Vous êtes sur la page 1sur 194

ABSTRACT

SMITH, JUSTIN WADE. Model Based Fault Locating on Distribution Feeders. (Under the
direction of Dr. Mesut E Baran.)
As electric power systems grow in complexity and size, ensuring system reliability and continuity has become a major concern. Failure to locate faults quickly can result in prolonged
outage times, customer safety concerns and lost revenues. Over the past few decades distribution networks have evolved into large and complex networks capable of carrying thousands
of customers. As of late, utilities have displayed particular interest in distribution fault location technology to reduce such widespread impacts. This thesis aims to develop a modern
distribution fault locating algorithm.
The proposed fault locating algorithm is a model based algorithm that uses a short circuit
model of the feeder to locate the fault. In the short circuit model, a fault is placed at every
node and the observed fault current at the substation is recorded in tabular format. During an
actual fault condition, the algorithm compares the recorded fault current at the substation to
the short circuit modelling data. This comparison allows the fault locator to identify the exact
node in the network that is faulty.
In many cases, the short circuit model will indicate that there are several unique nodes with
the same fault current magnitude. This problem is overcome with a proposed sub-algorithm
called the localization algorithm. This algorithm uses protective device data and load flow data
to determine the protective device that has interrupted the fault. By knowing the protective
device that has interrupted the fault, the fault can be localized to a specific region of the
network.
To test the fault locating algorithm, four different test cases are proposed. Three test cases
are performed using a Progress Energy Carolinas feeder model. Detailed fuse, recloser and
substation breaker models are inserted into the feeder model to test popular protective schemes
such as fuse saving and fuse blowing. Finally, the Stewart Street 12.47kV feeder is tested using
DEW models provided by Allegheny Power. The Stewart Street feeder is much larger than the
Progress Energy Carolinas feeder model and contains hundreds of nodes and tapped loads.
The proposed localization sub-algorithm directly relies on the accuracy of the load flow
solution. In the final test case, we assume the load to be a gaussian random variable by which
the known load is perturbed about a mean load operation point. This approach introduces load
uncertainty and measures the fault locating algorithms sensitivity to load variability.

Model Based Fault Locating on Distribution Feeders

by
Justin Wade Smith

A thesis submitted to the Graduate Faculty of


North Carolina State University
in partial fulfillment of the
requirements for the Degree of
Master of Science

Electrical Engineering

Raleigh, North Carolina


2013

APPROVED BY:

Dr. Subhashish Bhattacharya

Dr. Mesut E Baran


Chair of Advisory Committee

Dr. Srdjan Lukic

DEDICATION

To my parents Jeffrey W. Smith and Cindy A. Smith. Without you, none of this would have
been possible. You gave me strength when I had nothing left and encouragement when I
couldnt carry on. This one is for you.
To my grandfather Harold W. Smith who never got to see my thesis completed.
To my friend Asa Gray, who encouraged me. His hard work and dedication showed me that
the best things in life are the hardest to get.

ii

BIOGRAPHY

Justin Wade Smith was born in Salisbury, North Carolina, United States of America. Justin
attended high school at Northwood High in Chatham County, North Carolina and graduated
in 2004. He received a Bachelor of Science in Electrical Engineering from North Carolina State
University in 2009. Upon completion of his bachelors degree he joined the graduate school at
NCSU working towards his Master of Science degree in Electrical Engineering with an emphasis
on power system protection and fault locating under the direction of Dr. Mesut Baran with
the NSF FREEDM Systems Center.

iii

ACKNOWLEDGEMENTS

I would like to thank my advisor Dr. Mesut E. Baran, and my other committee members, Dr.
Subhashish Bhattacharya and Dr. Srdjan Lukic.
I would also like to acknowledge Larry Alesi at Schweitzer Engineering Laboratories(SEL) for
his guidance. Much of my knowledge in the field of power system protection can be attributed
to the mentorship and dedication of Mr. Alesi. He is true mentor and friend.
I would like to thank Bette Gray of Rodanthe, N.C. for helping me chase my dreams. My family
could not have been blessed with a greater friend.
I would like to thank my friends Nick Parks and Jon McDonald. Their support and friendship
led me to success.
I would also like to thank my friends at Progress Energy Carolinas: Ronald Chip Moore,
Juan Yancey and Lawrence Roberts.

iv

TABLE OF CONTENTS

LIST OF TABLES. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix
LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi
Chapter 1 Introduction . . . . . . . . . . . . . .
1.1 Introduction . . . . . . . . . . . . . . . . . .
1.2 Challenges of Distribution Fault Locating .
1.3 Proposed Solution: Model Based Algorithm
1.4 Glossary of Terms . . . . . . . . . . . . . .

.
.
.
.
.

.
.
.
.
.

1
1
1
2
4

Chapter 2 Modern Distribution Fault Locating Algorithms . . . . . . . . . . .


2.1 Introduction to Distribution Fault Locating Algorithms . . . . . . . . . . . . .
2.2 Technique with Two-port Network Section Representation(Das Method) . . . .
2.2.1 Overview of Das Method . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.2 Data Acquisition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.3 Fault Detection and Classification . . . . . . . . . . . . . . . . . . . . .
2.2.4 Developing an Equivalent Radial Network . . . . . . . . . . . . . . . . .
2.2.5 Load Modelling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.6 Estimating nodal pre-fault voltages and currents . . . . . . . . . . . . .
2.2.7 Estimating Voltages and Currents at the Remote End and at the Fault
2.2.8 Calculating the Distance to the Fault: Single Line to Ground . . . . . .
2.2.9 Assessment of the Das Algorithm: Advantages and Disadvantages . . .
2.3 Girgis Apparent Impedance Method . . . . . . . . . . . . . . . . . . . . . . . .
2.3.1 Overview of Girgis Method . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.2 Direct Determination of Distance to the Fault . . . . . . . . . . . . . . .
2.3.3 Assessment of the Girgis Algorithm: Advantages and Disadvantages . .
2.4 Fault Locating using Digital Fault Recorder Data(Saha Algorithm) . . . . . . .
2.4.1 Overview of Saha Method . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.2 Fault Loop Impedance Determination . . . . . . . . . . . . . . . . . . .
2.4.3 Determination of the Faulty Node . . . . . . . . . . . . . . . . . . . . .
2.4.4 Distance to the Fault . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.5 Assessment of the Saha Algorithm: Advantages and Disadvantages . . .
2.5 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

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

5
5
5
6
6
6
10
11
12
15
18
19
21
21
21
26
27
27
27
30
34
38
39

Chapter 3 Model Based Fault Locating . . . . . . . . . . . . .


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2 Sampled Data Conditioning . . . . . . . . . . . . . . . . . .
3.3 Steady State Fault Current Extraction . . . . . . . . . . . .
3.4 Fault Tables . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.4.1 Sliding Fault Resolution . . . . . . . . . . . . . . . .
3.4.2 Selecting Possible Fault Locations from Fault Tables
3.4.3 Fault Identification and Fault Table Selection . . . .

.
.
.
.
.
.
.
.

40
40
41
43
44
46
47
47

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

49
49
50
52
53
56
58
59
61
62
64

Chapter 4 Fault Locator Test Results . . . . . . . . . . . . . . . . . . . . . . . . .


4.1 Introduction to the Notional Feeder, Stewart Street Feeder and Test Cases . . .
4.2 Notional Feeder Fault Tables and Short Circuit Data . . . . . . . . . . . . . . .
4.3 Test Case 1: Notional Feeder Testing With No Localization . . . . . . . . . . .
4.3.1 Introduction: Test Conditions and Procedure . . . . . . . . . . . . . . .
4.3.2 FLA Testing for Line-to-Ground Faults . . . . . . . . . . . . . . . . . .
4.3.3 FLA Testing for Line-to-Line Faults . . . . . . . . . . . . . . . . . . . .
4.3.4 Test Case 1 Results: . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.4 Test Case 2: Localization using Load Flow and Protective Devices . . . . . . .
4.4.1 Introduction: Test Conditions and Procedures . . . . . . . . . . . . . .
4.4.2 FLA Testing for Line-to-Ground Faults . . . . . . . . . . . . . . . . . .
4.4.3 A-Ground Fault at Node 17 . . . . . . . . . . . . . . . . . . . . . . . . .
4.4.4 A-Ground Fault at Node 7 . . . . . . . . . . . . . . . . . . . . . . . . .
4.4.5 Limitations of the Localization Algorithm: A-Ground Fault at Node 16
4.5 Test Case 3: Fault Locating on Fuse Saving or Fuse Blowing Schemes . . . . .
4.5.1 Introduction: Notional Feeder with Fuse Blowing Coordination . . . . .
4.5.2 Introduction: Fuse Blowing Scheme Test Conditions . . . . . . . . . . .
4.5.3 FLA Testing for Line-to-Ground Faults: Fuse Blowing . . . . . . . . . .
4.5.4 Line-to-Ground Fault at Node 17 . . . . . . . . . . . . . . . . . . . . . .
4.5.5 Line-to-Ground Fault at Node 6 . . . . . . . . . . . . . . . . . . . . . .
4.5.6 Introduction: Notional Feeder with Fuse Saving Coordination . . . . . .
4.5.7 Introduction: Fuse Saving Scheme Test Conditions . . . . . . . . . . . .
4.5.8 FLA Testing for Line-to-Ground Faults: Fuse Saving . . . . . . . . . . .
4.5.9 Line-to-Ground Fault at Node 10 . . . . . . . . . . . . . . . . . . . . . .
4.5.10 Line-to-Ground Fault at Node 6 . . . . . . . . . . . . . . . . . . . . . .
4.5.11 Limitations of the FLA on Fuse Blowing or Fuse Saving Coordinated
Feeders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.6 Notional Feeder Test Results Summary . . . . . . . . . . . . . . . . . . . . . . .
4.6.1 Test Case 1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.6.2 Test Case 2 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.6.3 Test Case 3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.7 Test Case 4: Fault Locating on Large Scale Feeders . . . . . . . . . . . . . . . .
4.7.1 Introduction: Stewart Street 12.47kV Feeder . . . . . . . . . . . . . . .

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

65
65
66
67
67
67
68
69
70
71
71
72
75
77
79
79
80
80
80
83
85
85
85
86
88

.
.
.
.
.
.
.

90
91
91
91
91
93
93

3.5

3.6
3.7
3.8

3.4.4 Calculation of Fault Currents in the Fault Table: Fault Resistance


Localization Techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.5.1 Localization Using Protective Devices . . . . . . . . . . . . . . . .
3.5.2 Zones of Protection . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.5.3 Localization using Fuse Characteristics . . . . . . . . . . . . . . . .
3.5.4 Estimating Fault Current Through the Fuse . . . . . . . . . . . . .
Localization using Load-Flow Rejection . . . . . . . . . . . . . . . . . . .
3.6.1 Calculating Best Matched Device from Load Flow Rejection . . . .
Combining Localization Data for Best Match . . . . . . . . . . . . . . . .
3.7.1 Final Ranking of Each Possibility . . . . . . . . . . . . . . . . . . .
Localization Flow Chart . . . . . . . . . . . . . . . . . . . . . . . . . . . .

vi

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

4.8

4.7.2 Introduction: Test Conditions and Procedures . . . . . . . . . .


4.7.3 Fault Tables for the Stewart Street Feeder . . . . . . . . . . . . .
4.7.4 Load Flow Analysis on the Stewart Street 12.47kV Feeder . . . .
4.7.5 MATLAB Fault Modelling of the Stewart Street 12.47kV Feeder
4.7.6 Fault at Pole P4622 on the Stewart Street 12.47kV Feeder . . . .
4.7.7 FLA Test Results for P4622 Fault . . . . . . . . . . . . . . . . .
4.7.8 Fault at Pole P4266 on the Stewart Street 12.47kV Feeder . . . .
4.7.9 FLA Test Results for P4266 Fault . . . . . . . . . . . . . . . . .
4.7.10 FLA Test Results Summary for Stewart Street Feeder . . . . . .
Sensitivity Analysis: System Load Perturbations . . . . . . . . . . . . .
4.8.1 Introduction: Sensitivity Analysis . . . . . . . . . . . . . . . . .
4.8.2 System Load Perturbations: Test Conditions and Procedures . .
4.8.3 System Load Perturbations on Test Case 1-No Localization . . .
4.8.4 System Load Perturbations on Test Case 2 . . . . . . . . . . . .
4.8.5 System Load Perturbations on Test Case 3-Fuse Saving . . . . .
4.8.6 System Load Perturbations on Test Case 3-Fuse Blowing . . . .

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

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

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

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

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

94
95
95
96
98
103
105
109
111
112
112
112
114
116
118
120

Chapter 5 FLA Testing Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . 122


References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
Appendices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Appendix A Notional Feeder Load Flow Data . . . . . . . . . . . . .
A.1 Notional Feeder Real Power Flow Data . . . . . . . . . . . .
A.2 Notional Feeder Reactive Power Flow Data . . . . . . . . .
A.3 12kV Capacitor Bank Data . . . . . . . . . . . . . . . . . .
A.4 Transformer Bank Data . . . . . . . . . . . . . . . . . . . .
A.5 Source Impedance Data . . . . . . . . . . . . . . . . . . . .
A.6 Line Impedance Data . . . . . . . . . . . . . . . . . . . . . .
A.6.1 Positive Sequence Line Impedance Data . . . . . . .
A.6.2 Negative Sequence Line Impedance Data . . . . . . .
A.6.3 Zero Sequence Line Impedance Data . . . . . . . . .
A.6.4 Positive and Zero Sequence Shunt Capacitance Data
A.7 Relay Settings and Fuse Characteristics . . . . . . . . . . .
A.8 Fuse Characteristics . . . . . . . . . . . . . . . . . . . . . .
Appendix B Modelling of the Notional Feeder using MATLAB . . .
B.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . .
B.2 Modelling of Lines . . . . . . . . . . . . . . . . . . . . . . .
B.3 Modelling of System Loads . . . . . . . . . . . . . . . . . .
B.3.1 Load Modelling Under Fault Conditions . . . . . . .
B.4 Modelling of Sources . . . . . . . . . . . . . . . . . . . . . .
B.5 Modelling of Capacitor Banks . . . . . . . . . . . . . . . . .
B.6 Modelling of Multi-Winding Transformers . . . . . . . . . .
B.7 Modelling of Feeder Transformer . . . . . . . . . . . . . . .

vii

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

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

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

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

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

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

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

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

. 126
. 127
. 127
. 129
. 130
. 130
. 130
. 131
. 131
. 132
. 133
. 134
. 135
. 136
. 138
. 138
. 138
. 141
. 142
. 143
. 143
. 144
. 145

Appendix C Distribution Fault Analysis . . . . . . . . . . . . . . . . . . . . . . . . . 147


C.1 System Fault Currents and Voltages via Numerically Computed Thevenin
Equivalents and Sensitivity Matrices . . . . . . . . . . . . . . . . . . . . . . 147
C.1.1 Thevanins Theorem and Superposition Principle . . . . . . . . . . . 148
C.1.2 Pre-Fault and Faulted Systems in DEW . . . . . . . . . . . . . . . . 151
C.1.3 Forming the Phase Thevanin Matrix . . . . . . . . . . . . . . . . . . 155
C.1.4 System Fault Characteristics . . . . . . . . . . . . . . . . . . . . . . 158
C.1.5 3-Phase Faults . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160
C.1.6 Phase-to-Phase-Ground Faults . . . . . . . . . . . . . . . . . . . . . 162
C.1.7 Phase-to-Phase Faults . . . . . . . . . . . . . . . . . . . . . . . . . . 164
C.1.8 Single Line-to-Ground Faults . . . . . . . . . . . . . . . . . . . . . . 166
C.2 Verification of DEW Fault Current Calculation: Example Feeder . . . . . . 167
C.2.1 Validation with MATLAB Simulink Model . . . . . . . . . . . . . . 173
Appendix D Supplemental FLA Simulation Results for Stewart Street 12.47kV Feeder174
D.1 Pole P4622 Supplemental Recorded Results . . . . . . . . . . . . . . . . . . 174
D.2 Pole P4266 Supplemental Recorded Results . . . . . . . . . . . . . . . . . . 178

viii

LIST OF TABLES

Table 3.1
Table 3.2
Table
Table
Table
Table
Table
Table
Table
Table
Table
Table
Table
Table
Table
Table
Table
Table
Table
Table
Table
Table
Table
Table
Table
Table
Table
Table

Fault Table for Figure 3.7 . . . . . . . . . . . . . . . . . . . . . . . . . . . 46


Increased Resolution Fault Table for Figure 3.7 . . . . . . . . . . . . . . . 47

4.1
4.2
4.3
4.4
4.5
4.6
4.7
4.8
4.9
4.10
4.11
4.12
4.13
4.14
4.15
4.16
4.17
4.18
4.19
4.20
4.21
4.22
4.23
4.24
4.25
4.26

Notional Feeder Fault Table for SLG Faults on A Phase. . . . . . .


Line-to-Ground Simulation Results for Test Case 1 . . . . . . . . .
Phase-to-Phase Fault Simulation Results for Test Case 1 . . . . . .
Percent Mismatch Table for Protective Device Localization. . . . .
Percent Mismatch Table for Load Rejection Localization . . . . . .
Percent Mismatch Table for Protective Device Localization . . . .
Percent Mismatch Table for Load Rejection Localization . . . . . .
Percent Mismatch Table for Protective Device Localization . . . .
Percent Mismatch Table for Load Rejection Localization . . . . . .
Percent Mismatch Table for Protective Device Localization . . . .
Percent Mismatch Table for Load Rejection Localization . . . . . .
Percent Mismatch Table for Protective Device Localization . . . .
Percent Mismatch Table for Load Rejection Localization . . . . . .
Percent Mismatch Table for Protective Device Localization . . . .
Percent Mismatch Table for Load Rejection Localization . . . . . .
Substation Bus Load Flow Solution for Stewart Street Feeder. . . .
Pre-Fault Load Flow Values . . . . . . . . . . . . . . . . . . . . . .
Fault Analysis Calculated Parameters . . . . . . . . . . . . . . . .
MATLAB Model Parameters . . . . . . . . . . . . . . . . . . . . .
Pre-Fault DEW Load Flow Values . . . . . . . . . . . . . . . . . .
Success-Failure Rate on Node 10, Test Case 1(No Localization) . .
Success-Failure Rate on Node 17, Test Case 1(No Localization) . .
Success-Failure Rate on Node 4, Test Case 1(No Localization) . . .
Success-Failure Rate on Load Perturbation Test for Node 10 Fault
Success-Failure Rate on Load Perturbation Test for Node 17 Fault
Success-Failure Rate on Load Perturbation Test for Node 10 Fault
Fuse Saving Feeder Coordination. . . . . . . . . . . . . . . . . . . .
Table 4.27 Success-Failure Rate on Load Perturbation Test for Node 17 Fault
Fuse Saving Feeder Coordination. . . . . . . . . . . . . . . . . . . .
Table 4.28 Success-Failure Rate on Load Perturbation Test for Node 10 Fault
Fuse Blowing Feeder Coordination. . . . . . . . . . . . . . . . . . .
Table 4.29 Success-Failure Rate on Load Perturbation Test for Node 17 Fault
Fuse Blowing Feeder Coordination. . . . . . . . . . . . . . . . . . .

. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
. . .
with
. . .
with
. . .
with
. . .
with
. . .

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

66
69
69
74
74
76
77
82
82
84
84
87
87
89
89
96
98
100
102
107
114
115
115
117
117

. 118
. 119
. 120
. 121

Table A.1
Table A.2

Relay Settings for Substation Breaker A and Substation Breaker B . . . . 135


Relay Settings for Recloser A and Recloser B . . . . . . . . . . . . . . . . 135

Table C.1
Table C.2

Fault types and their frequency of occurrence[23]. . . . . . . . . . . . . . . 159


L-G Voltage Results of DEW vs. MATLAB Powerflow Script . . . . . . . 168

ix

Table C.3
Table C.4

Current Results of DEW vs. MATLAB Script . . . . . . . . . . . . . . . . 169


MATLAB Simulink Fault Current Results . . . . . . . . . . . . . . . . . . 173

Table D.1
Table D.2

Other Likely Fault Possibilities for P4622 . . . . . . . . . . . . . . . . . . 177


Other Likely Fault Possibilities for P4266 . . . . . . . . . . . . . . . . . . 180

LIST OF FIGURES

Figure 1.1
Figure 1.2
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure

2.1
2.2
2.3
2.4
2.5
2.6
2.7
2.8

Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure

2.9
2.10
2.11
2.12
2.13
2.14
2.15
2.16
2.17
2.18
2.19

Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure

3.1
3.2
3.3
3.4
3.5
3.6
3.7
3.8
3.9
3.10
3.11
3.12
3.13
3.14
3.15
3.16
3.17

12kV Radial Feeder. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .


FLA High Level Overview. . . . . . . . . . . . . . . . . . . . . . . . . . .

2
3

Radial Distribution Feeder[25]. . . . . . . . . . . . . . . . . . . . . . . . .


Flow Chart for Determining fault type[25]. . . . . . . . . . . . . . . . . .
Line section Bus M and Bus R. . . . . . . . . . . . . . . . . . . . . . . .
Radial Distribution Feeder[25]. . . . . . . . . . . . . . . . . . . . . . . . .
Voltage and Current relationship between M and R[25]. . . . . . . . . . .
Consolidated loads at the remote end, Node N[25]. . . . . . . . . . . . . .
Fault between Nodes x and x + 1(= y)[25]. . . . . . . . . . . . . . . . . .
Fault locator load model as constant impedance with system load model
as constant power [8]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Simple Feeder with no laterals or taps . . . . . . . . . . . . . . . . . . . .
Single line-to-ground fault sequence networks. . . . . . . . . . . . . . . .
B-to-C Fault . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Sequence Network Diagram for a Phase-to-Phase Fault. . . . . . . . . . .
Simple Substation with Two Parallel Feeders during Pre-Fault Conditions.
Simple Substation with Two Parallel Feeders during Fault conditions. . .
Cascaded Line Sections of Distribution Feeder. . . . . . . . . . . . . . . .
Faulted Distribution Feeder[1]. . . . . . . . . . . . . . . . . . . . . . . . .
Circuit Representation of Network for a Fault located between 1 and k. .
Circuit Representation of Network for a Fault located between k and k + 1.
Tapped Load inserted between node 2 and node k. . . . . . . . . . . . . .

7
8
9
10
13
15
16

High-Level Overview of FLA. . . . . . . . . . . . . . . . . . . . . . .


Handling of Raw Data Passed to the Fault Locator. . . . . . . . . .
Sampling of Fault Current Recorded at 64 Samples Per Cycle . . .
Bus Voltage During a Ground Fault. . . . . . . . . . . . . . . . . . .
RMS value of the 60Hz Fundamental during Line-to-Ground Fault.
Algorithm for detecting steady state fault current. . . . . . . . . . .
Generic 6 Node Feeder. . . . . . . . . . . . . . . . . . . . . . . . . .
Phase-to-Phase evolving into a 3-Phase fault[13]. . . . . . . . . . . .
Generic 6 Node Feeder with Fuses and Overcurrent Relay. . . . . .
Operating Times for a 8620 A Fault Through Current. . . . . . . .
Observed Current during Fault Conditions for Figure 3.9. . . . . . .
Zones of Protection for Figure 3.9. . . . . . . . . . . . . . . . . . . .
FLA Fuse Localization Algorithm. . . . . . . . . . . . . . . . . . . .
Superimposed Currents Through Faulted Fuse. . . . . . . . . . . . .
Fault Current Estimation Algorithm in FLA. . . . . . . . . . . . . .
Current Rejection after Recloser Lockout. . . . . . . . . . . . . . . .
Load Flow Rejection Algorithm in the FLA. . . . . . . . . . . . . .

xi

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

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

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

20
22
23
28
28
31
32
33
34
35
36
39
41
42
42
43
44
45
45
50
51
52
53
54
54
56
58
59
60

Figure 3.18 FLA Localization Algorithm Flow Chart. . . . . . . . . . . . . . . . . . . 64


Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure

4.1
4.2
4.3
4.4
4.5
4.6
4.7
4.8
4.9
4.10
4.11
4.12
4.13

Figure 4.32
Figure 4.33

Simulation Testing Procedure using MATLAB Simulink. . . . . . . . . . 67


Test Case 1 Notional Feeder One-Line Model. . . . . . . . . . . . . . . . 68
Test Case 2 Testing Procedure. . . . . . . . . . . . . . . . . . . . . . . . . 70
Notional Feeder One-Line Model with added Reclosers and Breakers. . . 72
Test Case 2: Substation Phase A RMS current during a fault at Node 17. 73
Test Case 2: Substation Phase A RMS current during a fault at Node 7. 75
Fault Locator Ranking of Fault at Node 16. . . . . . . . . . . . . . . . . 78
Fuse Blowing Scheme on Notional Feeder. . . . . . . . . . . . . . . . . . . 79
Fault Locator Observed Fault Current for Fault at Node 17. . . . . . . . 81
Fault Locator Observed Fault Current for a fault at Node 6. . . . . . . . 83
Fault Locator Observed Fault Current for a Fault at Node 10. . . . . . . 86
Fault Locator Observed Fault Current for a Fault at Node 6. . . . . . . . 88
Fault Locator Observed Fault Current for Node 8 Fault(Fuse Blowing
Coordination). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
Stewart Street 12.47kV Feeder Modelled in DEW. . . . . . . . . . . . . . 93
Test Procedure for Stewart Street Feeder using DEW. . . . . . . . . . . . 94
Reflected Fault Current at the Substation Bus for a Fault at Pole P4622. 95
Pre-Fault Short Circuit Model. . . . . . . . . . . . . . . . . . . . . . . . . 97
Faulty Short Circuit Model. . . . . . . . . . . . . . . . . . . . . . . . . . 97
Post-Fault Short Circuit Model. . . . . . . . . . . . . . . . . . . . . . . . 98
P4622 Node Location in Stewart Street 12.47kV Feeder Modelled in DEW. 99
Reflected Fault Current at the Substation due to the fault itself. . . . . . 100
Observed Fault Current at FUSE654421506 during loaded conditions. . . 101
Observed Fault Current at Substation during loaded conditions. . . . . . 101
Possible Fault Locations in Stewart Street 12.47kV Feeder for a fault at
P4622. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
Best Matched Locations for a fault at P4622 as calculated by the FLA. . 104
P4266 Node Location in Stewart Street 12.47kV Feeder Modelled in DEW.105
Upstream Fuse(FUSE1052482746) Load Flow Solution. . . . . . . . . . . 106
Pre-Fault Voltage at P4266. . . . . . . . . . . . . . . . . . . . . . . . . . 106
Reflected Fault Current at the Substation due to the fault itself. . . . . . 107
Observed Fault Current at upstream fuse(FUSE1052482746) during loaded
conditions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
Possible Fault Location in Stewart Street 12.47kV Feeder for a fault at
P4266. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
Best Matched Locations for a fault at P4266 as calculated by the FLA. . 110
Normally Distributed Per-Unit Load Power. . . . . . . . . . . . . . . . . 113

Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure

4.14
4.15
4.16
4.17
4.18
4.19
4.20
4.21
4.22
4.23
4.24

Figure
Figure
Figure
Figure
Figure
Figure

4.25
4.26
4.27
4.28
4.29
4.30

Figure A.1
Figure A.2

SandC T-Speed Fuse Minumum Melt Characteristics. . . . . . . . . . . . 136


SandC T-Speed Fuse Total Clearing Time Characteristics. . . . . . . . . 137

Figure B.1
Figure B.2

Nominal PI Line Model[28]. . . . . . . . . . . . . . . . . . . . . . . . . . 139


Line Model Block in MATLAB Simulink[28]. . . . . . . . . . . . . . . . . 140

Figure 4.31

xii

Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure

B.3
B.4
B.5
B.6
B.7
B.8
B.9
B.10

Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure

C.1
C.2
C.3
C.4
C.5
C.6
C.7
C.8
C.9
C.10
C.11

Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure
Figure

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

143
143
144
144
145
145
146
146

System Thevanin Equivalent Component Model. . . . . . . . . . . . . .


Power System containing a source, load buses and bus F . . . . . . . . .
Power System before a fault occurrence at bus F. . . . . . . . . . . . .
Open switch replaced by voltage source. . . . . . . . . . . . . . . . . . .
Closed Switch Replaced by Two Sources. . . . . . . . . . . . . . . . . .
Circuit during pre-fault conditions. . . . . . . . . . . . . . . . . . . . .
Calculation of currents due to the fault itself. . . . . . . . . . . . . . . .
Converting the network to a thevanin equivalent. . . . . . . . . . . . . .
Pre-Fault system showing several loads and bus N(future faulted bus).
Faulted System with Fault applied at Node N. . . . . . . . . . . . . . .
Pre-Fault system with scaled loads and test load inserted at the faulted
bus. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
C.12 Test load inserted on Phase A of the faulted node. . . . . . . . . . . . .
C.13 Test load inserted on Phase B of the faulted node. . . . . . . . . . . . .
C.14 Test load inserted on Phase C of the faulted node. . . . . . . . . . . . .
C.15 Fault Root Causes and Percent Occurrence[7]. . . . . . . . . . . . . . .
C.16 Model of a 3 Phase Fault. . . . . . . . . . . . . . . . . . . . . . . . . . .
C.17 Model of a Phase-Phase-Ground Fault. . . . . . . . . . . . . . . . . . .
C.18 Model of a Phase-Phase Fault. . . . . . . . . . . . . . . . . . . . . . . .
C.19 Model of a Phase-to-Ground Fault. . . . . . . . . . . . . . . . . . . . .
C.20 Simple Radial Feeder with three load buses. . . . . . . . . . . . . . . .
C.21 DEW Event Report showing the current seen at the substation for a
bolted 3-Phase fault at Bus 1. . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.

148
149
149
150
150
151
151
152
152
153

.
.
.
.
.
.
.
.
.
.

155
156
157
158
158
160
162
164
167
168

Figure D.1
Figure D.2

3-Phase Dynamic Load Representing Exponential Load Functions.


