16.08 212.31
10 7 100 25 60 0 1.025 1500 0.053 43.66 781.52
11 7 100 25 60 0 1.025 1500 0.053 43.66 781.52
12 13 197 69 80 0 1.02 1500 0.0072 48.58 832.76
13 13 197 69 80 0 1.02 1500 0.0072 48.58 832.76
14 13 197 69 80 0 1.02 1500 0.0072 48.58 832.76
15 14 0 0 200 -50 0.98 1500 0 0 0
16 15 12 2.4 6 0 1.014 1500 0.328 56.56 86.38
17 15 12 2.4 6 0 1.014 1500 0.328 56.56 86.38
18 15 12 2.4 6 0 1.014 1500 0.328 56.56 86.38
19 15 12 2.4 6 0 1.014 1500 0.328 56.56 86.38
20 15 12 2.4 6 0 1.014 1500 0.328 56.56 86.38
21 15 155 54.3 80 -50 1.014 1500 0.008 12.39 382.24
22 16 155 54.3 80 -50 1.017 1500 0.008 12.39 382.24
23 18 400 100 200 -50 1.05 1500 0.00021 4.42 382.24
24 21 400 100 200 -50 1.05 1500 0.00021 4.42 382.24
25 22 50 10 16 -10 1.05 1500 0 0.001 0.001
26 22 50 10 16 -10 1.05 1500 0 0.001 0.001
27 22 50 10 16 -10 1.05 1500 0 0.001 0.001
28 22 50 10 16 -10 1.05 1500 0 0.001 0.001
29 22 50 10 16 -10 1.05 1500 0 0.001 0.001
30 22 50 10 16 -10 1.05 1500 0 0.001 0.001
31 23 155 54.3 80 -50 1.05 1500 0.0083 12.39 382.24
32 23 155 54.3 80 -50 1.05 1500 0.0083 12.39 382.24
33 23 350 140 150 -25 1.05 1500 0.0049 11.85 665.11

198
Appendix A: Network Data

A.2.4 Branch Reliability Data


The branch reliability data for 24-bus network is shown in Table A-7Table A-2.

Table A-7 Branch Reliability Data of IEEE 24-bus Reliability Test System
OHL outage OHL outage UGC outage UGC outage
Branch Length
rate duration rate duration
Num miles
occ./yr hrs/yr occ./yr hrs/yr
1 3 0.2356 10 0.2446 16
2 55 0.506 10 0.567 94
3 22 0.3344 10 0.3624 44.5
4 33 0.3916 10 0.4306 61
5 50 0.48 10 0.536 86.5
6 31 0.3812 10 0.4182 58
8 27 0.3604 10 0.3934 52
9 23 0.3396 10 0.3686 46
10 16 0.3032 10 0.3252 35.5
11 16 0.3032 10 0.3252 35.5
12 43 0.4436 10 0.4926 76
13 43 0.4436 10 0.4926 76
18 33 0.4022 11 0.4254 48
19 29 0.3886 11 0.4102 44
20 33 0.4022 11 0.4254 48
21 67 0.5178 11 0.5546 82
22 60 0.494 11 0.528 75
23 27 0.3818 11 0.4026 42
24 12 0.3308 11 0.3456 27
25 34 0.4056 11 0.4292 49
26 34 0.4056 11 0.4292 49
27 36 0.4124 11 0.4368 51
28 18 0.3512 11 0.3684 33
29 16 0.3444 11 0.3608 31
30 10 0.324 11 0.338 25
31 73 0.5382 11 0.5774 88
32 18 0.3512 11 0.3684 33
33 18 0.3512 11 0.3684 33
34 27.5 0.3835 11 0.4045 42.5
35 27.5 0.3835 11 0.4045 42.5
36 15 0.341 11 0.357 30
37 15 0.341 11 0.357 30
38 47 0.4498 11 0.4786 62

199
Appendix A: Network Data

A.3. Roy Billinton Reliability Test System (RBTS bus 4)


The RBTS bus 4 test network consists of 7 feeders (F1–F7), 38 load points (LP1–
LP38), disconnecting switches on both sides for each feeder cable and 4 tie-switches.
Disconnecting switches are positioned before and after each load bus, the tie-
switches (TS1-TS4) are normally open. Only one circuit breaker is equipped on the
top of each feeder. The default peak load of the network is the actual 40MW and
13MVar. A single line diagram of RBTS bus 4 test network is shown in Figure 6-4 by
removing all the EV components.