System Thevanin Equivalent. . . . . . . . . . . . . . . . . . . . . .
Three Phase MATLAB Simulink Source. . . . . . . . . . . . . . .
12kV Capacitor Bank. . . . . . . . . . . . . . . . . . . . . . . . . .
Three Winding Transformer. . . . . . . . . . . . . . . . . . . . . .
MATLAB Multi-Winding Transformer Block. . . . . . . . . . . .
Per Phase Representation of 3-Phase Feeder Transformer. . . . . .
MATLAB Three Phase Transformer. . . . . . . . . . . . . . . . .

Worst
Other
FLA. .
Figure D.3 Worst
Figure D.4 Other
FLA. .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

. 172

Matched Locations for a fault at P4622 as calculated by the FLA.


Likely Fault Locations for a fault at P4622 as calculated by the
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Matched Locations for a fault at P4266 as calculated by the FLA.
Likely Fault Locations for a fault at P4266 as calculated by the
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

xiii

175
176
178
179

Chapter 1

Introduction
1.1

Introduction

Modern utilities are required to transmit and distribute electric power over vast regions dependably while keeping costs low for its customers. As electric power systems grow in complexity
and size, ensuring system reliability and continuity has become a major concern. Failure to
locate faults quickly can result in prolonged outage times, customer safety concerns and lost
revenues. Over the past few decades, distribution networks have evolved into large and complex
networks capable of carrying thousands of customers. As of late, utilities have displayed particular interest in distribution fault location technology to reduce such widespread impacts. With
the addition of fault locators, utilities are able to reduce the search radius substantially. This
improvement has allowed utilities to expedite restoration time and reduce economic impact.

1.2

Challenges of Distribution Fault Locating

Currently most of the research being performed in the field of fault location has been focused on transmission networks[1]. Transmission networks are generally very simple, homogeneous throughout and contain few tap lines or branches. The simplicity of transmission
system topology greatly reduces the complexity of the algorithm. Many transmission fault
locating algorithms have been shown to be accurate using basic fault locating techniques and
methods([10],[30]). Unlike transmission networks, distribution systems contain various conductor sizes in addition to many load taps, laterals, and branches. Fault locating on such complex
topologies presents many challenges not present in transmission systems. An example of a
modern radial distribution feeder is shown in Figure 1.1.
One of the challenges of fault locating on distribution networks is the presence of significant
system loading. Unlike transmission networks, distribution systems contain many tapped loads

Figure 1.1: 12kV Radial Feeder.

between the fault locator and the fault itself. Modern distribution networks make it infeasible
to place measurement devices at every branch on the network. This implies that system load
conditions throughout the network remain an unknown quantity. Therefore, distribution fault
location algorithms must be robust enough to handle load uncertainty.
Many modern distribution fault locating algorithms use apparent impedance at the substation to calculate the location of the fault. In many distribution networks there exist multiple
nodes that have the same apparent impedance during fault conditions. This problem results in
multiple calculated possibilities throughout the network. As a result, fault locating algorithms
must be able to localize and rank these possibilities from most to least likely.

1.3

Proposed Solution: Model Based Algorithm

The proposed fault locating algorithm(FLA) is a model based algorithm(MBA) that uses a
short circuit model of the feeder to locate the fault. In the short circuit model, a fault is
placed at every node and the observed fault current at the substation is recorded in tabular

format(fault tables). During an actual fault condition, the algorithm compares the recorded
fault current at the substation to the short circuit modelling data. This comparison allows the
fault locator to identify the exact node in the network that is faulty. A high level overview of
the FLA is shown in Figure 1.2.

Figure 1.2: FLA High Level Overview.

In many cases, the short circuit model will indicate that there are several unique nodes with
the same fault current magnitude. This problem is overcome with a sub-algorithm in the FLA
called the Localization Algorithm. This sub-algorithm uses protective device data and load
flow data to determine the protective device that has interrupted the fault. By knowing the
device that interrupted the fault, we can de-rank all other possibilities as a poor choices.

1.4

Glossary of Terms

MBA : Model Based Algorithm, which is the proposed algorithm in this paper. Short circuit
models are used to locate the faulty node.
FLA: Fault Locating Algorithm
IEEE: Institute of Electrical and Electronics Engineers
NF: Notional Feeder, which refers to the test network built by Progress Energy Carolinas in
MATLAB Simulink.
PEC: Progress Energy Carolinas, referring to the utility based in Raleigh, NC.
P.U.: A per-unit quantity, expressed as a quantity on a defined system base unit.
DFT: Discrete Fourier Transform, referring to the family of techniques based on signal decomposition into sinusoids.
DFR: Digital Fault Recorder.
RMS: Root Mean Square, referring to the process of calculating the quadratic mean, representing the measure of the magnitude of a varying function[12].
EPRI: Electric Power Research Institute
SandC: Refereeing to the fuse and protective device manufacturer.
DEW: Referring to the fault analysis and load flow software package. The Stewart Street
12.47kV feeder was built by Allegheny Power and tested in the software package.
ASPEN One-Liner: Software fault analysis and coordination tool used to check protective
device coordination.

Chapter 2

Modern Distribution Fault Locating


Algorithms
2.1

Introduction to Distribution Fault Locating Algorithms

Fault locating on transmission and distribution networks requires the algorithm to quickly
and accurately calculate the location of the fault. Transmission systems by comparison tend
to be much simpler, containing few lateral taps and homogeneous conductor sizing. Experimental data has shown that modern fault locating algorithms on transmission system to be
very accurate([10],[30]). Distribution systems present a unique challenge for fault locating due
to lateral taps(single and multiphase laterals), complex topologies, load uncertainty and nonhomogeneous nature of the system. This set of challenges makes distribution fault location
distinctly different from transmission[25].
The algorithms presented in this section attempt to confront many of the challenges of
distribution fault locating. Each algorithm is assessed and derived in full detail to show its
strengths and weaknesses. At the end of each algorithm derivation, a section is included to
review the advantages and disadvantages of each.

2.2

Technique with Two-port Network Section Representation(Das


Method)

The following method developed by Ratan Das([25]), uses fundamental voltages and currents
available via measurement devices at the substation. The authors method addresses many
of the known problems with distribution fault location including: non-homogeneous cabling,
system loading, fault impedance and multi-estimate conversion. In the following sections, a
detailed assessment of the method is performed to illustrate advantages and dis-advantage of

the Das method.

2.2.1

Overview of Das Method

The alogrithm proposed by Das uses fundamental voltages and currents at the substation bus
provided by utility measurement equipment. The location of the fault is estimated by computing
the apparent reactance from the fundamental voltage and current phasors at Bus M in Figure
2.1. Once all possible fault locations have been identified, estimates of the voltages and currents
at the fault and remote end are calculated. The final step of the algorithm is to estimate the
distance to the fault from the beginning of the line segment. The Das fault locating algorithm
can be decomposed into seven major steps.
Data Acquisition.
Preliminary estimate of the faulted line section.
Modification of the radial model(Equivalent Model Development).
Load Modelling.
Estimation of voltage and currents at the fault and at the remote end.
Estimating the distance to the fault from the line origination node.
Converting multiple fault locations into a single estimate.

2.2.2

Data Acquisition

After a fault is detected, the fundamental frequency voltage and currents at the substation
bus(Node M) are saved. The data saved includes the pre-fault(fault has not occurred) and
fault(fault has occurred) voltage and current phasors. Once a fault has been detected, pre-fault
voltage and current are saved 1 cycle before the fault occurrence. Fault data is saved 3 cycles
after the fault occurrence to minimize infeed by motors.

2.2.3

Fault Detection and Classification

The first step in many fault locating algorithms is detecting that a fault has occurred. The
Das algorithm uses current thresholds or pickup values to declare that a fault has occurred.
After the fault has been detected, the algorithm attempts to classify the fault condition as one
of the following types: 3-phase, Line-to-Line, Line-to-Ground or Line-to-Line-to-Ground. The
flow chart in Figure 2.2 is used to determine the fault type and the faulty phase A, B or C.

It must also be noted that zero sequence currents are used to discriminate between phaseto-phase faults and phase-to-phase-to-ground faults. If the zero sequence current threshold is
exceeded, this indicates the presence of a ground fault.

Figure 2.1: Radial Distribution Feeder[25].

Estimating the Faulted Line Section


After the fault has occurred, a preliminary estimate of the faulted line section is made. To
make this estimate, detailed knowledge of the line parameters in needed along with the fault
voltages and currents. We first consider a Phase A-to-Ground fault F between nodes x and
x + 1 = y in Figure 2.1. If load is neglected, the apparent impedance from node M to the fault
is defined as:
Zm =

Vam
Iam

(2.1)

We can define the apparent reactance as:



Xm = Im

Vam
Iam


(2.2)

Figure 2.2: Flow Chart for Determining fault type[25].

Let us define the line segment between Bus M and Bus R as the line segment in Figure 2.3.
And the sequence impedance matrix for the line segment between Bus M and Bus R:

Z00 Z01 Z02

Z012 = Z10 Z11 Z12


Z20 Z21 Z22

(2.3)

Assuming de-coupled terms(equal self and mutual impedances):

Z0

0
Z+

Z012 = 0
0

0
Z

(2.4)

Where Z00 = Z 0 , Z11 = Z + and Z22 = Z . We also assume that the negative and positive
sequence impedances are equal for this line segment: Z + = Z

Figure 2.3: Line section Bus M and Bus R.

We can transform the sequence impedance matrix to the phase impedance matrix using the
following transform:

Z0

1
Zabc = T 0
3
0

0 3T 1
Z+

Z+

(2.5)

Where T is:

T = 1 2
1

(2.6)

Where 2 = 16 120 . This yields the following result:

Zabc =

2Z (1)
Z (0)
3 + 3
Z (0)
Z (1)
3 3
Z (0)
Z (1)
3 3

Z (0)
Z (1)
3 3
2Z (1)
Z (0)
3 + 3
Z (0)
Z (1)
3 3

Z (0)
Z (1)
3 3
Z (0)
Z (1)
3 3
2Z (1)
Z (0)
3 + 3

Zaa Zab Zac


= Zba
Zca

Zbb
Zcb

Zbc
Zcc

(2.7)

We can now define the self impedance of phase A, Zaa as:


Zaa =

2Z (1) Z (0)
+
3
3

(2.8)

The author defines the modified reactance as:


Im (Zaa ) = Xaa =

2X (1) X (0)
+
3
3

(2.9)

If the modified reactace is less the apparent reactance than the fault must be located beyond

the remote node. To illustrate this point, if:


(

Xaa < Xm

: Fault is Beyond Node R

Xaa > Xm : Fault is Between Node M and Node R


If the fault is beyond the remote node of the first line section, the next section is added to
the first to obtain the total modified reactance defined by the following general summation:

Xtotal =

N
(1)
X
2Xn
n=1

Xtotal < Xm

(0)

Xn
3

(2.10)

: Fault is Beyond Current Line Section

Xtotal > Xm : Fault is Located on Current Line Section


This general procedure is continued into Xtotal > Xm , indicating the faulted line section.
Given the large size of distribution feeders, there is likely more than one possible fault location. If this occurs, all possible fault locations are recorded and analysed individually. In later
sections, ranking of these possibilities from most likely to lease likely is discussed.

2.2.4

Developing an Equivalent Radial Network

Once all possible fault locations have been established, the radial feeder model with laterals
is converted to a network without laterals. All lateral loads between Bus M and the fault
are consolidated at the tap origination. To illustrate this point we refer back to our radial
distribution feeder shown in Figure 2.1 with lateral taps K and L. To eliminate laterals K and
L, they are lumped with all other loads connected to x 1. The final result is shown in Figure
2.4.

Figure 2.4: Radial Distribution Feeder[25].

10

2.2.5

Load Modelling

System loading on distribution networks can introduce large errors when calculating distance
to the fault along a faulted line segment. To mitigate this, the effects of loads are taken into
account in the Das algorithm. It is assumed that system load is dependent on voltage. We
begin our analysis of system loading by considering the following static exponential load model
for real and reactive power consumed by the load:

P (V, p ) = P0

Q(V, q ) = Q0

V
V0

p

V
V0

 q

(2.11)

(2.12)

The above equations represent the consumed load at various voltages based on a nominal
power and voltage. The terms p and q terms represent the real and reactive power sensitivity
to changes in voltage. To illustrate the sensitivity terms, we begin by defining a relative
sensitivity function[31]:
F
S

F
S

F
F0

0 x0


F 0
=
x0 F0

(2.13)

(2.14)

For Active Power:


P
SV

P
SV

P
P0
V
V0 V0


P V0
=
V V0 P0

(2.15)

(2.16)

For Reactive Power:


Q
SV

Q
SV

Q
Q0
V
V0 V0


Q V0
=
V V0 Q0

11

(2.17)

(2.18)

If we take a derivative of the load model in Equation 2.11 active power with respect to
voltage we get:
P
P0
= p
V
V0

V
V0

p 1
(2.19)

Evaluating V = V0 we get:

P0
P
= p

V V0
V0

(2.20)

P
P0
V
V0

(2.21)

Q
Q0
V
V0

(2.22)

Solving for p :
P

p = S V =
We can also do the same for q :
q =

Q
SV

We can see intuitively that the coefficients p and q represent the load sensitivity to voltage.
Therefore, the apparent power absorbed by the load is:

S(V, p , q ) = P0

V
V0

p


+ jQ0

V
V0

q
(2.23)

Using Equation 2.23 formulation, a voltage dependent impedance value can be derived to
represent the load at various voltages. The values of G0 and B0 are calculated at nominal
voltage and corresponding value of Yload can be calculated at any voltage. This relationship is
extremely important when calculating load currents under fault conditions:
"
Yload (V, p , q ) = G0

2.2.6

V
V0

p 2


+ jB0

V
V0

q 2 #
(2.24)

Estimating nodal pre-fault voltages and currents

A vital part of the Das algorithm is the compensation for system loading. The algorithm compensates for system load by calculating the nominal load impedance and applying it to Equation
2.24. After the nominal pre-fault load impedance is calculated, the appropriate impedance at
any voltage can be calculated. With this result we can easily calculate the load impedance under faulted conditions and compensate for the effects of load. To calculate the load impedance
under pre-fault conditions, the following parameters are needed: pre-fault load voltage, prefault power factor(assumed to be known), percent of the total load at the load node(assumed

12

to be known). The pre-fault voltage is solved for by using a two-port network representation of
each distribution line. Using the two-port model, we can begin at the measurement node and
calculate the corresponding current injection and voltage at each node.
Before calculation of pre-fault values, the algorithm suggested by Das estimates the loads
at all nodes up to the fault using a database. The following formulation is used:
Sload=

Connected Load at Node X


Total Pre-Fault Load
Total Connected Load

(2.25)

The total pre-fault load is measured before the fault occurred by using the voltage and current
phasors measured by the fault locator. It can be see from the above equation that the measured
pre-fault load is apportioned to each node based on the percent loading of the node.
After the apparent power is calculated at each node(Equation 2.25), the load admittance
can be calculated:

Y0 =

V02
Sload

1
(2.26)

The above equation illustrates a very important point of this algorithm, the load impedance
cannot be calculated until the pre-fault load voltage V0 is solved for1 . Beginning at the measured
node, we can calculate the load voltage at each node using the following two-port network
relationship:

Figure 2.5: Voltage and Current relationship between M and R[25].

It is assumed that p and q are known from the power database

13

The two port equation for Figure 2.5:


"

Vr
Irm

"
=

Dmr Bmr
Cmr

#"

Amr

Vm

Imr

(2.27)

Where Dmr , Cmr , Bmr and Amr are as follows:


Dmr = cosh (mr Lmr )

(2.28)

sinh (mr Lmr )


s
Zmr

(2.29)

Cmr =

Bmr = Zmr sinh (mr Lmr )

(2.30)

Amr = cosh (mr Lmr )

(2.31)

s are the propagation constant and characteristic impedance of the


The terms mr and Zmr

line, respectively. We can can calculate each of these terms by using the resistance, reactance,
conductance and suseptance per unit length of the line.
mr =

p
(rmr + jxmr ) (gmr + jbmr )
s
s
Zmr
=

(rmr + jxmr )
(gmr + jbmr )

(2.32)

(2.33)

rmr resistance per unit length

xmr reactance per unit length

gmr conductance per unit length

bmr suseptance per unit length


For short cable lengths in distribution systems the following approximation can be made:
Amr = Dmr = 1

14

s
Bmr = Zmr
mr Lmr

mr Lmr
s
Zmr

Cmr =

Resulting in a two-port model for a short distribution line:


"

Vr

Irm

"
=

Bmr

Cmr

#"

Vm

Imr

(2.34)

All pre-fault voltages and currents are solved for up until the faulty line section(node x)
using the two-port in Equation 2.34. Once the faulty line section has been reached, all loads
beyond the fault are consolidated at the farthest remote node (node N). Using another two-port
equation similar to Equation 2.34, we can solve for the remote end voltage and current. Using
a cascaded two-port model, we can form a cascaded line section equivalent for all nodes from
x to N . This forms a equivalent two port model between the beginning of the faulted line
section(Node x) to the remote node(Node N ).
"

Vn
In

"
=

De Be
Ce

Ae

#"

Vx

Ixf

(2.35)

Where De , Ce , Be and Ae are cascaded line section equivalent constants from node x + 1 to
N.

Figure 2.6: Consolidated loads at the remote end, Node N[25].

2.2.7

Estimating Voltages and Currents at the Remote End and at the Fault

The next step requires the voltages and currents at the fault to be calculated. When the fault
occurs, the fault locator at Node M records the currents and voltages for later use. Since we
have a two-port model representation of each line up until the fault, we can easily calculate

15

the voltage at the beginning of the faulted line section(Node x). Also, as expected, the loads
are present during the fault and must be accounted for. Recall that Equation 2.24 allows us to
calculate the load current at any voltage, even during fault conditions.
Beginning at Node M , the two-port network model and Equation 2.24 can be used to calculate
currents and voltages present at Node R. Using the two-port network model for R to x 1 and
the load model for Node x 1 we can calculate the voltages and currents at node x 1. This
process is completed until you reach the beginning terminal of the faulted line section, Node x.

Figure 2.7: Fault between Nodes x and x + 1(= y)[25].

After the values of Vx and Ixf have been solve for, we can begin the process of calculating
the distance to the fault from Node x. The fault is considered to be s length from Node x and
1 s from Node x + 1(= y). Therefore, we break the two-port model of the line segment x to
x + 1 into two two-port models: one from x to F and the other from F to x + 1(= y).
"

Vf

#
=

If x
"

Vx+1(=y)
If n

"

"
=

sBxy

sCxy

#"

Vx

#
(2.36)

Ixf

(1 s)Bxy

(1 s)Cxy

#"

Vf
If n

#
(2.37)

With the above equations, the current flowing through the fault If is still unknown, along
with the remote end voltage and current. We can relate the currents at the remote node N and
the If n current by:
"

Vn
In

"
=

De Be
Ce

#"

Ae

Vx+1(=y)
If n

#
(2.38)

Where De , Ce , Be and Ae are cascaded line section equivalent constants from node x+1(= y)
to N . If we substitute Equation 2.37 into 2.38, we get:

16

"

Vn

"
=

In

#"

De Be
Ae

Ce

(1 s)Bxy

(1 s)Cxy

#"

Vf

(2.39)

If n

The above equation represents a very important relationship between currents flowing
around the fault and the remote end load voltage and current. If we simplify 2.39, we get:
"

Vn

"
=

In

Ka + sKb

Kc + sKd

Ke + sKf

Kg + sKh

#"

Vf

(2.40)

If n

We can substitute If n = If x If to eliminate If n :


"

Vn

"
=

In

Ka + sKb

Kc + sKd

Ke + sKf

Kg + sKh

#"

Vf

"

Kc + sKd

If

If x

#
(2.41)

Kg + sKh

And substituting 2.36:

"

Vn

"
=

In

Ka + sKb

Kc + sKd

Ke + sKf

Kg + sKh

#"

sBxy

sCxy

#"

Vx
Ixf

"

#
If

Kc + sKd

(2.42)

Kg + sKh

We can substitute In = Yn Vn and re-arrange:

"

Vn

"
+ If

Vn Yn

Kc + sKd
Kg + sKh

"
=

Ka + sKb

Kc + sKd

Ke + sKf

Kg + sKh

#"

sBxy

sCxy

#"

Vx

Ixf
(2.43)

Reducing:

"

Kc + sKd

#"

Yn Kg + sKh

Vn
If

"
=

Ka + sKb

Kc + sKd

Ke + sKf

Kg + sKh

#"

sBxy

sCxy

#"

Vx

Ixf
(2.44)

Solving for Vn and If while neglecting second order terms:


"

Vn
If

1
=
Kv + sKw

"

Km + sKn

sKp

Kq + sKr

Kv + sKu

#"

Vx
Ixf

#
(2.45)

Equation 2.45 represents the relationship between currents and voltages at the beginning of
the faulted line and the voltages and currents for the remote end load under fault conditions.
Equation 2.45 broken into equation form:

17

1
[Vx (Km + sKn ) + sKp Ixf ]
Kv + sKw

(2.46)

1
[Vx (Kq + sKr ) + (Kv + sKu ) Ixf ]
Kv + sKw

(2.47)

Vn =

If =

2.2.8

Calculating the Distance to the Fault: Single Line to Ground

For a single line to ground resistive fault, the fault voltage is described as:
Vf = If Rf

(2.48)

The fault voltage and current can be broken into corresponding sequence components:
(0)

(1)

(2)

Vf + Vf + Vf
Vf
= (0)
= Rf
(1)
(2)
If
If + If + If

(2.49)

Taking the imaginary parts of both sides:

Im

(0)

Vf

(0)

(1)

+ Vf

(1)

(2)

+ Vf

(2)

If + If + If

=0

(2.50)

Referring back to Equation 2.37 and Equation 2.47, these equations can also be broken into
sequence components:
(0)

(0)
= Vx(0) sBxy
Ixf

(1)

(1)
= Vx(1) sBxy
Ixf

(2)

(2)
= Vx(2) sBxy
Ixf

Vf
Vf
Vf
(0)

If

(1)

If

(2)

If

1
(0)

(2)

(2.53)

(1)

h

 

i
(1)
Vx(1) Kq(1) + sKr(1) + Kv(1) + sKu(1) Ixf

(2.55)

(2)

h

 

i
(2)
Vx(2) Kq(2) + sKr(2) + Kv(2) + sKu(2) Ixf

(2.56)

1
(2)

(2.52)

(2.54)

Kv + sKw
=

(1)

(0)

1
(1)

(2.51)

h

 

i
(0)
Vx(0) Kq(0) + sKr(0) + Kv(0) + sKu(0) Ixf

Kv + sKw
=

(0)

Kv + sKw

After substituting Equations 2.54-2.57 and Equations 2.51-2.53 into Equation 2.49 while
neglecting higher order terms we obtain the solution for the distance to the fault:

18

s=

2.2.9

KAR KCI KAI KCR


(KCR KBI KCI KBR ) + (KDR KAI KDI KAR )

(2.57)

Assessment of the Das Algorithm: Advantages and Disadvantages

The method proposed by Ratan Das, is a excellent fault location method with many advantages over other competing algorithms. In this section, a discussion of the advantages and
disadvantages of the Das algorithm are compared to other apparent impedance techniques used
in distribution fault location.
Fault Resistance
An excellent attribute to the Das algorithm is its ability to compensate for fault resistance. The
author choose a 23 mile long, 25kV radial feeder to preform tests. The tests performed were
single line-to-ground faults with fault resistances varying from 5 to 50. The Das algorithm
shows that for a SLG fault of 5, the error is less than 1.7%. For a 50 fault, the error was
shown to be less than 2.2% [1].
However, it can be easily shown that the Das algorithm does not work for a bolted fault(Rf = 0)
using Equation 2.49. Das does note this drawback: ...it is practically impossible to have a
fault with exactly zero resistance.[25]
Load Compensation
One of the major problems of distribution fault location is compensating for the effects of loads.
The Das algorithm does this by developing a voltage dependent load model that is calculated
under pre-fault conditions. In order to develop the load model, information about the load is
taken from a load database. This implies that the utility using the fault locator must have
monitoring equipment at the load or available load flow study data. The drawback of using a
load flow table is common for fault locators without communication systems such as SCADA
to report real time load data.
In [8], it was shown that inaccurate load studies can lead to significant errors in fault
location with the Das algorithm. In Figure 2.8, a radial distribution feeder was modelled with
constant power loads. The load were assumed to be constant impedance loads by the fault
locator, resulting in large errors. The author of [8] concluded that the performance of the Das
algorithm is not guaranteed without load behaviour studies preformed.

19

Figure 2.8: Fault locator load model as constant impedance with system load model as constant
power [8].

Estimation of the Faulted Line using the Loop Reactance Method


One of the notable drawbacks of this method is that a estimation of the faulted line section is
obtained via the Loop Reactance method. The loop reactance is summarized by Equation 2.9.
Under no or light load conditions, the loop reactance can be used to estimate the faulted line
section with acceptable accuracy. Although for heavily loaded feeders, the loop reactance can
provide inaccurate results for estimation of the faulted line section. In [9], for heavily loaded
feeders the error between actual reactance to the fault and estimated reactance to the fault were
greater than 73%. Under light load conditions the reactance method performed much better
with a error of 28%. Under no load conditions the error was 2%.Therefore we can conclude
the under lightly loaded conditions, the loop reactance method may provide acceptable results.
However on heavily loaded feeders this can lead to a incorrect preliminary estimation of the
faulted line section.
Localizing Multiple Fault Possibilities
One of the major drawback of impedance based methods is multiple possible locations of the
fault and the Das algorithm is no exception. Das recommends the use of FCIs(Faulted Circuit
Indicators) to isolate the correct location of the fault from a list of possibilities. However, given
the size and complexity of large distribution feeders this may not be an acceptable solution
1

most notably where the load current is a substantial part of the fault current at the remote end of a long
feeder

20

for a utility. Also, careful attention must be given to the placement of the FCIs in order to
maximize their effectiveness.

2.3

Girgis Apparent Impedance Method

The Girgis Apparent Impedance method was developed by Adly Girgis and uses symmetrical components to determine a distance to the fault from the measurement point(usually the
substation)[26]. The proposed algorithm addresses some of the issues with fault location such
as fault impedance, and fault localization. However, this algorithm does not address issues such
as: unbalanced mutual coupling and non-homogeneous lines.

2.3.1

Overview of Girgis Method

The Girgis method uses fundamental frequency voltage and current phasors available at the
substation bus to determine the distance to the fault. Before the distance to the fault can
be calculated, the fault type is needed: Line-to-Ground, Line-to-Line, Line-to-Line-to-Ground,
or 3-phase. To classify the fault type, changes in current magnitude are observed, indicating
the faulted phase(s). Once the fault type has been determined, the algorithm uses symmetrical
components to determine the distance to the fault. If the solution yields multiple fault locations
the algorithm will use localization techniques to determine the most likely candidate. Therefore,
we can break the Girgis algorithm into three easy steps: Fault Classification, Solution and
Localization.

2.3.2

Direct Determination of Distance to the Fault

To begin our analysis of the Grigis algorithm we consider a simple feeder with no laterals or
taps in Figure 2.9. A single line-to-ground fault is placed d distance from the measurement
bus. We will begin our analysis be assuming the system to be unloaded, and the only current
flowing at the time of the fault is due to the fault itself.

(0)

If a

If a

(1) 1

If a = 1 2 0
3
(2)
1 2
0
If a

(2.58)

Evaluating Equation 2.58 above yields:


(0)

(1)

(2)

If a = If a = If a =

If a
3
(1)

(2.59)
(2)

(0)

The above equation directly implies that the positive(If a ), negative(If a ) and zero sequence(If a )

21

currents are equal during the fault. Using this implication, we must connect the positive, negative and zero sequence networks in series as shown in Figure 2.10. The positive, negative
(1)

(2)

(0)

and zero sequence impedance of the line are represented in the figure as ZL , ZL , and ZL
(i)

respectively. The source voltage is also broken into sequence components Vs .

Figure 2.9: Simple Feeder with no laterals or taps

To calculate the distance d to the fault, we begin with a simple KVL loop around the
positive, negative and zero sequence parts of Figure 2.10 to solve for sequence voltage at the
fault:
(1)

= Vs(1) DIf a ZL

(2)

= Vs(2) DIf a ZL

(0)

= Vs(0) DIf a ZL

Vf

Vf

Vf
(0)

Vf

(1)

+ Vf

(2)

+ Vf

(1)

(1)

(2.60)

(2)

(2)

(2.61)

(0)

(0)

(2.62)

(0)

(2.63)

= 3If a Rf

Summing equations 2.60, 2.61 and 2.62.

(0)

Vf

(1)

+ Vf

(2)

+ Vf

h
i


(1) (1)
(2) (2)
(0) (0)
= Vs(0) + Vs(1) + Vs(2) D If a ZL + If a ZL + If a ZL
(0)

We can substitute: Vf

(1)

+ Vf

(2)

+ Vf

(0)

= Vf and Vs

(1)

+ Vs

(2)

+ Vs



(1) (1)
(2) (2)
(0) (0)
Vf = Vs D If a ZL + If a ZL + If a ZL

(2)

(2.64)

= Vs .
(2.65)

(1)

The Girgis method assumes that ZL = ZL .



h
i

(1)
(1)
(2)
(0) (0)
Vf = Vs D ZL If a + If a +If a ZL

22

(2.66)

Figure 2.10: Single line-to-ground fault sequence networks.

Rearranging yeilds:
(1)
DZL

Vf = Vs
(0)

(1)

(1)
If a

(0)
(0) ZL
If a (1)
ZL

(0)
(0) ZL
If a (1)
ZL

(2)
If a

If a

(0)
If a

If a +

(0)
(1)
Z
(0) Z
If a L (1) L
ZL

(2.67)

(2)

Substituting If a If a = If a + If a :
Vf = Vs

(1)
DZL

Vf = Vs

(1)
DZL

(2.68)

If we simplify:

Solving for Vs :

23

(2.69)

(1)
DZL

Vs = Vf +

If a +

(0)
(1)
Z
(0) Z
If a L (1) L
ZL

!
(2.70)

Substituting Equation 2.63:


Vs =
(0)

Let k =

(0)
3If a Rf

(1)
DZL

If a +

(0)
(1)
(0) ZL ZL
If a
(1)
ZL

!
(2.71)

(1)

ZL ZL
(1)
ZL

(0)

(1)

Vs = 3If a Rf + DZL

(0)

If a + If a k

(2.72)

(0)

Equation 2.72 can be divided by If a + If a k which gives:


(0)

Vs
If a +

(0)
If a k

3If a Rf

If a +

(1)

(0)
If a k

+ DZL

(2.73)

The above equation can be used to calculate the positive sequence reactance to the fault.
This is accomplished by taking the imaginary part of both sides of Equation 2.73. Assuming
(0)

(0)

If and If a + If a k are in phase, the imaginary part of

Im

3If a Rf
(0)

If a +If a k

would equal 0. This yeilds:

Vs
If a +

(0)
If a k

= DX (1)
L

(2.74)

The above solution is called the positive sequence reactance method[9]. The obvious drawback
(0)

of this solution is the assumption that If and If a + If a k are in phase. The Girgis method offers
a different solution which does not make this assumption.
To begin, we define the apparent impedance seen at the measurement bus as Zapp =

Vs
(0) .
If a +If a k

(0)

Zapp =

3If a Rf
If a +

(0)
If a k

(1)

+ DZL

(2.75)

(0)

We then substitute Icomp = 3If a , where Icomp is the compensating current fed into the
fault.
Zapp =

Icomp Rf
If a +

(0)
If a k

24

(1)

+ DZL

(2.76)

(0)

Next, the compensating current Icomp and the selected current If a + If a k are broken into
real and imaginary parts:
Icomp = Id + jIq

(2.77)

(0)

If a + If a k = Is1 + jIs2

(2.78)

(1)

The positive sequence impedance of the line, ZL is broken into real and imaginary parts
as well:
(1)

(1)

(1)

ZL = RL + jXL

(2.79)

Substituting 2.77,2.78, and 2.79 into 2.76:



 R (I + jI )
q
f
d
(1)
(1)
Zapp = D RL + jXL +
Is1 + jIs2

(2.80)

In the above equation, the fault resistance and distance to the fault are unknown. To solve
for the distance to the fault, the apparent impedance is broken in real and imaginary parts.
This yields two equations and two unknowns.
(1)

Re (Zapp ) = DRL + Rf

Id Is1 + Iq Is2
I2 + I2
| s1 {z s2 }

(2.81)

(1)

Im (Xapp ) = DXL + Rf

Iq Is1 Id Is2
I2 + I2
| s1 {z s2 }

(2.82)

Substituting M and N for the terms above:


(1)

(2.83)

(1)

(2.84)

Re (Zapp ) = DRL + N Rf

Im (Xapp ) = DXL + M Rf
Solving for D:
D=

Rapp M Xapp N
(1)

(1)

(2.85)

RL M XL N
The above equation represent the direct determination of the distance to the fault from the
measurement point.

25

2.3.3

Assessment of the Girgis Algorithm: Advantages and Disadvantages

The Girgis algorithm is a very simple algorithm that can be easily implemented by a utility for
basic fault locating. However, the method does have many negative attributes that make it an
unacceptable choice for many modern utilities. The following conditions were not considered by
the Girgis algorithm: non-homogenous lines, mutual coupling and system loading conditions.
System Loading
The Girgis algorithm considers the feeder to be unloaded at the time of the fault occurrence.
This assumption causes large errors in heavily loaded feeders. In the event that the fault were
to occur at the remote end of a long feeder, it is often the case that the current seen at the
substation during the fault is not much greater than the load current. In this case, the load
current presents a major issue resulting in degraded accuracy at the remote end of the network.
Non-Homogeneous Lines and Mutual Coupling
The direct determination of the distance to the fault assumes equal mutual coupling and self
impedances of the line. In the event that the we have equal self impedance and unequal mutual
impedance or visa-versa this results in coupling between sequence components1 . In practice,
equal self impedance and mutual coupling terms are rarely the case.
The Girgis algorithm also assumes the feeder conductors to be homogeneous. Many feeders
are composed of many different types and sizes of conductor, resulting in a non-homogeneous
system. This is a major disadvantage of this algorithm.
Localization of Multiple Fault Possibilities
The Girgis algorithm, much like many impedance based algorithms can return multiple fault
possibilities. For example, if the fault is found to be 1 mile from the measurement point,
there may be multiple locations that are 1 mile from the measurement point. The author does
recognize this as a limitation and presents a excellent solution. When multiple fault locations
are found, the operating characteristics of protective devices are used to eliminate possibilities.
This will be discussed in a later chapter.
1

The 3-phase system is no longer decoupled.

26

2.4

Fault Locating using Digital Fault Recorder Data(Saha Algorithm)

2.4.1

Overview of Saha Method

The algorithm proposed by Mourari Saha uses fundamental voltages and current available at
the substation before and after the fault[1]. The algorithm proposed by Saha can be broken
into two major steps: calculation of the fault loop impedance and calculation of impedance
along the feeder. By comparing the measured impedance with the calculated feeder impedance,
an indication of the fault location can be obtained[1].

2.4.2

Fault Loop Impedance Determination

The algorithm suggested by Saha requires that the positive sequence fault loop impedance be
calculated from the available voltages and currents at the substation. This impedance is later
used to determine the faulty node in the network. To begin our analysis we consider a Phaseto-Phase fault involving phases B and C at some node in the network. A phase-to-phase fault
is shown in Figure 2.11 at any arbitrary point in the distribution network. The parameters
If a , If b and If c are the fault currents measured at the substation.
Assuming the system to be unloaded at the time of fault we can easily show that for a Bto-C fault that If c = If b and If a = 0. We can now transform the fault currents flowing in the
network into their respective symmetrical components using the transform in Equation 2.86.

(0)

If a

(1) 1

If a = 1 2 If b
3
(2)
1 2
If b
If a

(2.86)

Reducing Equation 2.86 results in the following:


(1)

(2)

If a = If a

(2.87)

The above equation forms the foundation of solving for the positive sequence fault loop
impedance. Using Equation 2.87 we can form its circuit representation shown in Figure 2.12.
If we preform a simple KVL loop on Figure 2.12, we obtain:
(1)

(1)

(1)

(1)

(2)

(2)

(2)

Vf a If a Zkk Zf If a + If a Zkk Vf a = 0

(2.88)

(1)

Where the positive sequence measured voltage is Vf a and the negative sequence measured
(2)

voltage is Vf a . The thevanin positive and negative sequence impedances looking into the

27

(1)

(2)

system at some arbitrary bus k is defined as Zkk and Zkk .

Figure 2.11: B-to-C Fault

Figure 2.12: Sequence Network Diagram for a Phase-to-Phase Fault.

Using Equation 2.87, we can eliminate all negative sequence currents:


(1)

(1)

(1)

(1)

(1)

(2)

(2)

Vf a If a Zkk Zf If a If a Zkk Vf a = 0
If we solve Equation 2.89 for the fault loop impedance we get:

28

(2.89)

(1)

(2)

Vf a Vf a

(1)

(2)

= Zkk + Zf + Zkk

(1)
If a

(2.90)

We can break down Equation 2.90 in much simpler terms by analysing the numerator
portion of the equation. The pre-fault measured voltages can be broken into their corresponding
sequence components by using a equation similar to Equation 2.86:

Vaf

(0)

Vaf


(1)
Vbf = 1 2 Vaf
(2)
Vcf
1 2
Vaf

(2.91)

If we extract Vbf and Vcf we get the following equations:


(0)

(1)

(0)

(1)

(2)

Vbf = Vaf + Vaf 2 + Vaf


(2)

Vcf = Vaf + Vaf + Vaf 2

(2.92)
(2.93)

Subtracting Equation 2.92 from 2.93 yeilds:


(1)

Vbf Vcf = Vaf



(2)
2 + Vaf 2

(2.94)

 


(2)
3j) + Vaf
3j)

(2.95)

Substituting = 16 120 :
(1)

Vbf Vcf = Vaf


(2)

(1)

We can now solve for Vaf Vaf :


(2)

(1)

Vaf Vaf =

Vbf Vcf

3j

(2.96)

If we substitute 2.96 into 2.90 we get:


Vbf Vcf

3j
(1)
If a

(1)

(2)

= Zkk + Zf + Zkk

(2.97)

Using a the symmetrical component transform similar to Equation 2.86, we can show that
the positive sequence fault current yields:
(1)

If a =


1
If b 2 If b
3

29

(2.98)

Reducing yields:

(1)
If a

3
jIf b
3

(2.99)

Substituting Equation 2.99 into Equation 2.97:


Vbf Vcf

3j

3
3 jIf b

(1)

(2)

= Zkk + Zf + Zkk

(2.100)

Reducing Equation 2.100 yields:


Vbf Vcf
(1)
(2)
= Zkk + Zf + Zkk
If b

(2.101)
(1)

(2)

If we assume that the positive and negative sequence impedances are equal Zkk = Zkk :
Vbf Vcf
(1)
= 2Zkk + Zf
If b

(2.102)

During fault conditions 2If b = Ibf Icf :


Vbf Vcf
(1)
= Zkk + Zf
Ibf Icf

(2.103)

Assuming the fault to be bolted Zf = 0:


Vbf Vcf
(1)
= Zkk
Ibf Icf

(2.104)

The above equation represents the positive sequence impedance to the fault. We can use
the available voltages Vbf and Vcf and currents If b and If c to calculate the positive sequence
fault loop impedance. However, with the fault resistance being unknown, we must assume that
the fault is bolted Zf = 0.

2.4.3

Determination of the Faulty Node

After the positive sequence fault loop impedance has been found, we then search for the faulty
node. To begin our search for the faulty node we consider a simple substation with two feeders.
Let us assume that the only available measurements are the bus voltage and supply current
and I. The pre-fault positive sequence impedance of the fault feeder is
from the source: V
(1)

(1)

defined as Zk . The parallel connected feeder positive sequence impedance is defined as Zlk .

30

Figure 2.13: Simple Substation with Two Parallel Feeders during Pre-Fault Conditions.

During the pre-fault conditions, we can define the pre-fault positive sequence impedance
(1)

seen by the fault locator as Zpre . This is represented in equation form as:
(1)

(1)
Zpre

(1)

Zk Zlk

(1)

(2.105)

(1)

Zk + Zlk

During fault conditions, as shown in Figure 2.14, we represent the positive sequence impedance
seen by the fault locator as:
(1)

(1)
Zf

(1)

Zf k Zlk
Vb Vc
=
= (1)
(1)
Ib Ic
Zf k + Zlk

(2.106)
(1)

Our objective is to calculate the positive sequence loop impedance of the faulty feeder Zf k .
(1)

Solving Equation 2.106 for Zf k we get:


(1)

(1)
Zf k

(1)

Zf Zlk

(1)

(1)

(2.107)

Zlk Zf

(1)

If we solve Equation 2.105 for Zlk we get:


(1)

(1)

Zlk =

(1)

Zpre Zk
(1)

(1)

Zk Zpre

31

(2.108)

Figure 2.14: Simple Substation with Two Parallel Feeders during Fault conditions.

Substituting Equation 2.108 into Equation 2.107:


(1)

(1)
Zf k

(1)

Zk Zpre

(1)

(1)
Zpre

(1)

Zk Zpre

(1)
Zf

(2.109)

(1)

Zk
{z

kzk

(1)

(1)

Consequently, kzk can also be related to Zpre and Zlk :


(1)

kzk =

Zpre

(2.110)

(1)

Zlk

Substituting:
(1)

(1)

Zf k =

(1)

Zk Zpre
(1)

(1) Zpre

(1)

Zpre Zf
(1)

(1)

Zf k =

(1)

Zlk

(1)

Zk Zpre
(1)

(1)

Zpre Zf kzk

We previously defined Zf as:

32

(2.111)

(2.112)

(1)

Zf

V
Va Vb
=
Ia Ib
I

(2.113)

If we substitute Equation 2.113 into 2.117:


V

(1)

Zf k =

I kzk

(2.114)

V
(1)

Zpre

The above equation represents the positive sequence loop impedance of the faulty feeder
during fault conditions. It is assumed that the impedance of the adjacent parallel feeders do
not change during fault conditions(this may not always be the case).
Now that the positive sequence loop impedance of the faulty feeder has been calculated, we
can begin to search for the faulty line. The algorithm suggested by Saha sweeps the network
for the faulty node until a specific set of criterion exists. One the criterion has been met, the
faulty line section is found.

Figure 2.15: Cascaded Line Sections of Distribution Feeder.

We will begin our search for the faulty node at some node i1 in the network. We will represent
(1)

the impedance of the cable between i 1 and i by Zsi1 . The load at the remote node i is
(1)

represented by Zpi . By using network reduction, we can develop a relationship between the
(1)

(1)

impedance seen looking into the system at i 1 and i(Which is Zf i and Zf i1 respectively).
(1)

(1)

Zf i1 =

(1)

Zf i Zpi
(1)
Zf i

(1)

Solving for Zf i we get:

33

(1)
Zpi

(1)

+ Zsi1

(2.115)

(1)

(1)

Zf i =

Zpi

(1)

(1)

Zf i1 Zsi1

(1)

(1)


(2.116)

(1)

Zpi Zf i1 + Zsi1

Let us assume the substation bus to be i 1; assuming the load impedances to be known,
we can now traverse each line section and calculate the impedance seen looking into the system
(1)

(1)

from the remote bus. As expected |Zf i | < |Zf i1 | as we approach the faulty line. If we were
(1)

traverse down the feeder and the value of Im(Zf i ) would be become less than zero, we would
know the fault exists between i 1 and i.
To begin our search for our faulty line section, we always begin at the substation and work
outwards. In this particaluar scenario, the substation may represent i 1. As we move outward
(1)

(1)

on the feeder, the value of Im(Zf i ) decreases. When the value of Im(Zf i ) becomes negative,
the faulty line section has been found[32]. This part of the algorithm only identifies that the
line section between i 1 and i is faulty. The distance from i 1 to the fault will be covered
in the next section.

2.4.4

Distance to the Fault

After we have identified the fault line section(i and i 1) we must now determine the distance
from node i 1 to the fault(distance down the line to the fault). Using a change in notation,
we will call the faulted line k and k + 1 in the following example(k = i 1 and k + 1 = i).

Figure 2.16: Faulted Distribution Feeder[1].

Using Figure 2.16, we will place a the fault between nodes k and k + 1. Our objective is to
calculate the distance from node k to the fault.

34

(1)

Let us begin by assuming that the positive sequence loop impedance from node 1 to k, Z1k

is known from pre-calculated tabulated values. Also, we will assume that the impedance of the
(1)

shunt load at node 2 Zl2 is also known. Any impedance beyond node k is lumped together as
(1)

a shunt impedance called Zf . The circuit representation of the faulty network seen by the
fault locator is represented by Figure 2.17.

Figure 2.17: Circuit Representation of Network for a Fault located between 1 and k.

We will begin our derivation by assuming that the impedance seen by the fault locator is defined
(1)

as Z1f .
(1)

Z1f =

V
I

(2.117)

In the above figure, the variable m is used as a percentile over the total length of the cable
from node 1 to node k. In this example the load is 1 m distance from the measurement
point and m from the origination terminal of the faulted line section(node k). The equation
representation of impedance seen by the fault locator described by Figure 2.17:


(1)
Z1f

(1)

(1)

mZ1k + Zf

(1)

Zl2
V
(1)
=
=
+ (1 m) Z1k
(1)
(1)
(1)
I
mZ1k + Zf + Zl2
(1)

If we solve for the unknown variable, Zf

we get:

35

(2.118)


(1)

Zf

=
(1)

Previously we defined Zf

(1)


 

(1)
(1)
(1) 2
mZ1k + Zl2 mZ1k


(1)
(1)
(1)
(1)
mZ1k Zl2 Z1k Z1f
(1)

Z1k Z1f

(2.119)

as the lumped impedance representative of all elements beyond


(1)

k. In order to solve for the distance to the fault from k, we must define Zf

in much more

detail.

(1)

We can represent Zf

as the circuit shown in Figure 2.18. All elements beyond node k

are represented, including the fault resistance. In Figure 2.18, x represents the distance to the
(1)

fault from node k and ZL represents the positive sequence impedance of the line from node
(1)

k to k + 1. Also since a shunt load is present at node k, this is represented by Zlk . We also
must represent all elements beyond node k + 1; this positive sequence thevanin impedance is
(1)

represented by Zk+1 .

Figure 2.18: Circuit Representation of Network for a Fault located between k and k + 1.

Using network reduction we can reduce Figure 2.18 to the following equation:
(1)
Zlk
(1)
Zf

(1)
xZL

=
(1)
Zlk

(1)
xZL


!
(1)
(1)
Rf (1x)ZL +Zk+1

36

(1)

(1)

Rf +(1x)ZL +Zk+1


(1)
(1)
Rf (1x)ZL +Zk+1
(1)

(1)

Rf +(1x)ZL +Zk+1

(2.120)

If we set Equation 2.119 and Equation 2.120 equal and solving for Rf :
Rf = x2 A xB + C

(2.121)

Where A,B and C are:




 
(1) 2


(1)
(1)
Zlk Zf


A= 
(1)
(1)
(1)
(1)
(1) (1)
Zlk Zf
ZL + Zk+1 Zf Zlk

(1)

B=

ZL



ZL




(1)
(1)
(1)
(1)
(1) (1)
(1) (1)
Zlk Zf
ZL + Zk+1 Zf Zlk + 2Zf Zlk



(1)
(1)
(1)
(1)
(1) (1)
Zlk Zf
ZL + Zk+1 Zf Zlk



(1) (1)
(1)
(1)
Zf Zlk ZL + Zk+1


C=
(1)
(1)
(1)
(1)
(1) (1)
Zlk Zf
ZL + Zk+1 Zf Zlk

(2.122)

(2.123)

(2.124)

Taking the imaginary part of the quadratic function for the fault resistance Rf :
Im (Rf ) = x2 Im(A) + xIm(B) + Im(C) = 0

(2.125)

Where the imaginary parts of A,B and C are:

Im(A) = Ai

Im(B) = Bi

Im(C) = C

(2.126)

The resultant roots 1 and 2 of Equation 2.125 are:


Bi +

1 =

2Ai
Bi

x = 2 =

Bi2 4Ai Ci

q
Bi2 4Ai Ci
2Ai

37

(2.127)

(2.128)

Since 1 is considered a invalid solution, the distance to the fault from k is x = 2 .

2.4.5

Assessment of the Saha Algorithm: Advantages and Disadvantages

The algorithm suggested by Saha has many advantages and addresses many of the issues that
other algorithms fail to consider. The algorithm compensates for system loading, fault resistance, non-homogeneous feeder conductors and measurement issues(only available current is the
main transformer current). However, there are several key disadvantages of the Saha algorithm.
One of which is that the positive sequence fault loop impedance is used to calculate the faulty
line section assuming unloaded conditions and a bolted fault. The algorithm also presents a
direct determination for distance to the fault, but this solution is highly dependent on topology
of the system. Also, the author does not address localization of multiple fault possibilities if
(1)

there exists two possible paths on a feeder where Im(Zf i ) < 0.


Problems with Determining the Faulty Line Section
The Saha algorithm uses the positive sequence fault loop impedance to directly determine the
faulty line section. To achieve this, the algorithm must assume the fault to be bolted and
the system to unloaded at the time of fault. For systems that are lightly loaded, this is an
acceptable way of calculating the faulty line section. However, for heavily loaded feeder, the
load currents become a large portion of the fault currents making this assumption possibly
invalid.
Problems with Determining Distance to the Fault
Once the faulty line section has been determined, when then calculate the distance to the fault
from the origination node of the faulty line section(see Figure 2.18). However, the solution
presented in Equation 2.128 is dependent on toplogy. To illustrate this, a tap load could be
inserted between nodes 2 and k, as shown in Figure 2.19.
Inserting a tapped load between nodes 2 and k complicates the direct determination of the
positive sequence fault loop impedance shown in Equation 2.118. Therefore, we can conclude
that the direct determination of the distance to the fault in Equation 2.128 is a rigid solution
that is dependent on network topology. This issue was not addressed by the author.
Localization of Multiple Fault Possibilities
As with many impedance based algorithms, there are usually multiple possible locations of
the fault. When using the positive sequence fault loop impedance it is often the case that
(1)

there is more than one location to satisfy the criterion for a faulty line section(Im(Zf i ) < 0).
Localization or ranking of the possibilities was not considered by the author.

38

Figure 2.19: Tapped Load inserted between node 2 and node k.

2.5

Conclusion

The algorithms covered in this section use fundamental voltages and currents available at the
substation to calculate the location of the fault. Impedance based algorithms, like the ones
presented here often have the same key disadvantages: localization of multiple fault possibilities
and error introduced by system loading.
It is often the case that some uncertainty exists when modelling load. Many impedance based
algorithms rely on accurate load information throughout the network in order to calculate the
location of the fault. However, the dynamic nature of system load is very difficult to predict
and can introduce large errors in fault locating. If a reasonable amount of information is know
about the system loading, this is usually sufficient for most algorithms to produce acceptable
fault locating accuracies([8],[1]).
One of the major flaws of the impedance based algorithm is localizing multiple fault possibilities. It is often the case that there exists multiple locations where the Thevanin impedances
are very similar resulting in two or more solutions. In order to rank them, Girgis suggests using
the characteristics of the protective devices that interrupted the fault to localize the fault to
one solution. This is a simple and effective way of localizing multiple fault possibilities.

39

Chapter 3

Model Based Fault Locating


3.1

Introduction

In this section we propose a model based FLA(Fault Locating Algorithm) that uses recorded
fault data available at the substation bus. The FLA identifies all possibly fault nodes by
comparing model derived short circuit data to the observed fault current. In the event that
there are multiple possibilities, the FLA executes a localization sub-algorithm. The localization
algorithm uses load flow data and protective device characteristics(operating time) to rank all
possibilities. After all possibilities are ranked, the results are printed out in the MATLAB
console for the user.
Figure 3.1 shows a high level overview of the proposed FLA. The FLA can be broken into
several key steps:
Simulation
Sampled Data Conditioning
Steady State Fault Current Extraction
Fault Type Identification
Identify Possibly Faulted Nodes
Localization
Ranking
Printout

40

Figure 3.1: High-Level Overview of FLA.

3.2

Sampled Data Conditioning

Many modern relays and IEDs have the capability to record pre-fault and post-fault voltage
and current waveforms at very high resolutions. Most modern protective relays can capture
resolutions up to 64 samples per cycle or higher [11](depending on the vendor). One of the
excellent attributes of the proposed method is that it does not require high resolution sampling.
Testing performed in MATLAB indicated that 16 samples per cycle provides sufficient resolution
for the proposed method. Since most modern protective relays and DFRs have the capability
of 64 samples per cycle or more, all testing done in this paper was performed at 64 samples per
cycle. For example, the SEL(Schweitzer Engineering Labs) 451 series relay has the capability
to sample up to 133 samples per cycle. Manufacturers of protective relays and DFRs have
incorporated the ability to share raw sampled fault data in many of their software packages.
An excellent example of this is the SEL-5601 software package that allows the user view captured

41

Figure 3.2: Handling of Raw Data Passed to the Fault Locator.

oscillography from SEL relays. The method suggested here assumes that the data it receives is
unfiltered fault data(provided by MATLAB Simulink). A DFT(Discrete Fourier Transform) is
later used to analyse the fault data to extract the 60Hz fundamental component for analysis.
The complete FLA process of conditioning the sampled data is shown in Figure 3.2.

Figure 3.3: Sampling of Fault Current Recorded at 64 Samples Per Cycle

Figure 3.4 is a excellent example of harmonic content present during a fault condition. The
FLA method proposed uses fundamental frequency phasor quantities to perform all calculations. To remove these harmonics, we use the DFT algorithm built in MATLAB to extract

42

the magnitude and angle of the fundamental. Fourier analysis shows that periodic functions
can be reproduced as a combination of complex exponentials whose frequencies are multiples
of the fundamental frequency kf0 where f0 = 60Hz. We are particularly interested in the
fundamental k = 1. The DFT transform used by MATLAB is shown in Equation 3.1 [12].

Figure 3.4: Bus Voltage During a Ground Fault.

X (k) =

N
1
X

x(n)e2j N n

(3.1)

n=0

The magnitude output of the DFT |X(k)| can be directly divided by the

2 to obtain the

RMS value.
XRM S =

3.3

|X (k) |

(3.2)

Steady State Fault Current Extraction

The FLA suggested here uses phasor quantities to preform calculations. During the occurrence
of a fault, the current will go through a short transient period before reaching steady state

43

value(Shown in Figure 3.5). The fault locator must wait until the observed current reaches
steady state. Once steady state has been reached, the RMS fault current value is captured for
later use.

Figure 3.5: RMS value of the 60Hz Fundamental during Line-to-Ground Fault.

To calculate the steady state value of the observed current during a fault, a algorithm is
used that was suggested in [13] . At the nth cycle, the relative change of the current between
the nth + 1 cycle and the nth cycle is less than .1, then the second condition is checked. If the
relative change between the nth + 2 cycle and the nth cycle is less than .1, then the current is
considered to be in steady state in the nth cycle. The steady state current is considered to be
the average of the nth + 1 and nth cycle. The full process of detecting the steady state of the
observed fault current is shown in Figure 3.6.

3.4

Fault Tables

The location of the fault in the network is determined using fault tables. Fault tables are the
expected value of fault current for any given location of the fault in the network. The actual
measured fault currents are compared to the expected(fault tables) to determine the location of
the fault. The idea of the algorithm relies on the diversity of thevanin impedances throughout
the network. Given that each point throughout the network yields a unique fault current value,
the fault location is easily found.
To illustrate the formation of a fault table, we consider a generic 6 node feeder shown in

44

Figure 3.6: Algorithm for detecting steady state fault current.

Figure 3.7. To form the fault table, we apply a 3-Phase(Bolted) fault beginning at node A and
concluding at node F. Each time the fault is moved throughout the feeder, the fault current at
the substation is recorded. This forms the fault table shown in Table 3.1. This table represents
the expected values of fault current if a 3-Phase fault did occur at Node A-Node F. To locate
a fault, we simply compare our actual and expected values of fault current. For example, a
3-Phase fault occurs in our actual system and results in 7158 A of current flow. We would be
able to rule out nodes A-C and nodes E-F due to their high/low expected values. The most
likely location for the fault to have occurred is node D.

Figure 3.7: Generic 6 Node Feeder.

45

Table 3.1: Fault Table for Figure 3.7


Node
Node
Node
Node
Node
Node
Node

3.4.1

Fault Current If
A
B
C
D
E
F

9553
8201
7542
7118
6346
5690

A
A
A
A
A
A

Sliding Fault Resolution

In our previous 6 node network shown in Figure 3.7, we applied a fault at each node to form the
fault table. Therefore, there can only be 6 possible places that the fault can occur(according
to the table). It is a possibility that our 3-phase fault could occur between nodes A and B. Let
us re-form the fault tables in Table 3.1 to include a fault between each node shown in Table
3.2. We now have a much higher resolution fault table.
If we revisit our previous example of a fault occurrence resulting in 7158 A, we now have a
much different problem. If the fault were to occur at node D, we should see 7118 A of expected
current. It is evident that the most likely location is still node D, but other possibilities
must be considered. With the higher resolution table, a fault location between node C and
node D(7330 A expected value) or between node C and node E(6944 A expected value) are
now likely possibilities. Although we have increased our resolution and accuracy, we have
increased the number of possible fault locations. The lower resolution fault tables result in a
larger current diversity, which yields less possibilities. In later sections we will introduce the
concept of localization techniques. Localization techniques allow us the ability to choose high
resolution fault tables while being able to eliminate certain possibilities. This will be discussed
on later sections.
When choosing a fault table, it is important to consider the units of resolution. In Test
Case 4, a sliding fault is placed at each distribution pole. This will achieve two objectives:
accuracy and clarity. In a distribution system, poles are often located a short distance apart
usually within line of sight. Line crews will be able to exactly locate the fault by the pole
name or location(this achieves clarity of information). Using pole-to-pole sliding faults do not
offer significant current diversity(between 50-100A), however this will allow the fault locator to
select the exact pole that is faulted. For example, the output of the FLA will indicate that the
fault is located at Pole 5567 3 poles with Pole 5567 being the most likely.

46

Table 3.2: Increased Resolution Fault Table for Figure 3.7


Node
Node A
Node B
Between
Between
Node C
Node D
Between
Between
Node E
Node F
Between

3.4.2

Fault Current If

Node A/B
Node A/C

Node C/D
Node C/E

Node E/F

9553
8201
8877
8547
7542
7118
7330
6944
6346
5690
6018

A
A
A
A
A
A
A
A
A
A
A

Selecting Possible Fault Locations from Fault Tables

In a ideal model, the expected and actual currents would be identical. However, measurement,
modelling and other errors introduce differing values than expected. Therefore, a window of
3% was included to account for all associated error. Selecting possible fault locations from
the fault table is done by collection all values that are within 3% of the actual value. Each
fault table entry(Iexpected ) is compared to the actual value of the fault current to determine if
the location is considered to be faulted.
In our previous example, a 3-Phase fault produced a current of 7158 A measured at the
substation. Using Table 3.2, each fault table entry is evaluated to determine if the associated
location is faulted. According to the criterion all locations where the current is greater than
6943.26 A and less than 7332.74 A is considered a possible location.
The 3% window for selecting possible fault locations can be expanded or contracted depending on modelling confidence. If the window is expanded, this will result in much more
possible fault locations when traversing the fault tables. However, this will allow for more measurement and modelling uncertainties. If the window is contracted, this opens the possibility
that the fault location may not be found.