A.3.1 Load Data


The load point data for RBTS bus 4 network is shown in Table A-8.

Table A-8 Load Point Data of RBTS bus 4 network


Load point Customer Number of base
Pd Qd
Number Type Customer KV
1-4, 11-13, 18-21, 32-35 0.8869 0.2915 Residential 220 11
5, 14, 15, 22, 23, 36, 37 0.8137 0.2675 Residential 200 11
8, 10, 26-30 1.63 0.5358 Small user 1 11
9, 31 2.445 0.8036 Small user 1 11
6, 7, 16, 17, 24, 25, 38 0.6714 0.2207 Commercial 10 11

A.3.2 Reliability Data


The reliability data for RBTS bus 4 network is shown in Table A-9.

Table A-9 Reliability Data of RBTS bus 4 network


Component Outage Outage Maintenance Maintenance
Voltage
rate duration outage rate duration
Level
occ./yr.km hrs/yr.km occ./yr.km hrs/yr.km
Transformers 138/33 0.01 15 0.5 168
33/11 0.015 15 1.0 120
11/0.415 0.015 10
Breakers 138 0.0058 8 0.2 108
33 0.002 4 0.5 96
11 0.006 4 1.0 72
Busbars 33 0.001 2 0.5 8
11 0.001 2 1.0 8
Lines 33 0.046 8 0.5 8
11 0.065 5
Cables 11 0.04 30

200
Appendix A: Network Data

A.4. Chronological Load Profile


The conventional IEEE RTS-96 hourly chronological load profile was used for all
modified test systems in this thesis. The profile is presented below in Table A-10,
Table A-11, and Table A-12.

Table A-10 Hourly load profile in percentage of peak load


Winter Weeks Summer Weeks Spring/Fall Weeks
Hour 1-8 & 44-53 18-30 9-17 & 31-43
Weekly Weekend Weekly Weekend Weekly Weekend
1 67% 78% 64% 74% 63% 75%
2 63% 72% 60% 70% 62% 73%
3 60% 68% 58% 66% 60% 69%
4 59% 66% 56% 65% 58% 66%
5 59% 64% 56% 64% 59% 65%
6 60% 65% 58% 62% 65% 65%
7 74% 66% 64% 62% 72% 68%
8 86% 70% 76% 66% 85% 74%
9 95% 80% 87% 81% 95% 83%
10 96% 88% 95% 86% 99% 89%
11 96% 90% 99% 91% 100% 92%
12 95% 91% 100% 93% 99% 94%
13 95% 90% 99% 93% 93% 91%
14 95% 88% 100% 92% 92% 90%
15 93% 87% 100% 91% 90% 90%
16 94% 87% 97% 91% 88% 86%
17 99% 91% 96% 92% 90% 85%
18 100% 100% 96% 94% 92% 88%
19 100% 99% 93% 95% 96% 92%
20 96% 97% 92% 95% 98% 100%
21 91% 94% 92% 100% 96% 97%
22 83% 92% 93% 93% 90% 95%
23 73% 87% 87% 88% 80% 90%
24 63% 81% 72% 80% 70% 85%

201
Appendix A: Network Data

Table A-11 Daily load profile in percentage of peak load


Day Peak Load [%]
Monday 93%
Tuesday 100%
Wednesday 98%
Thursday 96%
Friday 94%
Saturday 77%
Sunday 75%