3.4.3

Fault Identification and Fault Table Selection

When a fault occurs in the system, the fault locator must identify the type of fault to determine
the correct choice of fault table. For example, if a line-to-line fault occurs in our network, we
will select the fault table for line-to-line faults. For a line-to-ground fault, we would select our

47

line-to-ground table. When identifying the fault type, we must also record the affected phases.
For example, if a phase-to-phase fault occurs, we would need to know that the fault involves
phases B and C.
Each type of fault(except 3-phase) will contain three fault tables:
Line-to-Ground Fault: A-G, B-G, C-G
Line-to-Line-Ground Fault: A-B-G, B-C-G, and C-A-G
Line-to-Line Fault: A-B, B-C, C-A
3-Phase: A-B-C
To classify the fault type, we set the fault locator to pickup similar to a standard overcurrent relay. A standard phase over-current relay is usually set to 1.5If ullload . A standard
ground over-current element is usually set above the maximum known 3I (0) residual imbalance.
This philosophy was used as a threshold current or pick-up to detect the fault type and affected
phases. The values IAT ,IBT ,ICT and IN T represent the user defined threshold values.
Identification of Phase-to-Phase Faults
Phase-to-Phase faults contain positive sequence I (1) and negative sequence I (2) currents only.
With the absence of zero-sequence currents we can detect a Phase-to-Phase faults using the
following criterion:
A-B Fault:|If a | > IAT and |If b | > IBT and |If c | < ICT and |In | < IN T
Identification of Phase-to-Phase-to-Ground Faults
Phase-to-Phase-to-Ground faults contain positive sequence I (1) , negative sequence I (2) and
zero sequence I (0) currents. Using this, we can detect Phase-to-Phase-to-Ground faults in a
similar fashion as Phase-to-Phase Faults.
A-B-G Fault:|If a | > IAT and |If b | > IBT and |If c | < ICT and |In | > IN T
Identification of Phase-to-Ground Faults
Phase-to-Ground faults contain positive sequence I (1) , negative sequence I (2) and zero sequence
I (0) currents. Using this, we can detect Phase-to-Ground faults in a similar fashion as Phaseto-Phase Faults.
A-G Fault:|If a | > IAT and |If b | < IBT and |If c | < ICT and |In | > IN T

48

Identification of 3-Phase Faults


Three phase faults contain positive sequence I (1) currents only(assuming a balanced 3-Phase
fault). The absence of negative and zero sequence currents indicate a three phase fault.
A-B-C Fault:|If a | > IAT and |If b | > IBT and |If c | > ICT and |In | < IN T
Evolving Faults
Evolving faults begin as one type of fault and end as another. Faults can begin involving one
phase and then flash-over to another phase. For example a fault may begin as a phase-to-phase
fault involving B and C phases; however, after a short amount of time, the arcing between B
and C may flash-over to A. This results in a fault beginning as a Phase-to-Phase fault and then
ending as a 3-phase fault.
The FLA here does not consider evolving faults. A simple MATLAB script can be included
in the FLA to detect an evolving fault. If a evolving fault does occur, this does not imply that
the fault locator cannot calculate the fault location. Figure 3.8 shows a Phase-to-Phase fault
involving Phases B and C and then transitioning into a 3-Phase fault. Since the final fault is
3-Phase, the fault locator can ignore the Phase-to-Phase sampled data, and consider the fault
to be 3-Phase. If the current has reached steady state before the downstream protective device
interrupts, the fault can assumed to be 3-phase.

3.4.4

Calculation of Fault Currents in the Fault Table: Fault Resistance

The calculation of the fault currents in the fault table are done via the process outlined in
Section 3.4. The system is considered to be under loaded conditions during the fault, and
all load currents are considered in the solution. However, the arcing resistance of the fault
is a unknown when calculating the fault current. According to EPRI studies, a bolted fault
assumption is an acceptable assumption for fault locating purposes([17],[7],[15]). In this paper,
all studies are done with a bolted fault assumption Rf = 0. Detailed distribution fault analysis
under loaded conditions is provided in Appendix C.

3.5

Localization Techniques

In large distribution networks, there likely exists multiple locations where if a fault were to
be applied would result in the same fault current observed at the substation. If we were to
form the thevanin equivalent for all of these locations we would find that there impedances
are very similar. The FLA suggested here relies on diversity of impedances throughout the
network(mutual coupling and non-homogeneous conductor sizing) to lessen the chance of two
nodes being chosen as possibilities. In the case that there exists more than one possibility of

49

Figure 3.8: Phase-to-Phase evolving into a 3-Phase fault[13].

the location of the fault, we can use localization techniques to rank the possibilities from best
match to worst match.

3.5.1

Localization Using Protective Devices

When a fault occurs in distribution system, a protective device such as a fuse, breaker or
recloser will interrupt the fault before damage can occur. The amount of time before the protective device operates and interrupts fault current is unique to each device. The operating
time of each device is determined by its operating characteristic. Protective devices such as
fuses can be selected with many different types of operating characteristics: T speed, K
speed, ect. Breakers and reclosers can also be set using time over-current curves to achieve the
desired operating characteristic. Generally speaking, most protective devices used in distribution networks have a operating time that is inversely proportional to current. Localizing using
protective devices narrows the possibilities to a specific zone of protection. If a recloser operated during the fault, then possibilities that are with a fuses zone of protection are much less
likely(assuming fuse blowing coordination). Therefore, the concept of localization using protective devices pinpoints a zone of protection that a fault occured in. Once a zone of protection
has been identified, all other possibility can be ranked as less likely.
If the fault locator knows the operating characteristics of the protective devices in the

50

network, the fault can be localized. To illustrate this, we will refer back to our feeder in Figure
3.7 with a few changes as shown in Figure 3.9 . Let us consider that a fault occurs in our network
with resulting in a magnitude of 8620 A. Using Table 3.2 there are only two possibilities between
nodes A/B and nodes A/C. Since we have two possible fault locations we must localize and
rank the possibilities.

Figure 3.9: Generic 6 Node Feeder with Fuses and Overcurrent Relay.

The time-current characteristics for the overcurrent relay and fuses used in the feeder are
shown in Figure 3.10. Using Figure 3.9 we can see that the two fault possibilities lie in differing
zones of protection. If the fault were to occur between nodes A and C, this would be interrupted
by the circuit breaker. If the fault were to occur between A and B, this would be interrupted by
the fuse. Using the time-current characteristics in Figure 3.10, we can see that the fuse operates
much faster than the overcurrent relay controlling the circuit breaker. If we see a fast clearing
of the fault, we know that the fault must have occurred in the fuses zone of protection. If we
see a slower clearing than the fuse, then the fault must have occurred in the overcurrent relays
zone of protection. Therefore, the concept of localization using protective devices pinpoints a
zone of protection that a fault occurred in.
If we draw a vertical line on Figure 3.10, we can graphically determine the operating time
of each device. For a fault of 8620 A the fuse operates in .1 seconds(total clearing time) and
the overcurrent relay operates in .2 seconds.
Let us assume that Figure 3.11 is the observed waveform that results in a steady state RMS
fault current of 8620 A. Using this observed current, we can see that the fault occurrence occurs
at .1 seconds and the fault is cleared by the protective device at approximately .2 seconds. This
data implies that the protective device observing the fault current operated in approximately
.1 seconds. This operation is much too fast for it to have occurred in the overcurrent relays
zone of protection, so the possibility of the fault being between nodes B and C is eliminated
or de-ranked. Therefore, the fault must have been interrupted by the fuse and occurred in the

51

Figure 3.10: Operating Times for a 8620 A Fault Through Current.

fuses zone of protection. This leaves the possible location of the fault being between nodes A
and B as the most probable location.

3.5.2

Zones of Protection

As we previously discussed, the fault locator uses protective devices to pinpoint a zone of
protection that a fault occurred in. By pinpointing the exact zone of protection, we can derank all other possibilities outside that zone. To identify the zone of protection, we use the
operating time observed to determine which device in the system has operated. Without zones
of protection, the fault locator would have no way of knowing each fault possibilitys upstream
protective device.
The main protection for a specific zone of protection is called primary protection. The
FLA requires that each possible fault location in the fault tables have an associated primary
protective device. Having an associated protective device allows the FLA to compare the
observed operating times with the expected operating times for each fault possibility. This
allows each fault possibility to have a associated time mismatch between observed and expected
operating times.
To illustrate the primary zones of protection for a feeder, we begin with the network shown
in Figure 3.9. To find the associated protective device, we pick a possible fault location and
traverse upstream until we arrive at the first protective device. Figure 3.12 shows the primary
zones of protection for each protective device in the network. There are a total of four zones:
three fuse zones(shown in blue) and one breaker zone(shown in red). The breakers zone of

52

Figure 3.11: Observed Current during Fault Conditions for Figure 3.9.

protection stretches from the terminals of the breaker to the line side of each fuse. The fuses
zone of protection stretches from the load side of the fuse to each remote node. In addition to
specifying the primary protective device for each entry in the fault table, the user must also
specify the type of protective device and subclass. For example, the user would program a
fuse as the protective device and 100KS as the subclass(100KS Fuse).

3.5.3

Localization using Fuse Characteristics

A fuse is a simple protective element with a metallic link that operates under fault conditions.
During fault conditions, excessive current causes the metallic link in the fuse to heat and begin
to melt. Once the metallic link begins to melt, the element will begin to separate into two parts
forming an open circuit. The amount of time required for the fuse to begin melting is referred
to as the melting time.
Fuse characteristics provided by the manufacturer have an associated minimum melting
curve. The minimum melt time for a fuse indicates that fault current has caused the metallic
link to melt and begin arcing within the fuse. When the metallic link has melted, an arc is
formed across the severed link. As a result, the arc provides a path for current to flow. In most
conventional fuses it is necessary to wait until zero-crossing until the arc is extinguished. As
a result, the time-current characteristics for fuses include a minimum melt tmelt and a arcing

53

Figure 3.12: Zones of Protection for Figure 3.9.

time tarc . After the arc has been extinguished, the total clearing time has been achieved. The
total clearing time of a fuse is the sum of the minimum melt and arcing time. For fault locating
purposes, the total clearing time time-current characteristic is used to predict operating times
of fuses.
ttct = tarc + tmelt

(3.3)

Fuse manufacturers, such as SandC, provide time-current characteristics for coordinating


purposes. This characteristic is usually provided in spreadsheet form, which can be plugged
into the FLA. The fuse time-current characteristics allow the FLA to calculate the estimated
operating time ttct . For more information on how the esitmated fuse observed fault current is
calculated see Section 3.5.4.

Figure 3.13: FLA Fuse Localization Algorithm.

To determine which fuse has operated in a network, the FLA compares our observed op-

54

erating time(measured by the fault locator) to the expected operating time(calculated by the
fault locator). The observed operating time of the device is determined by recording the point
in time at which the current magnitude crosses the 1.5If ullload value. When the current is
increasing and passes the 1.5If ullload threshold, this is considered the beginning of the fault.
When the protective device operates and the current drops below the threshold of 1.5If ullload ,
this time is recorded as the end of the fault. Subtracting the yields the total fault duration and
operating time top .
For each fault possibilitys upstream protective device, we calculate the time mismatch
between the operating time observed top and the calculated operating time ttct :
te = ttct top

(3.4)

The time mismatch equation allows the fault locator to calculate the difference between
the expected and observed value. This quantifies how closely each fault possibilitys upstream
protective device matches the actual observed conditions. If the time mismatch between expected
and observed is large, then the associated device is not likely to have interrupted the fault. If
the time mismatch is small, then the protective device associated with the fault possibility is
the most likely to have interrupted the fault. A high level overview of the FLA fuse localization
algorithm is given in Figure 3.13.
Ranking the Best Matched Fuse
After the FLA has calculated the mismatch te for all applicable fuses, it must choose the best
matched fuse. This will indicate the most likely zone of protection the fault occurred in. To
choose the best matched fuse, the FLA uses unity based min-max normalization. When the
vector of mismatches is normalized, all time mismatch values will take on a value between 0
and 1. After being normalized, the minimum time mismatch is normalized to 0, representing
the most likely candidate. The largest mismatch is normalized to 1.0, representing the most
unlikely candidate. All other values are appropriately scaled between 0 and 1. To illustrate
this, a test was run on a feeder with two fuses with different operating characteristics. The
time mismatch was calculated as the following:
te =

0.6557

0.0357

(3.5)

To normalize the mismatch vector te we use the following relation:


de =
t

xi min(te )
max(te ) min(te )

55

(3.6)

de represents the normalized time mismatch vector. Each data point within
The variable t
te vector is represented using xi and the min and max of the te vector is represented by
min(te ) and max(te ) respectively. After normalization, the normalized time mismatch
de is the following:
vector t
de =
t

(3.7)

The first element in the vector was normalized to 1.0 is the most unlikely candidate with
largest time mismatch. The last element in the vector was normalized to 0.0 which represents
smallest time mismatch and most likely candidate.

3.5.4

Estimating Fault Current Through the Fuse

During the fault, the only available measurements are at the substation. Due to loading, the
fault current observed at the substation and the fault current observed by the fuse are not
equivalent. Urban load can be generally classified as approximately 50% constant impedance
and 50% constant power, which behaves as a constant current load [16]. Modelling all loads as
constant current loads is considered a good approximation for most networks [17],[18].
If we assume constant current loads and the load currents do not change during fault
conditions([6]),[22]), the load observed at the substation is a superposition of the current flowing
due to the fault itself If and the steady state load currents Iload . These superimposed currents
are shown in Figure 3.14. A detailed explanation of fault analysis under loaded conditions is
covered in Appendix C. Load models presented in Appendix C describe load behaviour under
normal conditions. A piecewise load model is used to describe load behaviour under faulted
conditions.

Figure 3.14: Superimposed Currents Through Faulted Fuse.

56

Isub = If + Iload

(3.8)

The currents observed by the fuse also comprise a superposition of currents. The fault
current observed by the fuse is comprised of the fault current due to the fault itself and the
steady state load currents beyond the fuse Iload0 . If we consider only the current flowing due
to the fault itself, the fault locator at the substation and the downstream fuse are observing
the same current If . However, when the load currents are superimposed, the fault locator will
observe a larger fault current than the fuse. Therfore, the FLA estimates the fault current
observed by the fuse to be a superposition of the current flowing due to the fault itself If and
the steady state load currents beyond its terminals Iload0 .
If use = If + Iload0

(3.9)

To estimate the operating time of the downstream fuse, we are first required to approximate
the observed current during fault conditions. Since the downstream fuse and fault locator are
observing the same current flowing due to the fault itself, we begin with Equation 3.8. The
load current flowing just before the fault at t = 0 is defined as Iload and is subtracted from
the steady state observed substation fault current:
If =

If + Iload
| {z }

Iload
|{z}

(3.10)

Observed Pre-Fault Current

Observed Fault Current

Since the load current at the fuse just before the occurrence of the fault is not know, it must
be estimated from stored load flow data. We will utilize the calculated fault current computed
above(Equation 3.10) due to the fault itself to estimate the current observed by the fuse during
fault conditions. Using the database load current value, we can estimate the observed fault
current by summing the phasor quantities of the fault current due to the fault itself If and the
expected database load Iload0 1 :
If use =

If
|{z}

Iload0
| {z }

(3.11)

Expected Pre-Fault Current

Fault Current

The variable Iload0d can be generalized in a summation of all n loads tapped beyond the
terminals of the fuse:
If use = If +

n
X

Iload0d

(3.12)

d=1

In Equation 3.13, the load connected to the applicable phase beyond the fuse is represented
1

It likely that the actual fuse load current just before the fault occurrence and the expected current flow
Iload0 are different. The effects of Load Perturbations is explored in Chapter 4.8.1.

57

a
b
c
as a summation of individual tapped loads Iload0d
, Iload0d
, and Iload0d
. The fault currents

due to the fault itself on each phase is represented by If a ,If b and If c . For example, for
a Line-to-Ground fault involving Phase B, the currents flowing through the fuse calculated by
the FLA would be:

Pn


If use = If b +
0

a
Iload0d
Pd=1
n
b
Iload0d
Pd=1
n
c
d=1 Iload0d

Pn

a
d=1 Iload0d
Pn
b
If b + d=1 Iload0d
Pn
c
d=1 Iload0d

(3.13)

The full process of estimating the fault current through the fuse is shown in Figure 3.15.

Figure 3.15: Fault Current Estimation Algorithm in FLA.

3.6

Localization using Load-Flow Rejection

When a distribution feeder lacks protective device diversity, load-flow characteristics must be
used to localize the fault. Load flow rejection allows the FLA to identify the protective device
that operated even if they are identical. For example, if there are two fault possibilities and
both are in the zone of 80KS fuses, their operating times cannot be used to determine which
fuse has operated. This scenario is not limited to fuses. If there exists two or more reclosers
in the feeder that are set identically, it is not possible to localize using operating times. To
alleviate this particular problem, load-flow rejection is used to localize two protective devices
that have identical operating times.
When a fuse operates for a fault in its zone, the loads connected to that fuse are open
circuited. The load that was previously connected to the feeder is no longer present and results

58

in a load loss(rejection) observed at the substation. If a load-flow data table is formed, the
observed load loss at the substation can be compared to the table to determine which device
has operated.

3.6.1

Calculating Best Matched Device from Load Flow Rejection

To begin localizing using Load Flow Rejection, the FLA will calculate the observed load rejection
at the substation(Equation 3.14). This calculation is done in the fault locator by observing
several cycles before the current first exceeds 1.5If ullload and several cycles after the last drop
below 1.5If ullload . Several cycles before the first current exceeds 1.5If ullload is defined as the
pre-fault current Ipre . Several cycles after the last drop below 1.5If ullload is defined as the post
fault load Ipost . As mentioned the FLA saves all times where the current crosses 1.5If ullload .
Pre-fault(Ipre ) and Post-Fault(Ipost ) values are shown in Figure 3.16.
Iob = Ipre Ipost

(3.14)

Figure 3.16: Current Rejection after Recloser Lockout.

Since the observed load rejection is now known, the FLA will calculate the current mismatch

59

for each possibilitiess upstream device. The mismatch is defined as the difference between the
expected load rejection for each device and the observed load rejection. The expected load
rejection for each device comes from the load flow data table stored within the fault locator.
To calculate the current mismatch of each protective device Ie , we use the following relation:
Ie = Iob Iexp

(3.15)

An overview of the process of load flow rejection localization is shown in Figure 3.17.

Figure 3.17: Load Flow Rejection Algorithm in the FLA.

Ranking the Best Matched Fuse


In the previous sections, the time mismatch vector was normalized in order to rank the protective devices match to the observed operating time. The same process is performed here with
load mismatch instead of time mismatch. To choose the best matched fuse, the FLA uses unity
based min-max normalization. When the vector of mismatches is normalized, all load mismatch values will take on a value between 0 and 1. After being normalized the minimum load
mismatch is normalized to 0, representing the most likely candidate. The largest mismatch is
normalized to 1.0, representing the most unlikely candidate. All other values are appropriately
scaled between 0 and 1. To illustrate this, a test was run on a feeder with six fuses of different
load values. The load mismatches for each of the six fuses are as follows:

60

Ief use

3.1167

3.7360

0.5196

2.2753

1.8776

0.0476

(3.16)

To normalize the mismatch vector Ief use we use the following relation:
Id
ef use =

xi min(Ief use )
max(Ief use ) min(Ief use )

(3.17)

The variable Id
ef use represents the normalized load mismatch vector. Each data point
within Ief use vector is represented using xi and the min and max of the Ief use vector
is represented by min(Ief use ) and max(Ief use ), respectively. After normalization, the
normalized load mismatch vector Id
ef use is the following:

Id
ef use

0.8321

1.0000

0.1280

0.6040

0.4961

(3.18)

What we see in Id
ef use is a clear representation of the most likely fuse to have operated.
The third element in the vector was normalized to 0.1280, representing a likely candidate. The
second element in the vector was normalized to 1.0, representing the most unlikely candidate.
The last element in the vector was normalized to 0.0 which represents smallest load mismatch
and most likely candidate.

3.7

Combining Localization Data for Best Match

In previous sections, we were able to localize and rank a fault possibility from load flow rejection
and protective device characteristics. To provide the most accurate localization information,
the data from both the load flow rejection and protective device characteristics are combined.
The consideration of all localization data creates a final ranking vector in the FLA, which is
used to determine the best matched node.

61

3.7.1

Final Ranking of Each Possibility

To begin final ranking of a fault possibility, we again consider if the fault possibilitys protective
device is a recloser, breaker, or fuse. If the device is a recloser we must consider the load flow
rejection, short-time mismatch, long-time mismatch and open interval timing. If the device is
a fuse we must consider the load flow rejection and operating time mismatch.
Each possible fault location has a mismatch from the observed fault current which also must
be considered. The expected fault current If exp for each fault possibility is pulled from the
fault tables and compare to the observed fault current If ob . The current mismatch is defined
as:
If = |If ob If exp |

(3.19)

The fault current mismatch is normalized using the following formulation:


df =
I

xi min(If )
max(If ) min(If )

(3.20)

df represents the normalized fault current mismatch vector. Each data point
The variable I
within If vector is represented using xi and the min and max of the If vector is represented
by min(If ) and max(If ) respectively.
Each fault possibilities fault current mismatch and associated upstream protective device
mismatches are combined to calculate the final ranking vector n. For a fault possibility with a
upstream fuse:

nf use,k =

de
t
|{z}

Normalized Time Mismatch

Id
ef use
| {z }

Normalized Load Rejection Mismatch

df
I
|{z}

Normalized Fault Current Mismatch

(3.21)
For a fault possibility with a upstream recloser:

nrecloser,k =

be
|{z}

Normalized Time Mismatch

d
Ierecloser
|
{z
}

Normalized Load Rejection Mismatch

df
I
|{z}

Normalized Fault Current Mismatch

(3.22)
A master ranking vector is formed that contains all fault possibilities nr,1 , nr,2 , all the way
to k th and final fault possibility nr,k . The sub-script variable r is used to identify the protective
device type where r = f use or r = recloser.

62

nr,1

nr,2
=
..
.
nr,k

(3.23)

The matrix is then normalized using the following relation:


b =

xi min()
max() min()

(3.24)

The variable b represents the final normalized ranking vector. Each data point within
vector is represented using xi (nr,1 nr,k ) and the min and max of the vector is represented
by min() and max() respectively.
The final normalized ranking vector is used by the FLA to rank all the possibilities from
most likely to least likely. The normalized vector also allow the fault locator to quantify how
the intermediate possibilities are ranked relative to the best match.

63

3.8

Localization Flow Chart

To clarify the exact procedure used to localize the fault, a localization flow chart is shown in
Figure 3.18.

Figure 3.18: FLA Localization Algorithm Flow Chart.

64

Chapter 4

Fault Locator Test Results


4.1

Introduction to the Notional Feeder, Stewart Street Feeder


and Test Cases

To test the proposed FLA, a MATLAB Simulink model of a Progress Energy 12kV feeder
was used. Model parameters such as line impedances, transformer ratings and load data were
provided by Progress Energy. In this paper, we refer to this feeder as the Notional Feeder(NF).
MATLAB Simulink was chosen to simulate the NF because of its ability to perform discrete time
simulations and data capture. Current waveforms captured after simulation can be recorded in
Simulink and passed to the FLA for execution(Figure 4.1). A one-line of the NF is shown in
Figure 4.2.
This chapter is comprised of three different test cases using the NF model. Test Case 1 is
performed using the FLA with no localization algorithm. No protective devices are present in
the feeder model while testing. As a result, the FLA ranks multiple fault possibilities using the
observed fault current only.
During Test Case 2 and Test Case 3, fuse and recloser models are inserted into the NF.
With the addition of protective devices, the localization algorithm is used to rank possibilities.
During Test Case 3, the FLA is tested using both Fuse Saving and Fuse Blowing coordination
schemes.
During Test Case 4, the Stewart Street 12.47kV feeder is tested using DEW models provided
by Allegheny Power. The Stewart Street feeder is much larger than the NF, and contains
hundreds of nodes and tapped loads. Short circuit and load flow analysis is performed in DEW
and passed to the FLA localization algorithm. A detailed overview of the Stewart Street feeder
is given in Section 4.7.

65

4.2

Notional Feeder Fault Tables and Short Circuit Data

The fault tables needed by the FLA are gathered by applying a bolted fault(Rf = 0) at each
node in the NF. After a sufficient time has passed, the RMS current reaches a steady state
and it is recorded in Microsoft Excel spreadsheet format. The spreadsheet is automatically
imported into the FLA upon execution of the algorithm. The fault table used during Phase A
to ground faults is provided in Table 4.1.

Table 4.1: Notional Feeder Fault Table for SLG Faults on A Phase.
Node
Node
Node
Node
Node
Node
Node
Node
Node
Node
Node
Node
Node
Node
Node
Node
Node
Node
Node

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18

RMS Current

Fault Type

6950
5145
3196
3440
4422
4844
4989
4088
2585
2827
3325
2186
1886
4260
3080
3020
2819
8800

A-G
A-G
A-G
A-G
A-G
A-G
A-G
A-G
A-G
A-G
A-G
A-G
A-G
A-G
A-G
A-G
A-G
A-G

66

4.3
4.3.1

Test Case 1: Notional Feeder Testing With No Localization


Introduction: Test Conditions and Procedure

The first test case performed with the FLA is done with no localization on the Notional Feeder.
Simple fault tables are used to rank faults from best to worst match. Fault current data is
captured at the substation bus of the NF using the Goto Workspace block set in Simulink.
After the data is recorded in the MATLAB workspace, the FLA algorithm(written in MATLAB
code) is executed. After execution and the results are captured using a screen shot of the
MATLAB console. A high level overview of the test procedure is shown in Figure 4.1.

Figure 4.1: Simulation Testing Procedure using MATLAB Simulink.

Before simulating a system fault all initial states are started at steady-state conditions.
According to [7], most line-to-ground faults last 8.35 cycles before being cleared by protective
devices. As a result, all line-to-ground fault durations during Test Case 1 simulations last for
8.35 cycles being removed. During fault conditions, the fault is assumed to be bolted Rf = 0.
Since single line-to-ground faults are the most common, the most attention is given to this fault
type.

4.3.2

FLA Testing for Line-to-Ground Faults

In this section, we consider several line-to-ground faults at various nodes in the network. After
the fault is applied, the FLA executed and the MATLAB console printout is recorded. A lineto-ground fault was tested at the following nodes: Node 4, Node 11, Node 15, Node 17. For the
location of the faulted nodes, refer to the system one-line Figure 4.2. After the FLA is executed,
the calculated fault current mismatch percentage is recorded for ranking purposes. The percent
mismatch is the difference between the observed and expected fault current, relative to the
observed current, for each node. The node with the lowest mismatch is considered the best

67

Figure 4.2: Test Case 1 Notional Feeder One-Line Model.

match. A summary of the line-to-ground test results are in Section 4.3.4.

4.3.3

FLA Testing for Line-to-Line Faults

In this section, we also consider line-to-line faults at various nodes in the network. After the
fault is applied, the FLA executed and the MATLAB console printout is recorded. A line-toline fault was tested at the following nodes: Node 15 and Node 17. For the location of the
faulted nodes, refer to the system one-line Figure 4.2. After the FLA is executed, the fault
current mismatch percentage is recorded for ranking purposes. The percent mismatch is the
difference between the observed and expected fault current, relative to the observed current,
for each node. The node with the lowest mismatch is considered the best match. A summary
of the line-to-line test results are in Section 4.3.4.

68

4.3.4

Test Case 1 Results:

During Test Case 1, a total of 6 cases involving phase-to-phase and line-to-ground faults were
tested at various nodes in the network. In all test cases the faulted node was correctly identified.
A summary of the test results are in Tables 4.2 and 4.3.

Table 4.2: Line-to-Ground Simulation Results for Test Case 1


Simulation Number

A-G Fault at Node

Possibilities

1
2
3

Node 4
Node 11
Node 15

Node 17

Node 4
Node 11
Node 15(Best Match)
Node 16(Worst Match)
Node 17(Best Match)
Node 10(Worst Match)

Mismatch %
0.0347%
0.0343%
0.1499%
2.0951%
0.0456 %
0.2381%

Table 4.3: Phase-to-Phase Fault Simulation Results for Test Case 1


Simulation Number

A-G Fault at Node

Possibilities

Node 15

Node 17

Node 15(Best Match)


Node 16(Worst Match)
Node 17(Best Match)
Node 10(Worst Match)

69

Mismatch %
0.0493%
3.5486%
0.1266 %
2.702%

4.4

Test Case 2: Localization using Load Flow and Protective


Devices

During Test Case 1, the FLA only utilized fault current as a means of determining the fault
location and ranking(no protective devices were modelled). In Test Case 2, two substation
breakers and two reclosers are added to the NF model as recommended by Progress Energy.
With protective devices now present, load flow and protective device data is passed to the localization sub-algorithm to aid in the ranking process. The load flow data is captured using the
MATLAB load flow tool and imported into the fault locator using Microsoft Excel spreadsheets.
The fault locator also requires time-current characteristics on all reclosers and breakers present
in the feeder as well as their respective locations ([19],[20]). Locations of each device and their
time-current characteristics are imported into MATLAB using Microsoft Excel. Fault current
data is captured at the substation bus of the NF using the Goto Workspace block set in
Simulink. After the data is recorded in the MATLAB workspace, the FLA algorithm(written
in MATLAB code) is executed. An overview of the test procedure is shown in Figure 4.3.

Figure 4.3: Test Case 2 Testing Procedure.

Although all nodes were identified correctly in Test Case 1, many fault possibilities were
nearly indistinguishable in percentage mismatch. For example, the results for a line-to-ground
fault at Node 17 in Test Case 1 yielded a 0.045% mismatch and a 0.23% mismatch for Node
10. These two nodes are nearly indistinguishable and are only separated by a 0.185% margin.
During the proper system conditions, the FLA Algorithm could incorrectly choose Node 10
as the best matched since both node have very similar fault currents(this is investigated more
in Section 4.8.1). Given that Node 10 and Node 17 are located in two different branches of
the feeder, they are protected by different devices. If the fault were to have occurred at Node
17, Recloser A would have operated for the fault. If the fault were to have occurred at Node

70

10, Substation Breaker A would have interrupted the fault. Since Recloser A and Substation
Breaker A have different protective device settings, the fault location can be localized. Also,
Substation Breaker A and Recloser A both carry measurably different load levels, allowing yet
another parameter to localize with. After the recloser or breaker goes to the lockout state,
the load rejection is calculated by the fault locator and used to localize the fault. Using both
load flow and protective device characteristics allows the fault locator to localize two faults
that were nearly indistinguishable in Test Case 1. Some of the same cases in Test Case 1 are
run again to illustrate the effectiveness of the added localization techniques. The addition of
localization will show that many indistinguishable cases in Test Case 1(i.e Node 17 fault test)
are now resolved into much more confident choices. Most of the attention in this section will
be given to line-to-ground faults since they are the most common.

4.4.1

Introduction: Test Conditions and Procedures

During Test Case 1, no protective devices were present in the feeder and the fault was applied to
the network for a allotted amount of time. During Test Case 2, protective devices are modelled
and placed at designated points as recommended by Progress Energy Carolinas(as shown in
Figure 4.4). Each recloser and substation breaker is set to 1 shot of fast reclosing and 3 shots
of slow reclosing. Each recloser and substation breaker has two phase(fast and slow) and two
ground relays(fast and slow). After the final shot of reclosing, the breaker or recloser opens
and enters the lockout state. Progress Energy recommended breaker and recloser settings can
be found in Appendix A.7.
Given the computational intensity of these simulation and the length of time they take to
complete, only 1 shot of fast and 1 shot of slow was performed before lockout of the substation
breaker or recloser. Reducing the number of shots of each device did not effect the accuracy of
the fault locator since all devices are set to a total of four shots.

4.4.2

FLA Testing for Line-to-Ground Faults

In this section, we will consider line-to-ground faults at various nodes in the network. After the
fault is applied, the FLA executed and the MATLAB console printout is recorded. A line-toground fault was tested at the following nodes: Node 17, Node 7 and Node 16. For the location
of the faulted nodes, refer to the system one-line Figure 4.4. After the test simulation is run,
the protective device mismatch and load rejection mismatch percentage is recorded for ranking
purposes.

71

Figure 4.4: Notional Feeder One-Line Model with added Reclosers and Breakers.

4.4.3

A-Ground Fault at Node 17

A line-to-ground fault was place at Node 17 on Phase A as done in Test Case 1 resulting in
two possible nodes. The two possibilities, Node 17 and Node 10 are nearly indistinguishable
and are only separated by a few percent mismatch. In this section, we show that Node 10 can
easily be eliminated as a possibility with the addition of the localization algorithm.
After the ground fault was place at Node 17, the ground relay in Recloser A picked up and
operated two times(shown in Figure 4.5). The first shot was a fast(short time) trip, followed
by a open interval period of several cycles. The recloser then de-activates the fast ground curve
then switch to slow for the second trip. After the fault does not clear on the second shot, the
recloser goes into the lockout state and does not reclose. When the FLA has detected that the
fault has cleared, it begins searching for all possible fault locations. The FLA has been coded

72

Figure 4.5: Test Case 2: Substation Phase A RMS current during a fault at Node 17.

to use both fast and slow curves for localization.


After Recloser A proceeds to the lockout state, the observed current at the substation bus
decreases due to the disconnecting of load. The fault locator also uses this information to
determine the most likely device that has operated in the network in addition to the protective
device information. In Figure 4.5, the magnitude of the load rejection in shown.
Test Results
Data provided to the user of the FLA ranks the fault possibilities from best match to worst
match. For each fault location possibility, the FLA must determine the correct upstream protective device. Both Node 10 and 17s upstream protective devices were identified as Substation
Breaker A and Recloser A, respectively. Once a protective device is associated with a node, the
algorithm begins to localize using the observed operating times.
The protective device localizing data shows that the fault could not have occurred a Node
10. The percent error from the observed and estimated operating times for Node 10 is 243.5%,
indicating that for a fault at Node 10, the interrupting time would have been much slower.

73

The percent error from the observed and estimated operating times for Node 17 is only 30.4%,
indicating that the fault must have occurred within the zone of Recloser A(Node 17). A
summary of this data is shown in Table 4.4.

Table 4.4: Percent Mismatch Table for Protective Device Localization.


Protective Device

Percent Mismatch

Substation Breaker A(Node 10)


Recloser A(Node 17)

243.5%
30.4%

Load flow rejection provides a additional layer of security for localization. After the recloser
or breaker enters the lockout state, there is an expected current rejection at the substation for
each phase. Observed current rejections poorly matched a load rejection of Substation Breaker
A and is reflected in the mismatch calculated by the fault locator. For Node 10, the rejection
mismatch for Phase A, B and C are 458%, 175% and 455% respectively. This poor match
indicates that if the fault had occurred at Node 10, there would have been a much larger load
rejection. Node 17 is the preferred choice with a Phase A, B and C mismatch of 6.3%, 8.2%
and 6.84% respectively. This data is summarized in Table 4.5.

Table 4.5: Percent Mismatch Table for Load Rejection Localization


Protective Device

Percent Mismatch

Substation Breaker A(Node 10)


Substation Breaker A(Node 10)
Substation Breaker A(Node 10)
Recloser A(Node 17)
Recloser A(Node 17)
Recloser A(Node 17)

Phase A: 458%
Phase B: 175%
Phase C: 455%
Phase A: 6.3%
Phase B: 12.9%
Phase C: 5.54%

Using localization it is easy to distinguish that the fault cannot be at Node 10. In Test Case
1, the fault locator was able to still select Node 17 as the correct faulty node, however, there
were no measures of certainty. With the addition of localization data, Node 10 can be thrown

74

out as a possibility.

4.4.4

A-Ground Fault at Node 7

For this test a line-to-ground fault is applied to Node 7 in Substation Breakers B zone of protection. Using the fault tables for line-to-ground faults, only two possibilities can be established:
Node 6 and Node 7. Both of these nodes are nearly identical in fault current magnitude, making them indistinguishable if localization is not used. Since Node 6 is in Relcoser Bs zone of
protection, the localization algorithm can easily distinguish between the two locations using
load rejection and observed protective device operation.

Figure 4.6: Test Case 2: Substation Phase A RMS current during a fault at Node 7.

After the ground fault was place at Node 7(shown in Figure 4.4), the ground relay in
Substation Breaker B picked up and operated two times. The first shot was a fast(short time)
trip, followed by a open interval period of several cycles. The recloser then de-activates the fast
ground curve then switch to slow for the second trip(Shown in Figure 4.6) . After the fault
does not clear on the second shot, the recloser goes into the lockout state and does not reclose.
When the FLA has detected that the fault has cleared, it begins searching for all possible fault
locations.

75

Test Results
For a fault in the zone of protection of Substation Breaker B(Node 7), the interruption time is
much slower than Recloser B. Had the fault actually occurred in the zone of the recloser(Node 6)
the fault locator would have observed a much faster ground relay operation, thereby indicating
a fault at Node 6. Since the ground relay operation was much slower, the best matched device is
either Substation Breaker B or Substation Breaker A. However, since all faults in the Substation
Breaker As zone do not match the fault current magnitude of the observed fault, they can be
eliminated. This essentially eliminates Node 6 as a potential fault location and is therefore
ranked as worst match.
The protective device localizing data shows that the fault could not have occurred a Node
6. The percent error from the observed and estimated operating times for Node 6 is 138.5%,
indicating that for a fault at Node 6, the interrupting time would have been much faster. The
percent error from the observed and estimated operating times for Node 7 is only 18.24%,
indicating that the fault must have occurred within the zone of Substation Breaker B(Node 7).
A summary of this data is shown in Table 4.6.

Table 4.6: Percent Mismatch Table for Protective Device Localization


Protective Device

Percent Mismatch

Substation Breaker B(Node 7)


Recloser B(Node 6)

18.24%
138.5%

Load flow rejection provides a additional layer of security for localization. After the recloser
or breaker enters the lockout state, there is an expected current rejection at the substation for
each phase. Observed current rejections poorly matched a load rejection of Recloser B and is
reflected in the mismatched calculated by the fault locator. For Node 6, the rejection mismatch
for Phase A, B and C are 48.9%, 50.1% and 50.1% respectively. This poor match indicates
that if the fault had occurred at Node 6, there would have been a much smaller load rejection.
Node 7 is the preferred choice with a Phase A, B and C mismatch of 4.22%, 1.20% and 1.87%
respectively. A summary of this data is in Table 4.7.
Using localization it is easy to distinguish that the fault cannot be at Node 6. With the
above data, Node 6 can be thrown out as a possibility.

76

Table 4.7: Percent Mismatch Table for Load Rejection Localization


Protective Device

Percent Mismatch

Substation Breaker B(Node 7)


Substation Breaker B(Node 7)
Substation Breaker B(Node 7)
Recloser B(Node 6)
Recloser B(Node 6)
Recloser B(Node 6)

4.4.5

Phase
Phase
Phase
Phase
Phase
Phase

A:
B:
C:
A:
B:
C:

4.22%
1.20%
1.87%
48.9%
50.1%
50.2%

Limitations of the Localization Algorithm: A-Ground Fault at Node


16

In certain cases, the proposed localization algorithm built into the FLA will not be able to
provide any additional information on the true fault location. This phenomena occurs when
two possibilities exist in the same zone of protection for a given device. When a fault occurs
in the zone of protection for any given device, such as a fuse, the fuse will operate and result
in a load rejection at the substation. If two possibilities exist in the zone of the device that
operated, it will result in a identical load and protective mismatch vector for both possibilities.
Therefore, it can be said that if two possibilities exist within the same zone of protection, they
cannot be localized.
An excellent example of this in the NF is nodes 15 and 16. Both of these locations are
separated by a short line segment approximately .356 km long, offering minimal short circuit
difference between the nodes. This results in mutual possibilities when a fault occurs at either
of the nodes. If a fault occurs at Node 15, then a Node 16 will be a mutual possibility. To test
this, a line-to-ground fault was placed at Node 16, resulting in two ground relay operations(one
fast and one slow) of Recloser A. After two shots of reclosing, Recloser A enters the lockout
state and remains in the open position.
Since both nodes are in the same zone of protection, they will have the same protective
device mismatch and load rejection mismatch. The fault locator is forced to use only fault
current to localize the fault. When this occurs, the fault locator will print a warning to the
user that the localization algorithm detected two fault possibilities within the same zone of
protection. This is shown in Figure 4.7. In this particular test, Node 16 was still identified as
the correct faulted node. This was achieved by using fault tables only(Test Case 1).
We can conclude that the proposed localization algorithm is effective if all possibilities do

77

Figure 4.7: Fault Locator Ranking of Fault at Node 16.

not exist within the same zone of protection for a given device. If this is the case, only fault
tables can be used to localize. As of late, the popularity of fuse blowing schemes requires that
fault locating algorithms adapt to this protective scheme. A properly coordinated fuse blowing
scheme allows the fuse to interrupt the fault before any upstream device(such as a recloser) is
allowed to trip.

78

4.5

Test Case 3: Fault Locating on Fuse Saving or Fuse Blowing


Schemes

4.5.1

Introduction: Notional Feeder with Fuse Blowing Coordination

During Test Case 2, the NF was protected by both reclosers and reclosing substation breakers
without the presence of fuses. Given the popularity of fuse blowing schemes, a new case was
introduced to test this protection scheme. A properly coordinated fuse blowing scheme allows
the fuse to interrupt the fault before the upstream breaker can trip. The FLA presented here
is easily adaptable to fuse blowing schemes, however, the fault locator must be told that the
feeder is being protected by a fuse blowing scheme.

Figure 4.8: Fuse Blowing Scheme on Notional Feeder.

79

4.5.2

Introduction: Fuse Blowing Scheme Test Conditions

To implement a common fuse blowing scheme, the NF protection system was reconfigured
to contain 26 SandC T Speed Fuses and a main reclosing substation breaker. Pickups and
time-dial settings were changed in the substation breaker to allow fuse blowing coordination
between all downstream fuses. Each fuse is named in the feeder and is shown along with the
characteristic in Figure 4.8. The reclosing substation feeder breaker fast curve was disabled,
thereby allowing the fuse to blow. The reconfigured protection system is shown in Figure 4.8.
Each fuse link was modelled in MATLAB Simulink using test data provided by SandC [21]. To
ensure adequate coordination, ASPEN One-Liner coordination tool was used to ensure proper
fuse-fuse and fuse-breaker coordination per IEEE standards and recommendations[17]. All fuse
and breaker information was imported into the FLA for fault localization.

4.5.3

FLA Testing for Line-to-Ground Faults: Fuse Blowing

In this section, we will consider line-to-ground faults at various nodes in the network. After the
fault is applied, the FLA executed and the MATLAB console printout is recorded. A line-toground fault was tested at the following nodes: Node 6, Node 10 and Node 17. For the location
of the faulted nodes, refer to the system one-line Figure 4.8. After the test simulation is run,
the protective device mismatch and load rejection mismatch percentage is recorded for ranking
purposes.

4.5.4

Line-to-Ground Fault at Node 17

The first fuse blowing scheme test was performed with a line-to-ground fault at Node 17. For
a fault at this node, the fault locator will produce two possible fault locations: Node 10 and
Node 17. Node 10 can be easily eliminated using the localization algorithm due to the slower
speed 200T fuse protecting this node. As a result, the fault locator will observe a faster fuse
operation, indicating that the fault is in the zone of a 140T or faster fuse. Since the only other
fault possibility is Node 17, which is in the zone of a 140T fuse, it is ranked as best match.
During the fault, the fault locator records the observed fuse operating time which is used
for localization. During this part of the algorithm, the fault locator will notify the user that
it is localizing using a fuse blowing scheme. During testing, the fault locator has correctly
identified that the fault must be in the zone of a fuse, since no reclosing was observed. This
data indicates that the fault could not have occured on the main feeder, as it is protected only
by the substation breaker. As indicated in the Figure 4.9, the observed fuse operating time is
.353 seconds.

80

Figure 4.9: Fault Locator Observed Fault Current for Fault at Node 17.

Test Results
Data provided to the user of the FLA ranks the fault possibilities from best match to worst
match. For each fault location possibility, the FLA must determine the correct upstream
protective device. Both Node 10 and 17s upstream protective devices were identified as Fuse
N142 and Fuse N14, respectively. Once a protective device is associated with a node, the
algorithm begins to localize using the observed operating times.
The protective device localizing data shows that the fault could not have occurred a Node
10. The percent error from the observed and estimated operating times for Node 10 is 168.3%,
indicating that for a fault at Node 10, the interrupting time would have been much slower.
The percent error from the observed and estimated operating times for Node 17 is only 14.5%,
indicating that the fault must have occurred within the zone of Fuse N14. A summary of this
data is shown in Table 4.8.
Load flow rejection provides a additional layer of security for localization. After the fuse
blows, there is an expected current rejection at the substation for the applicable phase. Observed current rejections poorly matched a load rejection of Fuse N142 and is reflected in the
mismatched calculated by the fault locator. For Node 10, the rejection mismatch for Phase

81

Table 4.8: Percent Mismatch Table for Protective Device Localization


Protective Device

Percent Mismatch

Fuse N14(Node 17)


Fuse N142(Node 10)

14.5%
168.3%

A is 377.0%. This matching indicates that if the fault had occurred at Node 10, there would
have been a larger load rejection. Node 17 is the preferred choice with a Phase A load rejection
mismatch of 14.7%. A summary of this data is in Table 4.9.

Table 4.9: Percent Mismatch Table for Load Rejection Localization


Protective Device

Percent Mismatch

Fuse N14(Node 17)


Fuse N142(Node 10)

Phase A: 14.7%
Phase A: 377.0%

Using localization it is easy to distinguish that the fault cannot be at Node 10. With the
above data, Node 10 can be thrown out as a possibility.

82

4.5.5

Line-to-Ground Fault at Node 6

The second fuse blowing scheme test was performed with a line-to-ground fault at Node 6. For
a fault at this node, the fault locator will produce two possible fault locations: Node 6 and
Node 7. Node 7 can be easily eliminated using the localization algorithm due to the faster
speed 50T fuse protecting this node. As a result, the fault locator will observe a slower fuse
operation, indicating that the fault is in the zone of a slower speed fuse. Since the only other
fault possibility is Node 6, which is in the zone of a 200T fuse, it is ranked as best match.

Figure 4.10: Fault Locator Observed Fault Current for a fault at Node 6.

During the fault, the fault locator records the observed fuse operating time which is used
for localization. During this part of the algorithm, the fault locator will notify the user that it
is localizing using a fuse blowing scheme. Also, the fault locator correctly identified that the
fault must be in the zone of a fuse, since no reclosing was observed. This data indicates that
the fault could not have occurred on the main feeder, as it is protected only by the substation
breaker. As indicated by Figure 4.10, the observed fuse operating time is .229 seconds.

83

Test Results
Data provided to the user of the FLA ranks the fault possibilities from best match to worst
match. For each fault location possibility, the FLA must determine the correct upstream
protective device. Both Node 6 and 7s upstream protective devices were identified as Fuse N1
and Fuse N23, respectively. Once a protective device is associated with a node, the algorithm
begins to localize using the observed operating times.

Table 4.10: Percent Mismatch Table for Protective Device Localization


Protective Device

Percent Mismatch

Fuse N1(Node 6)
Fuse N23(Node 7)

32.31%
89.6%

The protective device localizing data shows that the fault could not have occurred a Node
7. The percent error from the observed and estimated operating times for Node 7 is 89.6%,
indicating that for a fault at Node 7, the interrupting time would have been much faster. The
percent error from the observed and estimated operating times for Node 6 is only 32.31%,
indicating that the fault must have occurred within the zone of Fuse N1. A summary of this
data is shown in Table 4.10.
Load flow rejection provides a additional layer of security for localization. After the fuse
blows, there is an expected current rejection at the substation for the applicable phase. Observed current rejections poorly matched a load rejection of Fuse N23 and is reflected in the
mismatched calculated by the fault locator. For Node 7, the rejection mismatch for Phase A
is 70.7%. This poor match indicates that if the fault had occurred at Node 7, there would
have been a much smaller load rejection. Node 6 is the preferred choice with a Phase A load
rejection mismatch of 8.27%. A summary of this data is in Table 4.11.

Table 4.11: Percent Mismatch Table for Load Rejection Localization


Protective Device

Percent Mismatch

Fuse N1(Node 6)
Fuse N23(Node 7)

Phase A: 8.27%
Phase A: 70.7%

84

Using localization it is easy to distinguish that the fault cannot be at Node 7. With the
above data, Node 7 can be thrown out as a possibility.

4.5.6

Introduction: Notional Feeder with Fuse Saving Coordination

Although fuse saving is falling out of favor with many utilities, it is still in mainstream use. A
properly coordinated fuse saving scheme allows the upstream recloser instantaneous overcurrent
element to trip before the fuse is allowed to blow. Reclosing before the fuse is allowed to blow
allows a temporary fault to be cleared by recloser rather than by the fuse. This reduces outages
and the amount of line crew dispatches into the field. The FLA presented here is easily adaptable to fuse saving schemes, however, the fault locator must be told that the feeder is being
protected by a fuse saving scheme. In the event that the feeder is set to a hybrid scheme(fuse
saving and fuse blowing), the fault locator must be set to fuse saving mode.

4.5.7

Introduction: Fuse Saving Scheme Test Conditions

To implement a common fuse saving scheme, the Notional Feeder protection system was reconfigured to contain 26 SandC T Speed Fuses and a main reclosing substation breaker. Pickups
and time-dial settings were changed in the substation breaker to allow fuse saving coordination
between all downstream fuses. The reclosing substation feeder breaker was set to one shot
of fast before switching to the slow curve and allowing the fuse to blow. The reconfigured
protection system is shown in Figure 4.8. Each fuse link was modelled in MATLAB Simulink
using test data provided by SandC [21]. To ensure adequate coordination, ASPEN One-Liner
coordination tool was used to ensure proper fuse-fuse and fuse-breaker coordination per IEEE
standards and recommendations. All fuse and breaker information was imported into the FLA
for fault localization.

4.5.8

FLA Testing for Line-to-Ground Faults: Fuse Saving

In this section, we will consider line-to-ground faults at various nodes in the network. After
the fault is applied, the FLA executed and the MATLAB console printout is recorded. A lineto-ground fault was tested at the following nodes: Node 10 and Node 6. For the location of
the faulted nodes, refer to the system one-line Figure 4.8. After the test simulation is run,
the protective device mismatch and load rejection mismatch percentage is recorded for ranking
purposes.

85

4.5.9

Line-to-Ground Fault at Node 10

The first fuse saving scheme test was performed with a line-to-ground fault at Node 10. For
a fault at this node, the fault locator will produce two possible fault locations: Node 10 and
Node 17. Node 17 can be easily eliminated using the localization algorithm due to the faster
speed 140T fuse protecting this node. As a result, the fault locator will observe a slower
fuse operation, indicating that the fault is in the zone of a 200T or slower fuse. Since the only
other fault possibility is Node 10, which is in the zone of a 200T fuse, it is ranked as best match.
As shown in Figure 4.11, the substation breaker instantaneous ground element interrupts
the fault before the fuse is allowed to operate. After a open interval time, the breaker recloses
with the instantaneous ground element disabled. The ground time-overcurrent element remains
enabled and allows the fuse to blow. After the fuse blows, the ground fault is cleared.

Figure 4.11: Fault Locator Observed Fault Current for a Fault at Node 10.

Test Results
Data provided to the user of the FLA ranks the fault possibilities from best match to worst
match. For each fault location possibility, the FLA must determine the correct upstream

86

protective device. Both Node 10 and 17s upstream protective devices were identified as Fuse
N142 and Fuse N14, respectively. Once a protective device is associated with a node, the
algorithm begins to localize using the observed operating times. The protective device localizing
data shows that the fault could not have occurred a Node 17. The percent error from the
observed and estimated operating times for Node 17 is 53.2%, indicating that for a fault at
Node 17, the interrupting time would have been much faster. The percent error from the
observed and estimated operating times for Node 10 is only 9.71%, indicating that the fault
must have occurred within the zone of Fuse N142. A summary of this data is shown in Table
4.12.

Table 4.12: Percent Mismatch Table for Protective Device Localization


Protective Device

Percent Mismatch

Fuse N14(Node 17)


Fuse N142(Node 10)

53.2%
9.71%

Load flow rejection provides a additional layer of security for localization. After the fuse
blows, there is an expected current rejection at the substation for applicable phase. Observed
current rejections poorly matched a load rejection of Fuse N14 and is reflected in the mismatched
calculated by the fault locator. For Node 17, the rejection mismatch for Phase A is 75.1%. This
poor match indicates that if the fault had occurred at Node 10, there would have been a much
smaller load rejection. Node 10 is the obvious choice with a Phase A load rejection mismatch
of 3.26%. A summary of this data is in Table 4.13.

Table 4.13: Percent Mismatch Table for Load Rejection Localization


Protective Device

Percent Mismatch

Fuse N14(Node 17)


Fuse N142(Node 10)

Phase A: 75.1%
Phase A: 3.26%

Using localization it is easy to distinguish that the fault cannot be at Node 17. With the
above data, Node 17 can be thrown out as a possibility.

87

4.5.10

Line-to-Ground Fault at Node 6

The second fuse saving scheme test was performed with a line-to-ground fault at Node 6. For
a fault at this node, the fault locator will produce two possible fault locations: Node 6 and
Node 7. Node 7 can be easily eliminated using the localization algorithm due to the faster
speed 50T fuse protecting this node. As a result, the fault locator will observe a slower fuse
operation, indicating that the fault is in the zone of a slower speed fuse. Since the only other
fault possibility is Node 6, which is in the zone of a 200T fuse, it is ranked as best match.
As shown in Figure 4.12, the substation breaker instantaneous ground element interrupts
the fault before the fuse is allowed to operate. After a open interval time, the breaker recloses
with the instantaneous ground element disabled. The ground time-overcurrent element remains
enabled and allows the fuse to blow. After the fuse blows, the ground fault is cleared.

Figure 4.12: Fault Locator Observed Fault Current for a Fault at Node 6.

Test Results
Data provided to the user of the FLA ranks the fault possibilities from best match to worst
match. For each fault location possibility, the FLA must determine the correct upstream pro-

88

tective device. Both Node 6 and 7s upstream protective devices were identified as Fuse N1
and Fuse N23, respectively. Once a protective device is associated with a node, the algorithm
begins to localize using the observed operating times.
The protective device localizing data shows that the fault could not have occurred a Node
7. The percent error from the observed and estimated operating times for Node 7 is 89.6%,
indicating that for a fault at Node 7, the interrupting time would have been much faster.
The percent error from the observed and estimated operating times for Node 6 is only 32.3%,
indicating that the fault must have occurred within the zone of Fuse N1. A summary of this
data is shown in Table 4.14.

Table 4.14: Percent Mismatch Table for Protective Device Localization


Protective Device

Percent Mismatch

Fuse N1(Node 6)
Fuse N23(Node 7)

32.3%
89.6%

Load flow rejection provides a additional layer of security for localization. After the fuse
blows, there is an expected current rejection at the substation for the applicable phase. Observed current rejections poorly matched a load rejection of Fuse N23 and is reflected in the
mismatched calculated by the fault locator. For Node 7, the rejection mismatch for Phase A
is 70.7%. This poor match indicates that if the fault had occurred at Node 7, there would
have been a much smaller load rejection. Node 6 is the preferred choice with a Phase A load
rejection mismatch of 8.27%. A summary of this data is in Table 4.15.

Table 4.15: Percent Mismatch Table for Load Rejection Localization


Protective Device

Percent Mismatch

Fuse N1(Node 6)
Fuse N23(Node 7)

Phase A: 8.27%
Phase A: 70.7%

Using localization it is easy to distinguish that the fault cannot be at Node 7. With the

89

above data, Node 7 can be thrown out as a possibility.

4.5.11

Limitations of the FLA on Fuse Blowing or Fuse Saving Coordinated


Feeders

One of the inherent drawbacks of the FLA is that the fault current must reach steady state
for the fault location to be determined. In fuse blowing and fuse saving schemes, high speed
fuses(N, QR and K) can melt before the current reaches a steady state value. This phenomena
was observed in the Notional Feeder at Node 8, which is protected by a 50T fuse. The available
fault current at this node makes the clearing time of this fuse very quick, and subsequently the
fault current does not reach steady state. As a result, the fault locator failed to determine the
location of the fault. Figure 4.13 shows the observed fault current by the fault locator for a
fault at Node 8, notice that the current never reaches steady state.

Figure 4.13: Fault Locator Observed Fault Current for Node 8 Fault(Fuse Blowing Coordination).

90

4.6

Notional Feeder Test Results Summary

The following sections give a brief summary of the test results obtained during each test case
using the Notional Feeder(NF).

4.6.1

Test Case 1 Summary

During Test Case 1, the proposed FLA was tested on the NF containing no protective devices
with no localization. Phase-to-Ground faults were applied at Node 4, Node 11, Node 15 and
Node 17. The FLA was also tested with Phase-to-Phase faults at Node 15 and Node 17. During
all Phase-to-Phase and Phase-to-Ground tests, the FLA correctly identified the faulty node.

4.6.2

Test Case 2 Summary

During Test Cases 2, substation breakers and reclosers were added to the NF as recommended
by Progress Energy. Phase-to-ground faults were placed in the system at Node 17, Node 7
and Node 16. During all Phase-to-Ground tests, the FLA correctly identified the faulty node.
Detailed test data for Node 17 and Node 7 showed that the localization algorithm can effectively
resolve multiple possibilities into one certain fault location. However, the FLA localization does
have limitations. During a ground fault at Node 16, the localization algorithm was unable to
localize multiple possibilities. This was due to multiple possibilities existing in the same zone
of protection.

4.6.3

Test Case 3 Summary

During Test Case 3, the NF was reconfigured to accommodate fuse saving and fuse blowing
schemes. Fuses were added at specific points in the NF using SandC T Speed fuse characteristics. With fuse blowing coordination, the FLA was tested by applying phase-to-ground faults
at Node 17, Node 6 and Node 8 in the feeder model. The FLA correctly identified the faulty
nodes for all tests, except Node 8. Detailed test data for Node 17 and Node 6 showed that the
localization algorithm can effectively resolve multiple possibilities into one certain fault location
for fuse blowing coordinated feeders. During a ground fault at Node 8, it was noticed that the
FLA failed to locate the fault. One of the inherent drawbacks of the FLA is that the fault
current must reach steady state for the fault location to be determined. The available fault
current at this node made the clearing time of the fuse very quick, and subsequently the fault
current does not reach steady state. As a result, the FLA was unable to identify the faulty
node for a line-to-ground fault at Node 8.

91

The final testing during Test Case 3 required re-coordinating the NF for fuse saving. With
fuse saving coordination, the FLA was tested by applying phase-to-ground faults at Node 10
and Node 6 in the feeder model. The FLA correctly identified the fault nodes for all tests. Test
results for Node 10 and Node 6 showed that the localization algorithm can effectively resolve
multiple possibilities into one certain fault location for fuse saving coordinated feeders.

92

4.7
4.7.1

Test Case 4: Fault Locating on Large Scale Feeders


Introduction: Stewart Street 12.47kV Feeder

During Test Cases 1, 2 and 3, the Notional Feeder system model only consisted of 18 nodes,
which is much smaller than most conventional feeders. Although it showed promising results,
the algorithm must be tested on a much larger feeder to determine its accuracy and performance.
To test the proposed FLA, a feeder was model was built in DEW by Allegheny Energy for short
circuit analysis and load flow studies. The model consists of 405 power poles, 233 Single Phase
Transformers, 211 3-Phase Line Sections, 51 2-Phase Lines and 480 Single Phase Lines. The
protective equipment in the network consists of one recloser, and 69 T Speed Fuse Cutouts with
fuse blowing coordination. A picture of the Stewart Street feeder is shown in Figure 4.14.

Figure 4.14: Stewart Street 12.47kV Feeder Modelled in DEW.

93

4.7.2

Introduction: Test Conditions and Procedures

The FLA proposed in this paper uses sampled waveform data to calculate the location of the
fault. As of present, DEW does not offer a time domain based solution when calculating faults.
As a result, a MATLAB Simulink equivalent model has been proposed to act as a interface
between DEW and the FLA. Short circuit information and power flow data are passed to the
MATLAB model to form the equivalent model. Details about the equivalent model are covered
in Section 4.7.5.
To obtain the short circuit data, the DEW Fault Analysis tool is used. This tool allows the
user to calculate the reflected fault current at the substation bus under loaded conditions. When
calculating the fault current under loaded conditions, DEW assumes that the load currents do
not change after the fault is applied(constant current load). In the background of DEW, the
power-flow algorithm is run on the feeder to obtain all pre-fault quantities. More detailed
information about the formation of the FLA fault tables is discussed in Section 4.7.3.

Figure 4.15: Test Procedure for Stewart Street Feeder using DEW.