Table A-12 Weekly load profile in percentage of peak load


Week Peak Load [%] Week Peak Load [%] Week Peak Load [%]
1 86.2% 18 83.7% 36 70.5%
2 90.0% 19 87.0% 37 78.0%
3 87.8% 20 88.0% 38 69.5%
4 83.4% 21 85.6% 39 72.4%
5 88.0% 22 81.1% 40 72.4%
6 84.1% 23 90.0% 41 74.3%
7 83.2% 24 88.7% 42 74.4%
8 80.6% 25 89.6% 43 80.0%
9 74.0% 26 86.1% 44 88.1%
10 73.7% 27 75.5% 45 88.5%
11 71.5% 28 81.6% 46 90.9%
12 72.7% 29 80.1% 47 94.0%
13 70.4% 30 88.0% 48 89.0%
14 75.0% 31 72.2% 49 94.2%
15 72.1% 32 77.6% 50 97.0%
16 80.0% 33 80.0% 51 100.0%
17 75.4% 34 72.9% 52 95.2%
35 72.6% 53 86.2%

202
Appendix B: Cable Design and Rating Data

Appendix B: Cable Design and Rating Data


This appendix provide the cable design and rating data of cables utilized on the
modified test systems within this thesis.

Table B-1 Design properties for cables in 14 and 24-bus network


Cable Type 14-bus 24-bus
Voltage level (kV) 69 69 69 13.8 13.8 13.8 230 138
Core number Single Single Single Single Single Single Single Single
Conductor material Cu Cu Cu Cu Cu Cu Cu Cu
Conductor area (mm2) 630 400 185 800 400 185 1200 400
AC resistance at 90°C (Ω/km) 0.038 0.063 0.128 0.035 0.063 0.128 0.024 0.062
DC resistance at 20°C (Ω/km) 0.028 0.047 0.099 0.022 0.047 0.099 0.015 0.047
Cable Dimensions (mm)
Cable outer diameter 76 62 60 52.5 40.5 32.5 113.9 75
Conductor diameter 30.3 23.5 15.9 34.2 23.3 16 42.9 23.5
Conductor screen outer diameter 33.3 25.5 18.9 35.8 24.9 17.2 45.5 25.5
Insulation outer diameter 55.5 47.5 40.9 44.8 33.9 26.2 89.5 59.5
Insulation screen outer diameter 58.1 49.5 43.5 46.2 35.5 27.4 91.9 61.5
Sheath outer diameter 69.2 56.6 54 47.5 36.1 28.5 104.2 68.8
Copper wire sheath area (mm2) 85 50 95 35 35 35 150 95
Thickness of PVC oversheath 3.4 2.7 3 2.5 2.2 2.0 4.4 3.1
Cable Material Properties
XLPE thermal resistance 3.5 3.5 3.5 3.5 3.5 3.5 3.5 3.5
XLPE thermal capacitance (106) 2.57 2.57 2.57 2.57 2.57 2.57 2.57 2.57
Soil density (kg/m^3) 1300 1300 1300 1300 1300 1300 1300 1300
Soil capacitance (J/kg.K) 500 500 500 500 500 500 500 500
Soil thermal resistance (°Cm/W) 0.9 0.9 0.9 0.9 0.9 0.9 1 1
Cable Ratings
Normal rating (A) 1010 790 530 1200 845 535 1130 715
LTE rating (A) 1114 871 584 1323 932 590 1246 789
STE rating (A) 1228 961 645 1371 966 651 1291 817
Laying Conditions
Laying depth (m) 1.5 1.5 1.3 0.8 0.8 0.8 1.2 1.2
Laying formation flat flat flat flat flat trefoil flat flat
Core separation (m) 0.18 0.18 0.12 0.1 0.1 touching 0.22 0.15
Duct or direct buried direct direct direct direct direct direct direct direct
Duct outer diameter (m) 0 0 0 0 0 0 0 0
Duct inner diameter (m) 0 0 0 0 0 0 0 0