To test the FLA, a bolted line-to-ground fault is placed at various places in the Stewart
Street Feeder. Faults are placed at a chosen power pole such as P4622(PXXX Format). The
short circuit data(fault current magnitude and angle) is passed to the MATLAB Simulink
equivalent model in addition to pre-fault load data. The pre-fault load current(magnitude and
angle) are obtained via the Power Flow Tool in DEW. After passing the short circuit and prefault data to MATLAB, the Simulink model is executed along with the FLA. The results of the
FLA are recorded in the proceeding sections. A high level overview of the testing procedure is
shown in Figure 4.15.

94

4.7.3

Fault Tables for the Stewart Street Feeder

As previously mentioned, the FLA determines the location of the fault from the observed fault
current at the substation bus. To obtain the fault current magnitude at the substation bus
under loaded conditions, DEW assumes that the load currents do not change during fault
conditions(constant current load model)[22]. The resultant fault current is the sum of the
currents before the fault occurred(pre-fault) and the currents due to the fault itself. The DEW
network fault analysis tool uses the pre-fault load flow solution and sums this with the fault
current due to the fault itself. Using the display toolbox in DEW, the reflected fault current
at the substation bus is calculated for a phase-to-ground fault under loaded conditions and is
shown in Figure 4.16.

Figure 4.16: Reflected Fault Current at the Substation Bus for a Fault at Pole P4622.

Since single line-to-ground are the most common fault type, focus was given to this fault
type. Each line section in the Stewart Street feeder was faulted with a bolted line to ground
fault and the resultant reflected fault current observed at the substation bus was recorded to
form the FLAs fault tables.

4.7.4

Load Flow Analysis on the Stewart Street 12.47kV Feeder

To obtain the fault current magnitude under loaded conditions, DEW assumes that the load
currents do not change during fault conditions(constant current load model)[22]. The load flow
solution at the substation is needed later when building the MATLAB equivalent model in
MATLAB. After executing the DEW power flow tool, the resultant power flow solution is given
in Table 4.16.

95

Table 4.16: Substation Bus Load Flow Solution for Stewart Street Feeder.

4.7.5

Bus

Current Phase A

Current Phase B

Current Phase C

Substation Bus
Total Load
Total Load

421.90 6 -30.95
2703.51 kW
1620.84 kVAR

310.96 6 -148.86
2029.53 kW
1118.56 kVAR

397.01 6 90.34
2577.58 kW
1467.84 kVAR

MATLAB Fault Modelling of the Stewart Street 12.47kV Feeder

DEW software package does not offer time domain waveform fault data that was previously
available in other test cases using MATLAB. Currently, DEW only provides fault current in
phasor form. The FLA designed here only accepts raw time domain data and will not accept a
phasor quantity. Therefore, to test the FLA on the Stewart Street feeder, a MATLAB interface
model was developed using current sources and T Speed fuse models.
To begin developing the interface model, we can say that the fault locator at the substation
bus will observe a sum of the currents before the fault occurred I 0 and the currents due to the
fault itself If .
If sub = I 0 + If

(4.1)

Using the thevanin impedances and pre-fault voltages, we can easily calculate the current
flow due to the fault itself indicated by the equation below(Equation 4.2). Using the impedance
tool in DEW, the thevanin impedance looking into the fault point Zth can be calculated. The
pre-fault voltage V 0 is available via the DEW power flow tool. This allows the user to easily
calculate the current flowing due to the fault itself.
If =

V0
Zth

(4.2)

We can also say the the fuse protecting the faulted line segment will also observe a sum of
the currents before the fault occurred If0use and the currents due to the fault itself If .
If f use = If0use + If

(4.3)

Also, we must take into account that the fault locator must observe the full load current
during the fault. Therefore a current source is placed in the circuit with a magnitude of the full
load current with the fuse load current subtracted(Equation 4.4). This allows the fault locator

96

to no only see the full load current during the fault but the correct load rejection after the fuse
operates.
Irej = I 0 If0use

(4.4)

During the pre-fault state, the If0use current source switch is closed which allows the fault
locator at the substation bus to observe full load current and the fuse to observe its full load
current. The pre-fault model is shown in Figure 4.17. In the faulted model shown in Figure
4.18, the If f use current source switch is closed which allows the fault locator at the substation
bus to observe the current due to the fault itself in addition to the full load. Also, the fuse is
observing the current due to the fault itself in addition to its connected load. As shown in the
post-fault model in Figure 4.19, the correct observed current by the fault locator is I 0 If0use .
This requires neither of the two current sources and both are disconnected.

Figure 4.17: Pre-Fault Short Circuit Model.

Figure 4.18: Faulty Short Circuit Model.

97

Figure 4.19: Post-Fault Short Circuit Model.

4.7.6

Fault at Pole P4622 on the Stewart Street 12.47kV Feeder

FLA Testing Introduction


The first test using the Stewart Street Feeder was performed using a single line-to-ground
fault at distribution Pole P4622. This node is located on a remote portion of a lateral tap
approximately midway down the feeder as shown in Figure 4.20. Node P4622 is protected by a
65T T-Link Fuse that is named FUSE654421506 by DEW(UID-Unique Identifier).
MATLAB Equivalent Model: Load Flow and Fault Analysis Data
To begin, several parameters were collected from the DEW load flow and fault analysis tools.
First, the load flow analysis tool was used to obtain the current flow through the fuse during
normal load conditions. By selecting the fuse after load flow is performed, we obtain the prefault voltage and current solution. The pre-fault values obtained via load flow is listed in Table
4.17.

Table 4.17: Pre-Fault Load Flow Values


Quantity

Value

Pre-Fault Voltage V 0
Pre-Fault Current If0use

6.9616 4.19 kV
27.576 31.63 A

One of the quantities needed in the MATLAB interface model is the current due to the fault
itself, If . To calculate this quantity, the thevanin impedance is needed Zth (Refer to Equation
4.2). The thevanin impedance can be obtained by using the Fault Analysis tool in DEW and

98

Figure 4.20: P4622 Node Location in Stewart Street 12.47kV Feeder Modelled in DEW.

is given in Table 4.18 . Using the pre-fault voltage and thevanin impedance, we can calculate
the reflected fault current observed at the substation due to the fault itself : If =

V0
Zth .

This

leads to the following equation:


If =

6.9616 4.19 kV
= 1400.866 71.52
4.9696 67.33

(4.5)

Using the DEW fault analysis tool, we can easily check this calculation to ensure that it is
valid. Referring to Figure 4.21, we can see that DEW obtains a solution of 1400.51 A observed
at the substation recloser. Again, this solution is the fault current due to the fault itself, and
load currents have not been considered.

99

Table 4.18: Fault Analysis Calculated Parameters


Quantity

Value

Thevanin Impedance Zth


Fault Current If

4.969096 67.33 Ohms


1400.866 71.52 A

Figure 4.21: Reflected Fault Current at the Substation due to the fault itself.

As we previously indicated, during a fault the upstream fuse will observe If f use = If0use +If
where If = 1400.866 71.52 and the pre-fault load flow through the fuse is If0use = 27.576
31.63 . If we sum these values, we will obtain the observed fault current under loaded conditions
for the fuse.

If f use = If0use + If = 1400.866 71.52 + 27.576 31.63 = 1422.126 70.80

(4.6)

We can quickly check this result in DEW by applying a fault at Node P4622 under loaded
conditions. DEW obtains a loaded fault solution of If f use = 14206 70.8 as shown in Figure
4.22.
Using load flow data for the substation bus in Table 4.16 at the beginning of the chapter, we
can calculate the observed fault current at the substation under loaded conditions. To calculate
the observed fault current we sum the current due to the fault itself and the pre-fault load flow:
If sub = I 0 + If . The pre-fault load flow I 0 = 421.96 30.95 and the fault current due to the
fault itself is If = 1400.866 71.52.

100

Figure 4.22: Observed Fault Current at FUSE654421506 during loaded conditions.

If sub = If0use + If = 421.96 30.95 + 1400.866 71.52 = 1743.076 62.46

(4.7)

We can quickly verify this result in DEW by applying a fault at P4622 and clicking the
substation recloser to obtain the reflected fault current at the substation. DEW calculates
If sub = 1741.526 62.46 for the observed fault current at the substation during loaded
conditions.

Figure 4.23: Observed Fault Current at Substation during loaded conditions.

101

The final value that is needed for the switching model is the bus current Irej . This current
is calcuated as:
Irej = I 0 If0use = 421.96 30.95 27.576 31.63 = 394.446 30.90
A summary of the derived MATLAB interface model values is given in Table 4.19.

Table 4.19: MATLAB Model Parameters


Quantity
Fuse Pre-Fault Load If0use
Fuse Loaded Fault Current If f use
Load Rejection Irej

102

Value
31.63 A
1422.126 70.80 A
394.446 30.90 A
27.576

(4.8)

4.7.7

FLA Test Results for P4622 Fault

The FLA calculates 30 possible fault locations throughout the feeder. In Figure 4.24, the fault
possibilities are shown plotted on the feeder map, each represented by a red circle.

Figure 4.24: Possible Fault Locations in Stewart Street 12.47kV Feeder for a fault at P4622.

The FLA identifies Node P4622 as the best matched possibility for the fault location. A
screen-shot of the MATLAB is console is taken after the FLA is run and is shown in Figure
4.25. The FLA also lists Node P4619 as a best match because P4619 and P4622 are only 1
power pole span apart. Since both nodes are protected by the same fuse, and the two nodes are
nearly electrically identical in short circuit current, the fault locator cannot localize any farther.
The worst matches occurred at nodes P4238, P4236, P4235, N7970, N7955, N7972, N7973,
P4240, P4247, and L1254805 which occurred in the zone of the recloser. The recloser phase and

103

Figure 4.25: Best Matched Locations for a fault at P4622 as calculated by the FLA.

ground curves operate much slower than the 65T fuse characteristic that protects P4622. Since
the recloser also carries more than 400A of load current, the fault locator ranks these nodes as
the worst matched.
Several other nodes were listed a likely possibilities: P4498, P4509, P4513, P4515, P4520,
P4522, L1279017, N7974, 1L5936, N7976, N7977, P4241, L1254804, L1254806, L1254807,
L1254808, L1254809, and P4268. These possibilities along with their total respective rank
is shown in Appendix D, Table D.1.

104

4.7.8

Fault at Pole P4266 on the Stewart Street 12.47kV Feeder

The second test using the Stewart Street Feeder was performed using a single line-to-ground
fault at distribution pole P4266. This node is located on a remote portion of a lateral tap at
the end of the feeder as shown in Figure 4.26. Node P4266 is protected by a 200T T Speed
Fuse that is named FUSE1052482746 by DEW(UID-Unique Identifier).

Figure 4.26: P4266 Node Location in Stewart Street 12.47kV Feeder Modelled in DEW.

MATLAB Equivalent Model: Load Flow and Fault Analysis Data


To begin, several parameters were collected from the DEW load flow and fault analysis tools.
To begin, the load flow analysis tool was used to obtain the current flow through the fuse during
normal load conditions. By selecting the fuse after load flow is performed, we get the current
solution shown in Figure 4.27. This indicates that our pre-fault load flow through the fuse is
If0use = 9.036 32.34 .

105

Figure 4.27: Upstream Fuse(FUSE1052482746) Load Flow Solution.

The next element we must fetch from the load flow tool is the pre-fault voltage at the fault
point, V 0 . To retrieve this value, the load flow tool is executed and pole P4266 is selected. As
shown in Figure 4.28, V 0 = 6.9476 4.45 kV. A summary of the pre-fault values are given in
Table 4.20.

Figure 4.28: Pre-Fault Voltage at P4266.

106

Table 4.20: Pre-Fault DEW Load Flow Values


Pre-Fault Voltage V 0
Pre-Fault Current If0use

6.9476 4.45 kV
9.036 32.34 A

Another associated element that is needed is the thevanin impedance, Zth . This impedance
can be obtained by using the Fault Analysis tool in DEW. The tool indicated that the thevanin
impedance for a fault at Node P4266 is Zth = 4.72666 77.40 Ohms. If we use the pre-fault
voltage and thevanin impedance, we can calculate the current due to the fault itself : If =

V0
Zth .

This leads to the following equation:


If =

6.9476 4.45 kV
= 1469.756 81.85
4.72666 77.40

(4.9)

Using the DEW fault analysis tool, we can easily check this calculation to ensure that it is
valid. Referring to Figure 4.29, we can see that DEW obtains a solution of 1469.24 A observed
at the substation recloser. Again, this solution is the fault current due to the fault itself, and
load currents have not been considered.

Figure 4.29: Reflected Fault Current at the Substation due to the fault itself.

As we previously indicated, during a fault the upstream fuse will observe If f use = If0use +If
where If = 1469.756 81.85 and the pre-fault load flow through the fuse is If0use = 9.036
32.34 . If we sum these values, we will obtain the observed fault current under loaded conditions

107

for the fuse.

If f use = If0use + If = 1469.756 81.85 + 9.036 32.24 = 1475.636 81.59

(4.10)

We can quickly check this result in DEW by applying a fault at Node P4266 under loaded
conditions and select the upstream fuse(FUSE1052482746). DEW obtains a loaded fault solution of If f use = 1473.26 81.54 , as shown in Figure 4.30.

Figure 4.30: Observed Fault Current at upstream fuse(FUSE1052482746) during loaded conditions.

Using load flow data previously given in the table at the beginning of the chapter(Table
4.16), we can calculate the observed fault current at the substation under loaded conditions. To
calculate the observed fault current we sum the current due to the fault itself and the pre-fault
load flow: If sub = I 0 + If . The pre-fault load flow I 0 = 421.96 30.95 and the fault current
due to the fault itself is If = 1469.756 81.85 .

If sub = I 0 + If = 421.96 30.95 + 1469.756 81.85 = 1766.46 71.17

(4.11)

We can quickly verify this result in DEW by applying a fault at P4266 and clicking the
substation recloser to obtain the reflected fault current at the substation. DEW calculates
If sub = 1764.246 71.24 for the observed fault current at the substation during loaded
conditions.
The final value that is needed for the switching model is the bus current Irej . This current

108

is calculated as:
Irej = I 0 If0use = 421.96 30.95 9.036 32.34 = 412.876 30.91

4.7.9

(4.12)

FLA Test Results for P4266 Fault

The fault locator calculates 35 possible fault locations throughout the feeder. In Figure 4.31,
the fault possibilities are shown plotted on the feeder map, each represented by a red circle.

Figure 4.31: Possible Fault Location in Stewart Street 12.47kV Feeder for a fault at P4266.

The FLA identifies Node P4266 as the best matched possibility for the fault location(Figure
4.32). The fault locator also lists Node L1254804(second) and Node P4268(third) as favourable
matches because P4266, L1254804 and P4268 are only 3 power pole spans apart. Since both
nodes are protected by the same fuse, and the three nodes are nearly electrically identical in
short circuit current, the FLA cannot localize any farther. Figure 4.32 shows the best matched
nodes as calculated by the FLA.

109

Figure 4.32: Best Matched Locations for a fault at P4266 as calculated by the FLA.

The worst matches occurred at nodes P4380, P4378, P4375, P4238, P4236, P4235, N7970,
N7955, N7972, N7973, P4240, P4247, and L1254805 which occurred in the zone of the recloser.
The recloser phase and ground curves operate much slower than the 200T fuse characteristic
that protects P4622. Since the recloser also carries more than 400A of load current, the fault
locator ranks these nodes as the worst matched. Figure D.3 in Appendix D.2 shows a few of
the worst matched nodes as calculated by the FLA.
Several other nodes were listed a likely possibilities: P4496, P4498, P4508, P4509, P4513,
P4515, P4520, L1279017, N7974, 1L5936, N7976, N7977, P4241, L1254806, L1254807, L1254808,
L1254809, P4619, and P4622. These possibilities along with their total respective difference
from the calculated best match is shown in Appendix D.2 Table D.2.

110

4.7.10

FLA Test Results Summary for Stewart Street Feeder

Two line-to-ground faults were run on the Stewart Street 12.47kV in DEW to test the FLA.
During both tests, more than 30 fault possibilities throughout the feeder were calculated by the
FLA. The FLA calculated the fault location correctly during both simulations and was able to
localize the fault to within 1 power pole span. Since these are ideal conditions, more testing
is needed with the introduction of measurement and modelling errors.

111

4.8
4.8.1

Sensitivity Analysis: System Load Perturbations


Introduction: Sensitivity Analysis

One of the disadvantages of the proposed FLA is that the fault current solution depends directly
on system loading at the time of the fault. In addition, the localization algorithm directly relies
on the accuracy of the load flow solution. During Test Cases 1-3, we showed that the FLA
was able to correctly identify all fault locations, however, we assumed the load was known and
static. In this section we assume the load to be a gaussian random variable by which the known
load is perturbed about a mean load operation point. By perturbing the load in this fashion,
this introduces uncertainty about system loading conditions.

4.8.2

System Load Perturbations: Test Conditions and Procedures

As we previously described, the uncertainty about system load can be modelled by a gaussian
random variable Xi . To create the forecasted load data, we perturb the actual load table by
the error variable ez . This forms the random variable equation:
Xi = Z ez

(4.13)

For a gaussian distribution, the probability that the load falls within n standard deviations
of the mean(n) is described by: P ( xn < Xi < + xn )
1
P ( xn < Xi < + xn ) =
2

+n

(x)2
2 2

dx

(4.14)

Evaluating the above integral leads to the following table via a error function:
xn

P ( xn < Xi < + xn )

0.6826895

0.9544997

0.9973002

Assuming that 3 are the limits of the probability function(.997 1), the load error is described
in Equation 4.15 below. It was assumed the perturbation error to be 30%, which gives a standard
deviation of = .1.
=

perturbation error
3

(4.15)

To simulate load uncertainty, each loads real and reactive power is assumed to be normally
distributed with a 30% error( = 0.1) with a mean power flow value of 1.0 per-unit( = 1.0).

112

A swing load was chosen in the feeder such that the net system load is the same during each
iteration. The swing load allows all loads in the system to be perturbed while assuring that
the net system load is the same every iteration. By keeping the net power flow the same, we
can test the FLA for every possible load scenario for a known load condition. Therefore, we
can say that if we sum all k buses power flow every iteration, it will be equal to the net system
power P0 , which will remain constant.
P0 =

k
X

Pn

(4.16)

n=1

In Figure 4.33, the histogram of 100 iterations of load perturbations is shown along with
the probability density function. All system loads(except the swing load) are perturbed in this
fashion and the fault locators performance is evaluated. In the next few sections, 100 iterations
or more were performed to determine the FLAs performance under various loading conditions.

Figure 4.33: Normally Distributed Per-Unit Load Power.

113

4.8.3

System Load Perturbations on Test Case 1-No Localization

During Test Case 1(Section 4.3), the proposed FLA was tested with no localization on the NF.
Simple fault tables are used to rank faults from best to worst match. During Test Case 1, the
FLA correctly identified the faulty node during each test. However, the system load conditions
were assumed to be known. In this section, we repeat Test Case 1 with the addition of load
uncertainty. As previously mentioned, the load uncertainty is modelled as a gaussian random
variable by which the known load is perturbed about a mean load operation point.
In this section, we consider several line-to-ground faults at various nodes in the network
with perturbed loading. A line-to-ground fault was tested at the following nodes: Node 10 and
Node 4. A simple MATLAB script was written to run the NF 100 times and record the results.
A 30% load error was assumed( = 0.1) with a mean power flow value of 1.0 per-unit( = 1.0).
The results are presented as percent success and failure over 100 iterations of perturbed loading.
For the location of the faulted nodes, refer to the system one-line Figure 4.2.
Fault at Node 10 with Perturbed Loading
A bolted line to ground fault is placed at Node 10. Exactly 100 iterations are run to evaluate
the FLAs performance under varying load conditions.
Results
Observing the fault table for the NF(Table 4.1), Nodes 10 and 17 are almost indistinguishable
under bolted ground fault conditions at either node. During the 100 iterations, the FLA was
able to correctly identify the fault location as Node 10 67.0% of the time. The remaining 33%,
the faulted node was identified as Node 17 incorrectly.

Table 4.21: Success-Failure Rate on Node 10, Test Case 1(No Localization)
Success/Failure

Percent

Identified Node

Success
Failure

67.00%
33.00%

Node 10 Identified Correctly


Node 17 Identified Incorrectly

114

Fault at Node 17 with Perturbed Loading


A bolted line to ground fault is placed at Node 17. Exactly 100 iterations are run to evaluate
the FLAs performance under varying load conditions.
Results
Observing the fault table for the NF(Table 4.1), Nodes 17 and 10 are almost indistinguishable
under bolted ground fault conditions at either node. During the 100 iterations, the FLA was
able to correctly identify the fault location as Node 17 49.0% of the time. The remaining 51%,
the faulted node was identified as Node 10 incorrectly.

Table 4.22: Success-Failure Rate on Node 17, Test Case 1(No Localization)
Success/Failure

Percent

Identified Node

Success
Failure

49.00%
51.00%

Node 17 Identified Correctly


Node 10 Identified Incorrectly

Fault at Node 4 with Perturbed Loading


A bolted line to ground fault is placed at Node 4. Exactly 100 iterations are run to evaluate
the FLAs performance under varying load conditions.
Results
During the 100 iterations, the FLA was able to correctly identify the fault location as Node 4
100.0% of the time. No other nodes were identified as faulty.

Table 4.23: Success-Failure Rate on Node 4, Test Case 1(No Localization)


Success/Failure

Percent

Identified Node

Success

100.00%

Node 4 Identified Correctly

115

4.8.4

System Load Perturbations on Test Case 2

In Test Case 2(Section 4.4), we introduce the capability of localization into the FLA which relies
on load flow data results from a power flow algorithm. After the occurrence of the fault, the
fault locator will attempt to determine which protective device operated to localize the fault
to a particular zone of protection. To do this, the FLA relies heavily on load flow information
that is used to calculate operating times and load rejection matching. Since measuring live
load flow data at every point in the feeder is impractical(especially in large networks), the
localization algorithm accepts tabulated load flow results from a radial power flow algorithm.
Given that the load flow data is not real-time, there likely exists a difference in the actual and
expected load values at any given point in the network. The adaptability of the FLA to perform
under changing system load conditions is very important for practical use in a live network. To
determine the algorithms flexibility, the localization algorithm was tested in the presence of
perturbing load conditions to simulate load uncertainty.
During Test Case 2, reclosers and substation breakers were inserted into the Notional
Feeder(see Section 4.4 for details). This provided additional information to the fault locator, allowing an exact zone of protection to be pinpointed to localize the fault location. In this
section, we repeat Test Case 2 with the addition of load uncertainty. As previously mentioned,
the load uncertainty is modelled as a gaussian random variable by which the known load is
perturbed about a mean load operation point.
In this section, we consider several line-to-ground faults at various nodes in the network
with perturbed loading. A line-to-ground fault was tested at the following nodes: Node 10 and
Node 17. A simple MATLAB script was written to run the NF 100 times and record the results.
A 30% load error was assumed( = 0.1) with a mean power flow value of 1.0 per-unit( = 1.0).
The results are presented as percent success and failure over 100 iterations of perturbed loading.
For the location of the faulted nodes, refer to the system one-line Figure 4.2.
Fault at Node 10 with Perturbed Loading
A bolted line to ground fault is placed at Node 10. Exactly 100 iterations are run to evaluate
the FLA performance under varying load conditions. Simulation conditions for this fault were
identical to the previous simulation testing in Section 4.8.3.
Results
During the 100 iterations, the FLA was able to correctly identify the fault location as Node 10
100.0% of the time. No other nodes were identified as faulty.

116

Table 4.24: Success-Failure Rate on Load Perturbation Test for Node 10 Fault
Success/Failure

Percent

Identified Node

Success

100.00%

Node 10 Identified Correctly

Fault at Node 17 with Perturbed Loading


A bolted line to ground fault is placed at Node 17. Exactly 100 iterations are run to evaluate
the FLAs performance under varying load conditions. Simulation conditions for this fault were
identical to the previous simulation testing in Test Case 1. Simulation conditions for this fault
were identical to the previous simulation testing in Section 4.8.3.
Results
During the 100 iterations, the FLA was able to correctly identify the fault location as Node 17
100.0% of the time. No other nodes were identified as faulty.

Table 4.25: Success-Failure Rate on Load Perturbation Test for Node 17 Fault
Success/Failure

Percent

Identified Node

Success

100.00%

Node 17 Identified Correctly

117

4.8.5

System Load Perturbations on Test Case 3-Fuse Saving

During Test Case 2, we introduced protective devices such as reclosers and breakers to the
Notional Feeder to protect it. During Test Case 3, we changed this and added fuses to the system
to implement a fuse saving. To implement a common fuse saving scheme, the Notional Feeder
protection system was reconfigured to contain 26 SandC T-Link Fuses and a main reclosing
substation breaker. Pickups and time-dial settings were changed in the substation breaker to
allow fuse saving coordination between all downstream fuses. The reclosing substation feeder
breaker was set to one shot of fast before switching to the slow curve and allowing the fuse to
blow. For more detailed information, see Section 4.5.6.
Given the popularity of fuse saving schemes, further tests were needed to evaluate the
effectiveness of the FLA under various load conditions. To test the FLA on a fuse saving
coordinated feeder, the simulation conditions for Test Case 2 are reproduced again here in Test
Case 3 with the addition of load perturbations. In this section, we consider several line-toground faults at various nodes in the network with perturbed loading. A line-to-ground fault
was tested at the following nodes: Node 10 and Node 17. A simple MATLAB script was written
to run the NF 100 times and record the results. A 30% load error was assumed( = 0.1) with
a mean power flow value of 1.0 per-unit( = 1.0). The results are presented as percent success
and failure over 100 iterations of perturbed loading. For the location of the faulted nodes, refer
to the system one-line Figure 4.2.
Fault at Node 10 with Perturbed Loading
A bolted line to ground fault is placed at Node 10. Exactly 100 iterations are run to evaluate
the FLAs performance under varying load conditions. Simulation conditions for this fault were
identical to the previous simulation testing in Section 4.8.3.
Results
During the 100 iterations, the FLA was able to correctly identify the fault location as Node 10
100.0% of the time. No other nodes were identified as faulty.

Table 4.26: Success-Failure Rate on Load Perturbation Test for Node 10 Fault with Fuse Saving
Feeder Coordination.
Success/Failure

Percent

Identified Node

Success

100.00%

Node 10 Identified Correctly

118

Fault at Node 17 with Perturbed Loading


A bolted line to ground fault is placed at Node 17. Exactly 100 iterations are run to evaluate
the FLAs performance under varying load conditions. Simulation conditions for this fault were
identical to the previous simulation testing in Section 4.8.3.
Results
During the 100 iterations, the FLA was able to correctly identify the fault location as Node 17
100.0% of the time. No other nodes were identified as faulty.

Table 4.27: Success-Failure Rate on Load Perturbation Test for Node 17 Fault with Fuse Saving
Feeder Coordination.
Success/Failure

Percent

Identified Node

Success

100.00%

Node 17 Identified Correctly

119

4.8.6

System Load Perturbations on Test Case 3-Fuse Blowing

During Test Case 2, we introduced protective devices such as reclosers and breakers to the
Notional Feeder to protect it. During Test Case 3, we changed this and added fuses to the
system to implement a fuse saving. To implement a common fuse saving scheme, the Notional
Feeder protection system was reconfigured to contain 26 SandC T-Link Fuses and a main
reclosing substation breaker. Pickups and time-dial settings were changed in the substation
breaker to allow fuse blowing coordination between all downstream fuses. For more detailed
information, see Section 4.5.
Given the popularity of fuse saving schemes, further tests were needed to evaluate the
effectiveness of the FLA under various load conditions. To test the FLA on a fuse saving
coordinated feeder, the simulation conditions for Test Case 2 are reproduced again here in Test
Case 3 with the addition of load perturbations. In this section, we consider several line-toground faults at various nodes in the network with perturbed loading. A line-to-ground fault
was tested at the following nodes: Node 10 and Node 17. A simple MATLAB script was written
to run the NF 100 times and record the results. A 30% load error was assumed( = 0.1) with
a mean power flow value of 1.0 per-unit( = 1.0). The results are presented as percent success
and failure over 100 iterations of perturbed loading. For the location of the faulted nodes, refer
to the system one-line Figure 4.2.
Fault at Node 10 with Perturbed Loading
A bolted line to ground fault is placed at Node 10. Exactly 100 iterations are run to evaluate
the FLAs performance under varying load conditions. Simulation conditions for this fault were
identical to the previous simulation testing in Section 4.8.3.
Results
During the 100 iterations, the FLA was able to correctly identify the fault location as Node 10
100.0% of the time. No other nodes were identified as faulty.

Table 4.28: Success-Failure Rate on Load Perturbation Test for Node 10 Fault with Fuse
Blowing Feeder Coordination.
Success/Failure

Percent

Identified Node

Success

100.00%

Node 10 Identified Correctly

120

Fault at Node 17 with Perturbed Loading


A bolted line to ground fault is placed at Node 17. Exactly 100 iterations are run to evaluate
the FLAs performance under varying load conditions. Simulation conditions for this fault were
identical to the previous simulation testing in Section 4.8.3.
Results
During the 100 iterations, the FLA was able to correctly identify the fault location as Node 17
100.0% of the time. No other nodes were identified as faulty.

Table 4.29: Success-Failure Rate on Load Perturbation Test for Node 17 Fault with Fuse
Blowing Feeder Coordination.
Success/Failure

Percent

Identified Node

Success

100.00%

Node 17 Identified Correctly

121

Chapter 5

FLA Testing Conclusions


A total of four test cases were proposed to test the FLA using models in MATLAB and DEW
software packages. During Test Cases 1-3, a Progress Energy modelled feeder in MATLAB was
used to test the FLA under various conditions(called the Notional Feeder). During Test Case 4,
the proposed FLA was tested on a large scale feeder(Stewart Street 12.47kV Feeder) modelled
in DEW by Allegheny Power. It was noted that during Test Cases 1-4, system loading was
considered to be known and static. Since this is not always the case in practice, a final test
case was proposed with the addition of perturbed loading to introduce load uncertainty.
During Test Case 1, the FLA was tested on the NF containing no protective devices with no
localization. Test results showed that the FLA correctly identified the faulty node for each
test during Test Case 1. However, test data showed that several nodes were nearly identical
in short circuit current, i.e. Node 17 Phase-to-Ground fault test(Section 4.3.4). This test case
illustrated the importance of the localization algorithm to accurately determine the true fault
location if there are multiple possibilities.
In Test Case 2, recloser and substation breaker models were added to the NF. Test results
during this test case showed that the localization algorithm can effectively resolve multiple possibilities into one certain fault location. However, it was discovered that the FLA localization
algorithm does have limitations. During a ground fault at Node 16(see Section 4.4.5 for details),
the localization algorithm was unable to rank multiple possibilities. It was concluded that the
localization algorithm does not provide any additional ranking information when multiple possibilities exist in the same zone of protection.
In Test Case 3, the NF protection system was reconfigured to contain 26 SandC T-Link fuses
and a main reclosing substation breaker to implement a fuse saving and fuse blowing protec-

122

tion scheme. During Test Case 3, the FLA correctly identified the faulty node for each test,
except Node 8. While testing a fault at Node 8(see Section 4.5.11 for details), another inherent
drawback was noted in the proposed FLA. In fuse blowing and fuse saving schemes, high speed
fuses(N, QR and K) can melt before the current reaches a steady state value. This phenomena
was observed in the NF model at Node 8, which is protected by a 50T fuse. Consequently, the
fault locator failed to determine the location of the fault. Therefore, it was noted that high
speed fuses may present a problem when fault locating with the proposed FLA.
During Test Cases 1, 2 and 3, the Notional Feeder system model only consisted of 18 nodes,
which is much smaller than most conventional feeders. In Test Case 4, a large scale feeder
model was provided by Allegheny Power for FLA testing. The DEW feeder model consisted of
more than 250 faulty node possibilities. During testing, the FLA correctly ranked and localized
during all tests. However, multiple fault possibilities did occur in the same zone of protection
during each test included in Test Case 4. This same drawback was noted during Test Case 2.
Final testing of the FLA was performed again using NF model. In Test Cases 1-4, the load was
assumed to be know and static, which is rarely the case in practice. Using the NF model, the
system loads were perturbed to introduce load uncertainty. Test data showed that the localization algorithm provided a significant increased performance for faults at certain nodes in the NF.
It was concluded that the proposed FLA, with the addition of localization, is not sensitive to
small load changes. However, large errors in load flow modelling can result in incorrect ranking.

123

REFERENCES
[1] M. M. Saha, J. J. Izykowski and E. Rosolowski. Fault location on Power Networks. pp.
333-360. 2010.
[2] M. Dilek, R. Broadwater and R. Sequin, Computing distribution system fault currents and
voltages via numerically computed Thevenin equivalents and sensitivity matrices, Power
Systems Conference and Exposition, 2004. IEEE PES, vol.1, pp. 244, 2004.
[3] J. J. Grainger, Power System Analysis, pp. 157. 1994.
[4] W. F. Tinney, Compensation Methods for Network Solutions by Optimally Ordered Triangular Factorization, Power Apparatus and Systems, IEEE Transactions on (Volume:PAS91 , Issue: 1 ), 1972.
[5] G. Gross and H. W. Hong, A Two-Step Compensation Method for Solving Short Circuit
Problems, Apparatus and Systems, IEEE Transactions on (Volume:PAS-101 , Issue: 6 ),
1982.
[6] R. C. Dorf , The Engineering Handbook, pp. 112.7-112.10, 2004.
[7] J. J. Burke and D. J. Lawre, , Characteristics of Fault Currents on Distribution Systems,
Power Apparatus and Systems, IEEE Transactions on (Volume:PAS-103 , Issue: 1 ), 1984.
[8] I. D. Serna-Suarez, C. D. Ferreira-Sequeda, S. A. Martnez-Gutierrez, M. F. Suarez-Sanchez
and G. Carrillo-Caicedo, Impact of static load models on the power distribution fault location problem , Transmission and Distribution Conference and Exposition: Latin America
(TandD-LA), 2010 IEEE/PES, 2010.
[9] N. Karnik, S. Das, S. Kulkarni and S. Santoso, Effect of load current on fault location
estimates of impedance-based methods, Power and Energy Society General Meeting, 2011
IEEE, 2011.
[10] K. Zimmerman and D. Costello, Impedance-Based Fault Location Experience, Rural
Electric Power Conference, 2006 IEEE, 2006.
[11] P. Gill, Electrical Power Equipment Maintenance and Testing, pp. 544-545, 2008.
[12] (7/1/2012). Discrete Fourier Transform. http : //en.wikipedia.org
[13] K. Jinsang, M. E. Baran and G. C. Lampley, Estimation of Fault Location on Distribution
Feeders using PQ Monitoring Data, Power Engineering Society General Meeting. IEEE,
2007.
[14] T. A. Short, D. D. Sabin and M. F. McGranaghan, Using PQ Monitoring and Substation
Relays for Fault Location on Distribution Systems, Rural Electric Power Conference,
IEEE, 2007.

124

[15] Electric Power Research Institute, Distribution Fault Current Analysis, EPRI 1209-1,
1983.
[16] H. L. Willis, Power Distribution Planning Reference Book, pp. 51-53, 2004.
[17] T. A. Short, Electric Power Distribution Equipment and Systems, pp. 252-253, 2004.
[18] J. Machowski, J. Bialek and J. Bumby, Power System Dynamics: Stability and Control,
pp. 3.4.4.2-3.4.4.3, 2008.
[19] Schweitzer Engineering Laboratories, SEL-351S Protection System Instruction Manual,
pp. 354-355, 2013.
[20] Schweitzer Engineering Laboratories, SEL-651R Protection System Instruction Manual,
pp. 413-414, 2012.
[21] S&C, S&C T-Speed Positrol Fuse Links Time-Current Characteristic Curves, 1984.
[22] M. Dilek, R. Broadwater and R. Sequin , Calculating short-circuit currents in distribution
systems via numerically computed Thevenin equivalents, Transmission and Distribution
Conference and Exposition, IEEE PES (Volume:3 ), 2003.
[23] Schweitzer Engineering Laboratories, SELU: PROT 401 Course Handbook, pp. 18, 2013.
[24] J. OMalley, Schaums Outline of Basic Circuit Analysis, pp. 295-296, 1992.
[25] R. Das, Determining the Locations of Faults in Distribution Systems, 1998.
[26] A. A. Girgis, C. M. Fallon and D. L. Lubkeman, A Fault Location Technique for Rural
Distribution Feeders, Industry Applications, IEEE Transactions on (Volume:29 , Issue: 6
), 1993.
[27] J. S. Jung, Branch Current State Estimation Method for Power Distribution Systems,
2009.
[28] MathWorks, MATLAB 2010A Help File, 2010.
[29] U. Rao, Computer Techniques And Models In Power Systems, pp. 444-447, 2008.
[30] E. O. Schweitzer, A Review of Impedance-Based Fault Locating Experience,
Nebraska System Protection Seminar, 1990.

Iowa-

[31] W. Chen, Active Network Analysis, pp. 284-285, 1991.


[32] M. M. Saha, F. Provoost and E. Rosolowski, Fault location method for MV cable network
Developments in Power System Protection, Seventh International Conference, 2001.

125

APPENDICES

126

Appendix A

Notional Feeder Load Flow Data


A.1

Notional Feeder Real Power Flow Data


Bus

Power Phase A

Power Phase B

Power Phase C

Voltage

Load 1

57.7500 kW

165.5500 kW

112.4200 kW

12.0 kV

Load 2

651.4200 kW

171.7100 kW

100.1000 kW

12.0 kV

Load 3

544.3900 kW

56.9800 kW

254.1000 kW

12.0 kV

Load 4

145.5300 kW

112.4200 kW

245.6300 kW

12.0 kV

Load 5

597.5200 kW

510.5100 kW

733.8100 kW

12.0 kV

Load 6

535.1500 kW

530.5300 kW

520.5200 kW

12.0 kV

Load 7

532.0700 kW

599.8300 kW

338.8000 kW

12.0 kV

Load 8

14.6300 kW

12.0 kV

Load 9

20.0200 kW

10.0100 kW

85.4700 kW

12.0 kV

Load 10

13.0000 kW

0.240 kV

Load 11

26.0000 kW

0.240 kV

Load 12

13.0000 kW

0.240 kV

Load 13

19.0000 kW

0.240 kV

Load 14

25.0000 kW

0.240 kV

Load 15

26.0000 kW

0.240 kV

Load 16

13.0000 kW

0.240 kV

Load 17

26.0000 kW

0.240 kV

Load 18

13.0000 kW

0.240 kV

Load 19

13.0000 kW

0.240 kV

Load 20

13.0000 kW

0.240 kV

Load 21
..
.

..
.

19.0000 kW
..
.

..
.

0.240 kV
..
.

127

Bus

Power Phase A

Power Phase B

Power Phase C

Voltage

Load 22

19.0000 kW

0.240 kV

Load 23

19.0000 kW

0.240 kV

Load 24

19.0000 kW

0.240 kV

Load 25

32.7250 kW

25.5640 kW

50.7430 kW

12.0 kV

Load 26

455.8400 kW

715.1760 kW

841.8410 kW

12.0 kV

Load 27

15.8620 kW

12.0 kV

Load 28

235.0810 kW

207.4380 kW

196.1190 kW

12.0 kV

Load 29

319.3960 kW

317.8560 kW

95.3260 kW

12.0 kV

Load 30

357.6650 kW

207.2070 kW

59.5980 kW

12.0 kV

Load 31

499.5760 kW

424.4240 kW

846.4610 kW

12.0 kV

Load 32

168.9380 kW

608.8390 kW

227.2270 kW

12.0 kV

Total Load

5168.933 kW

4940.044 kW

4722.795 kW

=14831.772 kW

128

A.2

Notional Feeder Reactive Power Flow Data


Bus

Power Phase A

Power Phase B

Power Phase C

Voltage

Load 1

81.6200 kVAR

110.1100 kVAR

66.9900 kVAR

12.0 kV

Load 2

332.6400 kVAR

80.0800 kVAR

56.9800 kVAR

12.0 kV

Load 3

278.7400 kVAR

29.2600 kVAR

120.1200 kVAR

12.0 kV

Load 4

74.6900 kVAR

57.7500 kVAR

120.1200 kVAR

12.0 kV

Load 5

304.9200 kVAR

253.3300 kVAR

386.5400 kVAR

12.0 kV

Load 6

272.5800 kVAR

261.8000 kVAR

264.8800 kVAR

12.0 kV

Load 7

273.3500 kVAR

297.9900 kVAR

173.2500 kVAR

12.0 kV

Load 8

7.7000 kVAR

12.0 kV

Load 9

10.0100 kVAR

1.5400 kVAR

43.1200 kVAR

12.0 kV

Load 10

7.0000 kVAR

0.240 kV

Load 11

13.0000 kVAR

0.240 kV

Load 12

6.0000 kVAR

0.240 kV

Load 13

9.0000 kVAR

0.240 kV

Load 14

14.0000 kVAR

0.240 kV

Load 15

12.0000 kVAR

0.240 kV

Load 16

6.0000 kVAR

0.240 kV

Load 17

14.0000 kVAR

0.240 kV

Load 19

7.0000 kVAR

0.240 kV

Load 20

6.0000 kVAR

0.240 kV

Load 21

6.0000 kVAR

0.240 kV

Load 22

19.0000 kVAR

0.240 kV

Load 23

19.0000 kVAR

0.240 kV

Load 24

19.0000 kVAR

0.240 kV

Load 25

32.7250 kVAR

25.5640 kVAR

50.7430 kVAR

12.0 kV

Load 26

455.8400 kVAR

715.1760 kVAR

841.8410 kVAR

12.0 kV

Load 27

15.8620 kVAR

12.0 kV

Load 28

235.0810 kVAR

207.4380 kVAR

196.1190 kVAR

12.0 kV

Load 29

319.3960 kVAR

317.8560 kVAR

95.3260 kVAR

12.0 kV

Load 30

357.6650 kVAR

207.2070 kVAR

59.5980 kVAR

12.0 kV

Load 31

499.5760 kVAR

424.4240 kVAR

846.4610 kVAR

12.0 kV

Load 32

168.9380 kVAR

608.8390 kVAR

227.2270 kVAR

12.0 kV

Total Load

2569.644 kVAR

2384.478 kVAR

2248.4 kVAR

=7202.522 kVAR

129

A.3

12kV Capacitor Bank Data


Bus

Power Phase A

Power Phase B

Power Phase C

Voltage

Capacitor Bank 1

400 kVAR

400 kVAR

400 kVAR

12 kV

Capacitor Bank 2

200 kVAR

200 kVAR

200 kVAR

12 kV

Capacitor Bank 3

400 kVAR

400 kVAR

400 kVAR

12 kV

Capacitor Bank 4

400 kVAR

400 kVAR

400 kVAR

12 kV

Total

1400 kVAR

1400 kVAR

1400 kVAR

=4200 kVAR

A.4

Transformer Bank Data


Transformer

MVA Rating

Winding 1 Imp.

Winding Imp. 2

69kV/12kV(-Wye)

20 MVA

Z1 =0+0.055j pu

Z2 =0+0.005j pu

12kV/240V(Single Phase)

262kVA

Z1 =0+0.0192j pu

Z2 =Z3 =0+0.0048j pu

A.5

Source Impedance Data


Source

Source Resistance Rs

Source Reactance jXs

Source 1 69kV

0.0075 Ohms

0.0582j Ohms

130

A.6
A.6.1

Line Impedance Data


Positive Sequence Line Impedance Data
(1)

(1)

Line

Resistance RLine

Inductance jXLine

Length(km)

Line 1

0.1502

0.6309

7.1410

Line 2

0.0618

0.2597

2.9390

Line 3

0.0546

0.2294

2.5960

Line 4

0.0366

0.1536

1.7390

Line 5

0.0851

0.3575

4.0470

Line 6

0.0978

0.4106

4.6470

Line 7

0.0919

0.3861

4.3700

Line 8

0.0075

0.0315

0.3560

Line 9

0.0261

0.1095

1.2390

Line 10

0.0538

0.0066

0.2030

Line 11

0.1196

0.0146

0.4510

Line 12

0.0859

0.0105

0.3240

Line 13

0.1965

0.0240

0.7410

Line 14

0.0920

0.0113

0.3470

Line 15

0.2259

0.0276

0.8520

Line 16

0.1254

0.0153

0.4730

Line 17

0.0780

0.0095

0.2940

Line 18

0.0260

0.0032

0.0980

Line 19

0.0838

0.0103

0.3160

Line 20

0.1175

0.0144

0.4430

Line 21

0.1188

0.0145

0.4480

Line 22

0.0949

0.0116

0.3580

Line 23

0.3065

0.0375

1.1560

Line 24

0.0939

0.0115

0.3540

Line 25

0.0711

0.2984

3.3780

Line 26

0.0476

0.1998

2.2610

Line 27

0.0050

0.0211

0.2390

Line 28

0.0439

0.1846

2.0890

Line 29

0.1197

0.5028

5.6913

Line 30

0.0593

0.2491

2.8190

Line 31

0.0186

0.0780

0.8830

Line 32

0.0667

0.2801

3.1700

131

A.6.2

Negative Sequence Line Impedance Data


(2)

(2)

Line

Resistance RLine

Inductance jXLine

Length(km)

Line 1

0.1502

0.6309

7.1410

Line 2

0.0618

0.2597

2.9390

Line 3

0.0546

0.2294

2.5960

Line 4

0.0366

0.1536

1.7390

Line 5

0.0851

0.3575

4.0470

Line 6

0.0978

0.4106

4.6470

Line 7

0.0919

0.3861

4.3700

Line 8

0.0075

0.0315

0.3560

Line 9

0.0261

0.1095

1.2390

Line 10

0.0538

0.0066

0.2030

Line 11

0.1196

0.0146

0.4510

Line 12

0.0859

0.0105

0.3240

Line 13

0.1965

0.0240

0.7410

Line 14

0.0920

0.0113

0.3470

Line 15

0.2259

0.0276

0.8520

Line 16

0.1254

0.0153

0.4730

Line 17

0.0780

0.0095

0.2940

Line 18

0.0260

0.0032

0.0980

Line 19

0.0838

0.0103

0.3160

Line 20

0.1175

0.0144

0.4430

Line 21

0.1188

0.0145

0.4480

Line 22

0.0949

0.0116

0.3580

Line 23

0.3065

0.0375

1.1560

Line 24

0.0939

0.0115

0.3540

Line 25

0.0711

0.2984

3.3780

Line 26

0.0476

0.1998

2.2610

Line 27

0.0050

0.0211

0.2390

Line 28

0.0439

0.1846

2.0890

Line 29

0.1197

0.5028

5.6913

Line 30

0.0593

0.2491

2.8190

Line 31

0.0186

0.0780

0.8830

Line 32

0.0667

0.2801

3.1700

132

A.6.3

Zero Sequence Line Impedance Data


(0)

(0)

Line

Resistance RLine

Inductance jXLine

Length(km)

Line 1

0.8004

2.5107

7.1410

Line 2

0.3294

1.0333

2.9390

Line 3

0.2910

0.9127

2.5960

Line 4

0.1949

0.6114

1.7390

Line 5

0.4536

1.4229

4.0470

Line 6

0.5209

1.6338

4.6470

Line 7

0.4898

1.5365

4.3700

Line 8

0.0399

0.1252

0.3560

Line 9

0.1389

0.4356

1.2390

Line 10

0.1238

0.0838

0.2030

Line 11

0.2751

0.1861

0.4510

Line 12

0.1976

0.1337

0.3240

Line 13

0.4520

0.3057

0.7410

Line 14

0.2117

0.1432

0.3470

Line 15

0.5197

0.3515

0.8520

Line 16

0.2885

0.1951

0.4730

Line 17

0.1793

0.1213

0.2940

Line 18

0.0598

0.0404

0.0980

Line 19

0.1928

0.1304

0.3160

Line 20

0.2702

0.1828

0.4430

Line 21

0.2733

0.1848

0.4480

Line 22

0.2184

0.1477

0.3580

Line 23

0.7052

0.4769

1.1560

Line 24

0.2160

0.1460

0.3540

Line 25

0.3786

1.1877

3.3780

Line 26

0.2534

0.7949

2.2610

Line 27

0.0268

0.0840

0.2390

Line 28

0.2342

0.7345

2.0890

Line 29

0.6379

2.0010

5.6913

Line 30

0.3160

0.9911

2.8190

Line 31

0.0990

0.3105

0.8830

Line 32

0.3553

1.1145

3.1700

133

A.6.4

Positive and Zero Sequence Shunt Capacitance Data


(1)

(0)

Line

CLine (F)

CLine (F)

Length(km)

Line 1

1015

1015

7.1410

Line 2

1015

1015

2.9390

Line 3

1015

1015

2.5960

Line 4

1015

1015

1.7390

Line 5

1015

1015

4.0470

Line 6

1015

1015

4.6470

Line 7

1015

1015

4.3700

Line 8

1015

1015

0.3560

Line 9

1015

1015

1.2390

Line 10

1015

1015

0.2030

Line 11

1015

1015

0.4510

Line 12

1015

1015

0.3240

Line 13

1015

1015

0.7410

Line 14

1015

1015

0.3470

Line 15

1015

1015

0.8520

Line 16

1015

1015

0.4730

Line 17

1015

1015

0.2940

Line 18

1015

1015

0.0980

Line 19

1015

1015

0.3160

Line 20

1015

1015

0.4430

Line 21

1015

1015

0.4480

Line 22

1015

1015

0.3580

Line 23

1015

1015

1.1560

Line 24

1015

1015

0.3540

Line 25

1015

1015

3.3780

Line 26

1015

1015

2.2610

Line 27

1015

1015

0.2390

Line 28

1015

1015

2.0890

Line 29

1015

1015

5.6913

Line 30

1015

1015

2.8190

Line 31

1015

1015

0.8830

Line 32

1015

1015

3.1700

134

A.7

Relay Settings and Fuse Characteristics

Table A.1: Relay Settings for Substation Breaker A and Substation Breaker B
Operation Number(Type)
1(Phase Relay)
2(Phase Relay)
1(Ground Relay)
2(Ground Relay)

Pickup
1440 Amps
720 Amps
480 Amps
360 Amps

Curve
SEL-351S-6
SEL-351S-6
SEL-351S-6
SEL-351S-6

U5
U3
U5
U3

Time-Dial
Curve
Curve
Curve
Curve

TD=1.1
TD=2.3
TD=1.8
TD=4.0

Table A.2: Relay Settings for Recloser A and Recloser B


Operation Number(Type)
1(Phase Relay)
2(Phase Relay)
1(Ground Relay)
2(Ground Relay)

Pickup

Curve

1200 Amps
600 Amps
480 Amps
240 Amps

ABB PCD 2000 STI


ABB PCD 2000 EI
ABB PCD 2000 STI
ABB PCD 2000 EI

135

Time-Dial
TD=0.9
TD=1.0
TD=1.3
TD=3.0

A.8

Fuse Characteristics

Figure A.1: SandC T-Speed Fuse Minumum Melt Characteristics.

136

Figure A.2: SandC T-Speed Fuse Total Clearing Time Characteristics.

137

Appendix B

Modelling of the Notional Feeder


using MATLAB
B.1

Introduction

To test the proposed FLA, a PEC(Progress Energy Carolinas) example feeder was used(called
the Notional Feeder). All parameters such as line impedance, transformer ratings and load
data were provided and modelled in the MATLAB simulink software package. MATLAB was
selected due to its capability to perform discrete simulations and packaged detailed models of
transformers, lines and exponential loads. In this chapter, careful attention was given to the
modelling of the system to provide the most accurate system model for testing the FLA.

B.2

Modelling of Lines

The modelling of lines in the Notional Feeder is done via the nominal PI line model. This line
model is often used for short to medium length lines and includes a line inductance, resistance
and capacitance as lumped parameters. When the line becomes appreciably long, this modelling
introduces known errors. Since the distribution system contains short line sections, it was
assumed that this model would suffice. Figure B.1 shows the nominal PI model used.
In Figure B.1 the terms Rs , Rm and Ls ,Lm represent the self and mutual resistances and
inductances due to mutual coupling between each of the three conductors. Phase and ground
capacitances are also represented by Cp and Cg respectively. The RLC terms(Rs , Rm ,Ls ,Lm ,Cp
and Cg ) are calculated from positive, negative and zero sequence impedance parameters of the
line. Since MATLAB uses sequence components to calculate RLC terms, it must be assumed
that the mutual coupling and self impedance terms are all equal. This means all the mutual
impedances i.e., Zab = Zba and all the self-impedances Zaa = Zbb = Zcc are equal. Therefore,

138

Figure B.1: Nominal PI Line Model[28].

this reduces the line impedance matrix Zabc to the matrix shown in Equation B.5.
And the sequence impedance matrix for the line segment between the two ends of Figure
B.1:

Z00 Z01 Z02

Z012 = Z10 Z11 Z12


Z20 Z21 Z22

(B.1)

If we assume that all self and mutual terms are equal:

Z0

0
Z+

Z012 = 0
0

0
Z

(B.2)

Where Z00 = Z 0 , Z11 = Z + and Z22 = Z . We also assume that the negative and positive
sequence impedances are equal for this line segment: Z + = Z
We can transform the sequence impedance matrix to the phase impedance matrix using the
following transform:

Z0

1
Z012 = T 0
3
0

0
Z+
0

139

0 3T 1
Z+

(B.3)

Where T is:

T = 1 2
1

(B.4)

Where 2 = 16 120 . This yields the following result:

Zabc =

2Z (1)
Z (0)
3 + 3
Z (0)
Z (1)
3 3
Z (0)
Z (1)
3 3

Z (0)
Z (1)
3 3
2Z (1)
Z (0)
3 + 3
Z (0)
Z (1)
3 3

Z (0)
Z (1)
3 3
Z (0)
Z (1)
3 3
2Z (1)
Z (0)
3 + 3

Zaa Zab Zac


= Zba
Zca

Zbb
Zcb

Zbc
Zcc

(B.5)

Using the above matrix, MATLAB calculates the RLC parameters as follows:
Rs =

2R(1) R(0)
+
3
3

(B.6)

Ls =

2L(1) L(0)
+
3
3

(B.7)

Rm =

R(0) R(1)

3
3

(B.8)

Lm =

L(0) L(1)

3
3

(B.9)

The line capacitances can also be derived as:


Cp = C (1)

(B.10)

3C (1) C (0)
(B.11)
C (1) C (0)
In MATLAB Simulink each line is modelled using the Three Phase Nominal PI Line Section
Cg =

Block show in Figure B.2.

Figure B.2: Line Model Block in MATLAB Simulink[28].

140

B.3

Modelling of System Loads

In the NF, all loads are modelled as static loads. Each load is described by a exponential
function that directly varies the power flow as a function of voltage. Each composite load is
described by the following power flow function:

P (V, p ) = P0

Q(V, q ) = Q0

V
V0

p

V
V0

 q

(B.12)

(B.13)

The above equations represent the consumed load at various voltages based on a nominal
power and voltage. The terms P0 and Q0 represent the real and reactive power consumed a
voltage V0 . The load voltage during operating conditions is represented by V . The terms p
and q terms represent the real and reactive power sensitivity to changes in voltage. If we solve
for the terms p and q using a relative sensitivity function we get:
p =

P
P0
V
V0

(B.14)

q =

Q
Q0
V
V0

(B.15)

We can see intuitively that the coefficients p and q represent the normalized load sensitivity to voltage. Each of these terms represent an important factor in modelling consumer
composite loads. As we approach p = q = 0 the load power consumption becomes less dependent on voltage and acts a constant power load. If we to approach p = q = 1, this would
cause the load power flow to directly depend on voltage, thereby acting as a constant current
load. When the terms p = q = 2, this causes the power flow to directly depend on the square
of the voltage, thereby acting as a constant impedance load.
Utilities and power research institutes have dedicated significant research to modelling typical consumer loads. Software to model these loads and implement them in algorithms are
relatively inexpensive compared to obtaining real-time data. Modern protective relays and meters have become much cheaper, but the cost of real-time recording remains a significant cost.
Usually most utilities estimate feeder load data using the following rules suggested by [16]:
Industrial areas are most typically closest to constant power loads (about 80/20 constant
power/constant impedance).

141

Residential areas with very strong summer-peaking loads usually have a 70/30 split of
constant power/constant impedance loads.
Winter-peaking residential areas or those summer-peaking areas with low summer load
densities per household (little air conditioning), tend to have a mix closer to 30/70 constant power/constant impedance loads.
Urban areas are usually about 50/50 constant power/constant impedance loads. A 50/50
mixture of power and impedance loads looks very much like and can be modelled as
constant current.
In rural and residential areas of developing countries, and commercial areas in economically depressed regions, loads generally tend to be about 20/80 constant power/constant
impedance.
It was assumed that the NF load could be represented as a urban load, representing 50/50
constant power/constant impedance loads. As indicated by [16], a 50/50 mix of constant
power(motor loads) and constant impedance loads(resistive heating) can be approximated as
a constant current load. Therefore, under un-faulted conditions, the loads in the NF are
represented as constant current loads.

B.3.1

Load Modelling Under Fault Conditions

The load models given above in Equations B.12 and B.13 do not represent correct load behaviour under rapid voltage drop(fault) conditions[18]. Therefore, under faulted conditions, it
is customary to approach load modelling as a piecewise function[29]. During voltage depressions, the constant current load model is switched to constant impedance. As recommended by
[18], if the voltage is below 0.7 pu, the load models given above in Equations B.12 and B.13 are
no longer valid and the load is then switched to constant impedance.

= 1 if V
p
load
P (V, p ) =
= 2 if V
p
load

= 1 if V
q
load
Q(V, p ) =
= 2 if V
q
load

>0.7 pu

(B.16)

<0.7 pu
>0.7 pu

(B.17)

<0.7 pu

The equations above are used in the 3-Phase Dynamic Load model in MATLAB Simulink.
The block representing each load is shown in Figure B.3.

142

Figure B.3: 3-Phase Dynamic Load Representing Exponential Load Functions.

B.4

Modelling of Sources

The system source is modelled a thevanin equivalent. The thevanin equivalent was provided by
PEC and is represented by the thevanin equivalent source impedance Zs and thevanin voltage
Vs :

Figure B.4: System Thevanin Equivalent.

To model this, a simple three phase source is used in MATLAB containing a series resistance
and series inductance to model a thevanin equivalent of the system. The block configuration is
show in Figure B.5.

B.5

Modelling of Capacitor Banks

The notional feeder consists of a total of four 12kV capacitor banks. Each capacitor bank is
modelled as a constant impedance load with p = q = 2 in the exponential load equation

143

Figure B.5: Three Phase MATLAB Simulink Source.

proposed in previous sections. Reactive power flow is proportional to the square of the voltage
applied to the bank. The bank is represented by the exponential load as follows:
Q(V, 2) =

Q0
V = X0 V 2
V02

(B.18)

Figure B.6: 12kV Capacitor Bank.

B.6

Modelling of Multi-Winding Transformers

All multi-winding transformers used in the notional feeder are modelled as show in Figure B.9.
Each of the transformers serve two 150kVA loads. The transformer winding leakage inductance
and resistance is specified for each winding as [L1 , L2 , L3 ] and [R1 , R2 , R3 ], respectively. The
nominal winding voltages must also be specified as [U1 , U2 , U3 ] in addition to the system nominal

144

frequency fn and power rating Sn . Each multi-winding transformer is modelled with a nonsaturable core therefore the magnetization inductance and resistance Lm , Rm must be specified.
All parameters can be found in Appendix A.

Figure B.7: Three Winding Transformer.

Figure B.8: MATLAB Multi-Winding Transformer Block.

B.7

Modelling of Feeder Transformer

The three phase feeder transformer used in the notional feeder are modelled as show in Figure B.9. This transformer steps 69kV sub-transmission down to 12kV for distribution. The
transformer winding leakage reactance and resistance is specified for each winding as [X1 , X2 ]
and [R1 , R2 ], respectively. The nominal winding voltages must also be specified as [V1 , V2 ] in
addition to the system nominal frequency fn and 3-phase power rating Sn . Each multi-winding
transformer is modelled with a non-saturable core therefore the magnetization reactance and
resistance Xm , Rm must be specified. All parameters can be found in Appendix A.

145

Figure B.9: Per Phase Representation of 3-Phase Feeder Transformer.

Figure B.10: MATLAB Three Phase Transformer.