203
Appendix B: Cable Design and Rating Data

Table B-2 Design properties for cables in RBTS and windfarm circuit
Cable Type Windfarm RBTS bus 4
Voltage level (kV) 33 11 11 11 11
Core number Single Single Single Single Single
Conductor material Cu Cu Cu Cu Cu
Conductor area (mm2) 630 300 240 185 150
AC resistance at 90°C (Ω/km) 0.04 0.062 0.077 0.1 0.124
DC resistance at 20°C (Ω/km) 0.028 0.0601 0.0754 0.0991 0.124
Cable Dimensions (mm)
Cable outer diameter 60.5 34.5 32.5 30 28
Conductor diameter 31 20.6 18.4 16 14.3
Conductor screen outer
32.6 22.2 20 17.6 15.9
diameter
Insulation outer diameter 48.6 29 26.8 24.4 22.7
Insulation screen outer diameter 50.2 30.6 28.4 26 24.3
Sheath outer diameter 55.1 31.3 29.3 26.96 25.14
Copper wire sheath area (mm2) 35 35 35 35 35
Thickness of PVC oversheath 2.7 1.6 1.6 1.52 1.43
Cable Material Properties
XLPE thermal resistance 3.5 3.5 3.5 3.5 3.5
XLPE thermal capacitance (106) 2.57 2.57 2.57 2.57 2.57
Soil density (kg/m^3) 1300 1300 1300 1300 1300
Soil capacitance (J/kg.K) 500 500 500 500 500
Soil thermal resistance (°Cm/W) 1.2 0.9 0.9 0.9 0.9
Cable Ratings
Summer normal rating (A) 745 683 608 525 467
Summer cyclic rating (A) 772 687 593 528
Summer emergency rating (A) 849 756 653 580
Winter normal rating (A) 706 629 544 484
Winter cyclic rating (A) 798 711 615 547
Winter emergency rating (A) 878 782 676 602
Laying Conditions
Laying depth (m) 0.8 0.8 0.8 0.8 0.8
Laying formation trefoil trefoil trefoil trefoil trefoil
Core separation (m) touching touching touching touching touching
Duct or direct buried Ducted Direct Direct Direct Direct
Duct outer diameter (m) 160 0 0 0 0
Duct inner diameter (m) 150 0 0 0 0

204
Appendix C: Publications

Appendix C: Publications

International Journal Publications

K. Kopsidas and S. Liu, "Power Network Reliability Framework for Integrating Cable
Design and Ageing," IEEE Transactions on Power Systems, vol. 33, no. 2, pp. 1521-1532,
March 2018. Accepted and Published

Submitted Journal Publications

S. Liu and K. Kopsidas, "Increasing Wind Integration through Cable Flexible Current
Rating," IEEE Transactions on Power Systems, Sept 2018. Under Review

S. Liu and K. Kopsidas, "Boosting Network Flexibility using Extensive Emergency


Ratings," IEEE Transactions on Power Systems, 2018. Under Revision

International Conference Publications

S. Liu, C. Cruzat, and K. Kopsidas, "Impact of transmission line overloads on network


reliability and conductor ageing," in 2017 IEEE Manchester PowerTech, 2017, pp. 1-6.
Accepted for oral presentation and published.

M. Abogaleela, K. Kopsidas, C. Cruzat, and S. Liu, "Reliability evaluation framework


considering OHL emergency loading and demand response," in 2017 IEEE PES
Innovative Smart Grid Technologies Conference Europe (ISGT-Europe), 2017, pp. 1-6.
Accepted for oral presentation and published.

S. Liu and K. Kopsidas, "Reliability evaluation of distribution networks incorporating


cable electro-thermal properties," in 2016 Power Systems Computation Conference
(PSCC), 2016, pp. 1-7. Accepted for oral presentation and published.

C. Cruzat, K. Kopsidas, and S. Liu, "Evaluating the impact of information and


communication technologies on network reliability," in 2016 18th Mediterranean
Electrotechnical Conference (MELECON), 2016, pp. 1-7. Accepted for oral presentation
and published.

205

You might also like