146

Appendix C

Distribution Fault Analysis


Introduction to Short Circuit Analysis There are a variety of causes for faults occurring in distribution networks. Over-head lines are often exposed to environmental interference
such as trees, lightning, and human intervention. Fault studies are often preformed to better
understand how the system responds to environmental interference as well as protection of the
system. Large scale distribution systems pose a unique set of challenges when preforming short
circuit studies. It is often the case that these systems are very complex, containing thousands
of customers and tens-of-thousands of nodes.
Most classical short circuit programs using symmetrical components work well for balanced
3-phase systems. The main computational advantage for symmetrical components exists since
the positive, negative and zero sequence networks can be considered separately. However, when
unequal self or mutual impedances are considered, symmetrical components loose many of their
advantages. Unequal self or mutual impedances create sequence component coupling. For this
reason, short circuit analysis based on 3-phase models are more attractive. The following sections describe a fault analysis technique based on numerically calculated thevanin equivalents
in the phase frame of reference.

C.1

System Fault Currents and Voltages via Numerically Computed Thevenin Equivalents and Sensitivity Matrices

There are several assumptions about the network that must be made before preforming any
analysis. The following assumptions are presumed to be true[6]:
Circuit is comprised of linear components only; this is comprised of inductive, capacitive
and resistive elements.
Slow acting control devices are considered to be frozen. This means that that all actively

147

controlled devices such as tap-changers or protective relays are in the same state at t = 0
as t = 0+ .
The substation bus is replaced by a thevanin equivalent. The thevanin equivalent component model is shown in Figure C.1.

Figure C.1: System Thevanin Equivalent Component Model.

Load currents do not change during the fault.

DEWs algorithm for calculation of the loaded fault current can be broken down into several
major steps:
1. Obtain all pre-fault voltages and currents. This requires a 3-phase power flow algorithm
and detailed knowledge of the network(line parameters, load parameters,ect.)
2. Formation of the phase thevanin matrix.
3. Calculation of the fault voltages and currents.
4. Summation of pre-fault and fault currents using superposition.

C.1.1

Thevanins Theorem and Superposition Principle

We begin fault analysis by considering a typical power system shown below consisting of a
source, loads, and a bus F .
The basic power system fault can be modelled as a simple switch in series with a fault
impedance. When the switch is in the closed position, the power system is considered to be
faulted and fault current flows through the switch path. This is shown below in C.3.

148

Figure C.2: Power System containing a source, load buses and bus F .

Figure C.3: Power System before a fault occurrence at bus F.

If we measure the voltage across the switch in the open position, the voltage will be equal to
the pre-fault bus voltage. If we replace the open switch by a voltage source equal in magnitude
and angle as the pre-fault voltage, this will have no effect on the circuit. Therefore, we can
model the open position of the switch by inserting a voltage source equal in magnitude and
angle as the pre-fault voltage. This is show graphically in Figure C.4.
If we want to represent a fault at bus f , we simply close the switch. We can represent the
switch in the closed position by inserting a second voltage source equal in magnitude, but 180
degrees out of phase with the pre-fault voltage. This results in a voltage drop across the two
sources as 0, yielding a closed switch. This is shown in Figure C.5.
According to the principle of superposition, in a linear circuit containing several independent
sources, the current or voltage of a circuit element equals the algebraic sum of the component
voltages or currents produced by each individual source acting alone [24]. Using this theorem,
we can conclude that the resultant voltages or currents are the sum of the currents/voltages
before the fault occurred and the currents/voltages due to the fault itself. This is represented

149

Figure C.4: Open switch replaced by voltage source.

Figure C.5: Closed Switch Replaced by Two Sources.

graphically as the addition of Figure C.6 and Figure C.7.


To begin let the pre-fault bus voltages and pre-fault bus current be denoted as V 0 and
I 0 respectively. Also let the faulted bus voltages and fault current be denoted as Vf and
If respectively. Using Figure C.8, we can establish a relationship between the pre-fault bus
voltage, faulted bus voltage and the fault current. This relationship can be seen by reducing
power network in Figure C.8 to a thevanin equivalent. We begin reducing the power network
in Figure C.8 by open circuiting all current sources and short circuiting all voltage sources. In
this model, we assume that the substation can be modelled as a thevanin equivalent source.
Also, we assume the load model can be modelled as a constant current source. In Figure C.8
we are considering only the effect of the fault current itself, so we short circuit the substation
source and open circuit the load current sources.
Performing a simple KVL loop yields the following equation[22]:
Vf = V0 Zth If

150

(C.1)

Figure C.6: Circuit during pre-fault conditions.

Figure C.7: Calculation of currents due to the fault itself.

This equation represents the relationship between the system pre-fault voltages and faulted bus
voltages at any bus k. This fundamental relationship is used later in this section to derive the
phase thevanin matrix.

C.1.2

Pre-Fault and Faulted Systems in DEW

To better understand how DEW calculates fault currents under loaded conditions, we start by
investigating the pre-fault and faulted networks. In Section C.1 we said that the superposition
theorem states that the resultant currents are the sum of the currents before the fault occurred
and the currents due to the fault itself. This is respesented by Equation C.2 below.
If loaded = I 0 + If

(C.2)

We also showed in Section C.1 that Figure C.6 and Figure C.7 represented the currents
flowing before the fault and the currents due to the fault itself respectively. In this paper,

151

Figure C.8: Converting the network to a thevanin equivalent.

we will refer to these circuits as the pre-fault and faulted networks. In order begin our fault
analysis we need to obtain the pre-fault voltage at the faulted bus N . In order to do this, we
call a 3-phase power flow algorithm to solve for the system voltages and currents(MATLAB and
DEW contain a powerflow tool). The pre-fault system to be solved by the power flow algorithm
is in Figure C.9, consisting of a substation(converted to a thevanin equivalent), several load
buses, and bus N .

Figure C.9: Pre-Fault system showing several loads and bus N(future faulted bus).

Let us consider bus N to be the future faulted node. The pre-fault system voltages returned

152

by the power flow algorithm are of the form:

i
Van

i
Vbn
i
Vcn
The pre-fault load currents are expressed as:

i
ILoadA

ILoadB
i
ILoadC
The stored load matrix holds all load flow of every bus up to the nth bus:

1
SLoadA

SLoadB

1
SLoadC

..

n
SLoadA

n
SLoadB

n
SLoadC

We can represent the faulted system, by short circuiting voltage sources, open circuiting
current sources and adding a voltage source of the same magnitude but 180 degrees out of phase
with the pre-fault voltage at bus N . This is shown in Figure C.10.

Figure C.10: Faulted System with Fault applied at Node N.

153

Before we can begin to calculate fault current in Figure C.10, we must convert the system
into a thevanin equivalent impedance looking into bus N. This requires a 3x3 matrix consisting of
self and mutual equivalent impedances for all 3-phases. This is called the phase thevanin matrix.
The matrix terms Zaa , Zbb , Zcc represent the self impedances of phase a, b and c respectively.
The remaining terms(off diagonal) represent mutual impedances between phases, for example
Zab represents the mutual impedance between phases a and b.

Zaa Zab Zac

Zth = Zba

Zbb

Zca

Zcb

Zbc
Zcc

(C.3)

Many times, manufactures of cabling for distribution systems will only supply positive, negative and zero sequence impedances for a line segment(s). By applying a reverse transformation,
we can go from the symmetrical component frame of reference to the phase frame.

Z0

Z012 = 0

Z+

0
Z

(C.4)

We can define the relation between the sequence component frame and phase frame as:

T = 1 2
1

(C.5)

Where 2 = 16 120 . Transformation of the impedance is not a straightforward procedure,


so we begin here:

Vabc = Zth Iabc


T V012 = Zth T I012
V012 = T1 Zth T I012
V012 = Z012 I012
Therefore:
Z012 = T1 Zth T
Zth = T Z012 T1
In the DEW software package, this conversion can be done for you. When clicking on a

154

distribution cable, we can define the positive, negative, and zero sequence impedances or you
can specify Zabc .

C.1.3

Forming the Phase Thevanin Matrix

Forming the phase thevanin matrix(PTM) begins by analysing the system network. Using the
system network model, we follow the following steps to obtain the PTM:
Load: The constant current load is scaled such that the load current flow is 0. This is
done via a scaling factor i .

1
ILoadA

ILoadB

1
ILoadC

..
0
.

n
ILoadA

n
ILoadB

n
ILoadC

Fault: The fault is removed from the faulted bus, and a test load is inserted.

Figure C.11: Pre-Fault system with scaled loads and test load inserted at the faulted bus.

When forming the phase thevanin matrix, a test load is inserted sequentially at each phase
of the faulted node and power flow is performed. The result of the power flow on the test load is
the test load voltage and current. The phase thevanin matrix represents the relationship between
the voltages changes and the current changes at the faulted node before and after the test load
is inserted. By making the applied system pre-fault load very small, this greatly reduces the

155

time needed to run power flow on the test load. Before the test load is inserted, a power flow
is performed and the pre-test load voltages are determined.
Building the phase thevanin matrix can be broken into the following steps:
1. Remove the fault from the faulted bus.
2. Scale the loads using the scaling factor i .
3. Run a power flow algorithm on the system to obtain the pre-test voltages Vkn where
k = a, b, c

Van

Vbn
Vcn
4. Insert a test load at the faulted bus on Phase A, then run power flow to determine test
(1)

the test load current Ia

(1)

and voltage Vkn where k = a, b, c. This is show in Figure C.12.

(1)

Van

(1)
Vbn
(1)
Vcn

Figure C.12: Test load inserted on Phase A of the faulted node.

156

5. Insert a test load at the faulted bus on Phase B, then run power flow to determine test
(2)

the test load current Ib

(2)

and voltage Vkn where k = a, b, c. This is show in Figure C.13.

(2)

Van

(2)
Vbn
(2)
Vcn

Figure C.13: Test load inserted on Phase B of the faulted node.

6. Insert a test load at the faulted bus on Phase C, then run power flow to determine test
(3)

the test load current Ic

(3)

and voltage Vkn where k = a, b, c. This is show in Figure C.14.

(3)

Van

(3)
Vbn
(3)
Vcn
7. Form the phase thevanin martix using equation C.6.

Zth

(1)

Van Van
(1)
Ia
(1)
Vbn Vbn
(1)

Ia
(1)
Vcn Vcn
(1)
Ia

(2)

Van Van
(2)
Ib
(2)
Vbn Vbn
(2)

Ib
(2)
Vcn Vcn
(2)
Ib

157

(3)

Van Van
(3)
Ic
(3)
Vbn Vbn
(3)

Ic
(3)
Vcn Vcn
(3)
Ic

(C.6)

Figure C.14: Test load inserted on Phase C of the faulted node.

C.1.4

System Fault Characteristics

Before performing analysis of the system under faulted conditions, we must study the fault
itself. Distribution systems are often constructed using overhead cabling, exposing it to a
wide variety of human and environmental interference. Utilities often record the root cause
of faults for future study: Lightning, Tree Contact, Ice/Snow, Animals, Vandalism, Construction, Vehicle Accidents, Wind, Equipment Failure, and Conductor Dig-Ins are most common.

Figure C.15: Fault Root Causes and Percent Occurrence[7].

158

Table C.1: Fault types and their frequency of occurrence[23].


Fault Type

Occurrence Percentage

Three Phase Fault


Phase-to-Phase
Phase-to-Phase-Ground
Single Line to Ground

5%
15%
10%
70%

Distributions faults are usually classified by the number of conductors that are involved.
For example, a single conductor shorted to ground is called a single line to ground fault(SLG).
When multi-phase faults occur we specify these as phase-to-phase(LL), phase-to-phase-toground(LLG) or 3-phase faults(3P).
3-Phase Fault: A general three phase fault involves short circuiting all three phases to one
common point, hence the name: three phase fault. Three phase faults can normally be
regard as the rarest but most severe type of fault. Usually 3-phase faults begin as phaseto-phase or phase-to-ground faults. This characteristic is important to note because most
fault location algorithms assume the fault type is non-evolving1 .
Phase-to-Phase and Phase-to-Phase-Ground Faults: A general phase-to-phase fault involves short circuiting two phases to one common point. If the fault is phase-to-phaseground, the common point is short circuited to ground. Phase-to-phase faults and Phaseto-Phase Ground faults are considered less severe than a 3-phase fault but more severe
than single phase fault. Phase to phase faults rarely occur on underground line segments
and mostly occur on overhead lines.
Single Line to Ground Fault: A general single line to ground fault involves short circuiting
a single line directly to ground. Single line to ground faults are the most common fault
occurring on power systems. For fault locating purposes, most analysis is centered around
this fault type.
Many faults occur by contacting trees, asphalt, gravel and dirt with quite different and
unpredictable fault impedances. Many tests have shown that conductors contacting the wet
soil, dry soil, concrete and rocks result in highly variable fault impedances. In general most
utitlites assume the fault impedance to be negligible or approximately zero. According to
1

This can be extremely problematic for most fault locating algorithms. The reasons for this will be cover in
later chapters.

159

studies done by the EPRI and utilties in [7], a fault impedance of 0 ohms proved to be a valid
assumption. EPRI stated the following: Good fault location can be done by assuming a bolted
faultno fault arcing resistance. Results from the EPRI fault study published in the early 1980s
showed that actual fault currents were close to the calculated value. The EPRI study found
that calculated fault currents were approximately 2 percent higher than the measured value.
Therefore, we assume that fault resistance cannot play a drastic role, but some faults may have
enough arc length to make the bolted-fault assumption less accurate than desired.

C.1.5

3-Phase Faults

We first consider the general case of a 3-phase fault shown in Figure C.16. Each phase consists
of its own fault impedance Za , Zb , Zc and a common point impedance Zf . We begin our
analysis by re-arranging Equation C.1:
V0 Vf = Zth If

Figure C.16: Model of a 3 Phase Fault.

Where, our pre-fault voltages are:

i
Van

i
V0 = Vbn

i
Vcn
We also define are fault voltages as:

f
Van

f
Vf = Vbn

f
Vcn

160

(C.7)

And our resultant fault currents as:

Ia

If = Ib
Ic
If we evaluate Equation C.1.8 we get the following equation:

i
Van

f
Van

Ia

i f

Vbn Vbn = Zth Ib


f
i
Vcn
Vcn
Ic
In order to solve for the If vector, we must solve our faulted bus voltage vector Vf in terms
of fault currents and fault impedances. This can be done by applying a KVL to Figure C.16.

V f = Ia Za + Zf If

an
f
KV L Vbn
= Ib Zb + Zf If

f
Vcn
=I Z +Z I
c

We can equate If in terms of Ia , Ib and Ic : If = Ia + Ib + Ic

V f = Ia Za + Zf (Ia + Ib + Ic )

an
f
KV L Vbn
= Ib Zb + Zf (Ia + Ib + Ic )

f
Vcn
= I Z + Z (I + I + I )
c

If we substitute we get:

i
Van

Ia Za + Zf (Ia + Ib + Ic )

Ia

i
Vbn Ib Zb + Zf (Ia + Ib + Ic ) = Zth Ib
i
Vcn
Ic Zc + Zf (Ia + Ib + Ic )
Ic
To simplify notation, we will define Zth as:

a11 a12 a13

Zth = a21 a22 a23


a31 a32 a33
Substitution yields:

161

i
Van

Ia Za + Zf (Ia + Ib + Ic )

a11 a12 a13

Ia


i
Vbn Ib Zb + Zf (Ia + Ib + Ic ) = a21 a22 a23 Ib
i
a31 a32 a33
Ic
Ic Zc + Zf (Ia + Ib + Ic )
Vcn
Solving for the pre-fault voltage vector gives the following equation:

i
Van

i
Vbn =
i
Vcn
|

Za + Zf g + a11

Zf g + a12

Zf g + a13

a21 + Zf g

a22 + Zb + Zf g

a23 + Zf g

a31 + Zf g

a32 + Zf g
{z

a33 + Zcc + Zf g

Ib
Ic
}

Ia

Solving for the fault currents gives:

Ia


Ib =
Ic
|

Za + Zf g + a11

Zf g + a12

Zf g + a13

a21 + Zf g

a22 + Zb + Zf g

a23 + Zf g

a31 + Zf g

a32 + Zf g
{z

a33 + Zcc + Zf g

i
Van

i
Vbn
i
Vcn

(C.8)

i1

For a bolted fault with no arcing resistance:

Ia

a11 a12 a13

i
Van


i
Ib = a21 a22 a23 Vbn

i
Ic
a31 a32 a33
Vcn
|
{z
}
1
Zth

C.1.6

Phase-to-Phase-Ground Faults

Figure C.17: Model of a Phase-Phase-Ground Fault.

162

(C.9)

We first consider the general case of a phase-to-phase-ground fault shown in Figure C.17.
Each phase consists of its own fault impedance Zb , Zc and a common point impedance Zf .
In this case phase A is the only un-faulted phase, with Ia = 0. We begin our analysis by
re-arranging Equation C.1:
V0 Vf = Zth If
Where, our pre-fault voltages are:
!

i
Vbn

V0 =

i
Vcn

We also define are fault voltages as:


f
Vbn

Vf =

f
Vcn

And our resultant fault currents as:


Ib

If =

Ic

If we evaluate Equation C.1.8 we get the following equation:


i
Vbn
i
Vcn

f
Vbn

f
Vcn

Ib

= Zth

Ic

In order to solve for the If vector, we must solve our faulted bus voltage vector Vf in terms
of fault currents and fault impedances. This can be done by applying a KVL to Figure C.17.

V f = I Z + Z I
b
b
f
f
bn
KV L
f
Vcn
= Ic Zc + Zf If
We can equate If in terms of Ib and Ic : If = Ib + Ic

V f = I Z + Z (I + I )
c
b
b
f
b
bn
KV L
f
Vcn
= Ic Zc + Zf (Ib + Ic )
If we substitute we get:
i
Vbn
i
Vcn

Ib Zb + Zf (Ib + Ic )
Ic Zc + Zf (Ib + Ic )

163

!
= Zth

Ib
Ic

To simplify notation, we will define Zth as:


Zth =

a22 a23

a32 a33

Substitution yields:
i
Vbn

Ib Zb + Zf (Ib + Ic )

i
Vcn

Ic Zc + Zf (Ib + Ic )

!
=

a22 a23

Ib

a32 a33

Ic

Solving for the pre-fault voltage vector gives the following equation:
i
Vbn

a22 + Zb + Zf

i
Vcn

a23 + Zf

a32 + Zf
{z

Ic

a33 + Zc + Zf

Ib

Solving for the fault currents gives:


Ib
Ic

!
=

!1

a22 + Zb + Zf

a23 + Zf

a32 + Zf

a33 + Zc + Zf
{z

i
Vbn
i
Vcn

!
(C.10)

i1

For a bolted fault with no arcing resistance:


Ib

Ic

a22 a23
a32 a33
{z
|
1
Zth

C.1.7

!1

i
Vbn

i
Vcn

Phase-to-Phase Faults

Figure C.18: Model of a Phase-Phase Fault.

164

(C.11)

We first consider the general case of a phase-to-phase fault shown in Figure C.18. In this
case the fault impedance between Phase A and Phase B is Zf . In this case phase A is the only
un-faulted phase, with Ia = 0. We begin our analysis by re-arranging Equation C.1:
V0 Vf = Zth If
Where, our pre-fault voltages are:
!

i
Van

V =

i
Vbn

We also define are fault voltages as:


f
Van

Vf =

f
Vbn

And our resultant fault currents as:


Ia

If =

Ib

If we evaluate Equation C.1.8 we get the following equation:


i
Van
i
Vbn

f
Van

f
Van

= Zth

Ia
Ib

!
(C.12)

To simplify notation, we will define Zth as:

Zth =

a11 a12

a21 a22

Substituting for Zth gives the following equations:


i
f
Van
Van
= a11 Ia + a12 Ib

(C.13)

f
i
Vbn
Vbn
= a21 Ia + a22 Ib

(C.14)

We begin by identifying that during a phase-to-phase fault, Ia = Ib = If . Subtracting


C.13 from C.14 and substituting Ia = Ib :

165

f
i
i
f
Van
Vbn
(Van
Vbn
) = [(a11 + a12 ) (a12 + a21 )] Ia

(C.15)

f
f
If we measure Van
and Vbn
,then subtracting gives:

V f Vbn
= If Zf

an

(C.16)

f
f
Van
Vbn
= Ia Zf
We can see that Equation C.16 give us the voltage across the fault impedance. We can
easily substitute If = Ia . We can use this relationship by substituting the second equation in
C.16 into Equation C.15. When we solve for the fault current we get:
Ia = Ib = If =

i Vi
Van
bn
(a11 + a22 ) (a12 + a22 ) Zf

(C.17)

For a bolted fault with no arcing resistance:


Ia = Ib = If =

C.1.8

i Vi
Van
bn
(a11 + a22 ) (a12 + a22 )

(C.18)

Single Line-to-Ground Faults

Line to ground faults are the easiest of fault types to calculate. We first consider the general
case of a phase-to-ground fault shown in Figure C.18. In this case the fault impedance between
Phase A and Ground is Zf . In this case phase A is the only faulted phase, with Ib = Ic = 0.
We begin our analysis by re-arranging Equation C.1:
V0 Vf = Zth If
i
Where, our pre-fault voltages are: V 0 = Van
f
We also define are fault voltages as: Vf = Van

And our resultant fault currents as: If = Ia


We can also reduce Zth to: Zth = a11
i Vf =a I
If we evaluate Equation C.1.8 we get the following equation: Van
an
11 f

166

Figure C.19: Model of a Phase-to-Ground Fault.

f
f
Substituting If = Ia and Van
= Za If gives: Van
Za If = a11 Ia

We can solve for the fault current If = Ia :


Ia = If =

i
Van
a11 + Zf

(C.19)

i
Van
a11

(C.20)

For a bolted fault with no arcing resistance:


Ia = If =

C.2

Verification of DEW Fault Current Calculation: Example


Feeder

When performing short circuit studies in DEW, the phase-thevanin matrix, load-flow analysis
and calculation of the fault current magnitude are all done without presentation to the user. In
this section, we show the step-by-step execution of DEWs fault analysis under loaded conditions
while also illustrating that it is sufficient for fault locating applications.
We begin our analysis by considering a simple radial feeder shown in Figure C.20. To
maintain the focus on the algorithm itself, simple values were chosen for line impedance values
and load conditions. As shown in the figure, this feeder consists of three load buses connected
to a 13.8kV source via three line segments.
As described in previous sections, we begin our fault analysis by considering the pre-fault
conditions of the circuit. To obtain pre-fault quantities, the DEW 3-Phase Power Flow tool is
utilized. This tool performs power flow on the network model and reports voltages, currents

167

Figure C.20: Simple Radial Feeder with three load buses.

and their corresponding angles. As shown in Figure C.20, three-phase loads are connected to
buses 1,2 and 3 rated at 90kW/43.6kVAR. Power flow was performed using a constant current
load model. To verify that the DEW power flow solution is accurate, the solution is verified
with MATLAB Simulink Power Flow Tool. A MATLAB script was written to calculate 3-phase
pre-fault voltages, currents, and their corresponding angles. The power flow results from DEW
and MATLAB are shown in Tables C.2, C.3.

Table C.2: L-G Voltage Results of DEW vs. MATLAB Powerflow Script
Quantity
Bus
Bus
Bus
Bus
Bus
Bus
Bus
Bus
Bus

1
1
1
2
2
2
3
3
3

L-G
L-G
L-G
L-G
L-G
L-G
L-G
L-G
L-G

DEW Power Flow Tool


Phase
Phase
Phase
Phase
Phase
Phase
Phase
Phase
Phase

A Voltage
B Voltage
C Voltage
A Voltage
B Voltage
C Voltage
A Voltage
B Voltage
C Voltage

7.962
7.962 6
7.962 6
7.958
7.958 6
7.958 6
7.956
7.956 6
7.956 6

-.0812 kV
-120.0812 kV
119.9188 kV
6 -.1354 kV
-120.1354 kV
119.8646 kV
6 -.1625 kV
-120.1625 kV
119.8375 kV

MATLAB Power Flow Script


7.9619
7.9619 6
7.9619 6
7.9583
7.9583 6
7.9583 6
7.9565
7.9565 6
7.9565 6

-0.0814 kV
-120.0814 kV
119.9186 kV
6 -0.1355 kV
-120.1355 kV
119.8645 kV
6 -0.1626 kV
-120.1626 kV
119.8374 kV

We begin our fault analysis by choosing a fault location and a fault type. In this example we
will be using a 3-Phase bolted fault(Rf = 0) at Bus 1. Referring back to Equation C.9 it can
be seen that the Phase Thevenin matrix and Pre-Fault Bus Voltages are needed to calculate
the resultant fault current. Since we have obtained the pre-fault voltages via power-flow(Tables
C.2, C.3), our next step will require the formation of the phase thevenin matrix.

168

Table C.3: Current Results of DEW vs. MATLAB Script


Quantity
Source Phase A Current
Source Phase B Current
Source Phase C Current
Line 2 Phase A Current
Line 2 Phase B Current
Line 2 Phase C Current
Line 3 Phase A Current
Line 3 Phase B Current
Line 3 Phase C Current
Load 1 Phase A Current
Load 1 Phase B Current
Load 1 Phase C Current
Load 2 Phase A Current
Load 2 Phase B Current
Load 2 Phase C Current
Load 3 Phase A Current
Load 3 Phase B Current
Load 3 Phase C Current

DEW Power Flow Tool

MATLAB Power Flow Script

12.55 6 25.9224 A
12.55 6 -145.9224 A
12.55 6 94.0776 A
8.36 6 -25.9449 A
8.36 6 -145.9449 A
8.36 6 94.0551 A
4.18 6 -25.9585 A
4.18 6 -145.9585 A
4.18 6 94.0415 A
4.18 6 -25.8772 A
4.18 6 -145.9300 A
4.18 6 94.0686 A
4.18 6 -25.9314 A
4.18 6 -145.9300 A
4.18 6 94.0686 A
4.18 6 -25.9585 A
4.18 6 -145.9585 A
4.18 6 94.0551 A

12.5328 6 -25.9222 A
12.5328 6 -145.9222 A
12.5328 6 94.0778 A
8.3536 6 -25.9446 A
8.3536 6 -145.9446 A
8.3536 6 94.0554 A
4.1763 6 -25.9582 A
4.1763 6 -145.9582 A
4.1763 6 94.0418 A
4.1792 6 -25.8774 A
4.1792 6 -145.8774 A
4.1792 6 94.1226 A
4.1773 6 -25.9316 A
4.1773 6 -145.9316 A
4.1773 6 94.0684 A
4.1763 6 -25.9582 A
4.1763 6 -145.9582 A
4.1763 6 94.0418 A

169

To form the Phase Thevanin Matrix we refer back to the procedure outlined in Section
C.1.3. Building the phase thevanin matrix can be broken into the following steps:
1. Scale the loads using the scaling factor i . A scaling factor of i = 103 was used [22].
2. Run a power flow algorithm on the system to obtain the pre-test voltages at the faulted
node(Bus 1) Vkn where k = a, b, c.

Van

7.962kV 6 .0182

Vbn = 7.962kV 6 120.0182


Vcn
7.962kV 6 119.9188
3. Insert a test load at the faulted bus(Bus 1) on Phase A, then run power flow to determine
(1)

test the test load current Ia

(1)

and voltage Vkn where k = a, b, c. This is show in Figure

C.12. A test load value of Rtest = 100 was used. The test load current was found to
(1)

be: Ia = 792.7523 A6 5.7327 .

(1)

Van

7.9275 kV 6 5.7327

(1)

Vbn = 7.9674 kV 6 120.0008


(1)
7.9674 kV 6 119.9992
Vcn
4. Insert a test load at the faulted bus(Bus 1) on Phase B, then run power flow to determine
(2)

test the test load current Ib

(2)

and voltage Vkn where k = a, b, c. This is show in Figure

C.13. A test load value of Rtest = 100 was used. The test load current was found to
(2)

be: Ib

= 792.7523 A6 125.7327 .

(2)

Van

7.9674 kV 6 0.0008

(2)

Vbn = 7.9275 kV 6 125.7327


(2)
7.9674 kV 6 119.9992
Vcn
5. Insert a test load at the faulted bus(Bus 1) on Phase C, then run power flow to determine
(3)

test the test load current Ic

(3)

and voltage Vkn where k = a, b, c. This is show in Figure

C.14. A test load value of Rtest = 100 was used. The test load current was found to
(3)

be: Ic

= 792.7523 A6 114.2673 .

(3)

Van

7.9674 kV 6 0.0008

(3)

Vbn = 7.9674 kV 6 120.0008


(3)
7.9275 kV 6 114.2673
Vcn

170

6. Form the phase thevanin matrix using equation C.6. The result is as follows:

(7.962 kV 6 .0182 )(7.9275 kV 6 5.7327 )


7.9286 5.7327 kA

Zth =

(7.962 kV 6 120.0182 )(7.9275 kV 6 125.7327 )


7.92756 125.7327 kA

Zth = 0

1j

0
1j

1j

1j

1j

Z1
th =

(C.21)

1j

(C.22)

(C.23)

Using Equation Equation C.9, we get the result:

Iaf

1j

1j

0
{z


Ibf =
Icf
|

1
Zth

7.962 kV 6 .0182

7.962kV 6 120.0182
7.962 kV 6 119.9188
1j
}

(C.24)

Which yields the following:

Iaf

7962 A6 90.0182

Ibf = 7962 A6 149.9818


7962 A6 29.9188
Icf
We can calculated the fault current seen by the substation under loaded conditions by using
the principle of superposition as described in previous sections: If loaded = I 0 + If

Iaf loaded

7962 A6 90.0182

12.5328 A6 25.9222


Ibf loaded = 7962 A6 149.9818 + 12.5328 A6 145.9222
Icf loaded
7962 A6 29.9188
12.5328 A6 94.0778
Therefore the current seen at the substation during loaded fault conditions are:

171

Iaf loaded

7967.32 A6 90.01


Ibf loaded = 7967.32 A6 149.98
7967.31 A6 29.98
Icf loaded
To validate this calculation, a 3-Phase bolted fault was place at Bus 1. DEW gives the
result shown in Figure C.21 which is If loaded = 7967 A.

Figure C.21: DEW Event Report showing the current seen at the substation for a bolted
3-Phase fault at Bus 1.

172

Table C.4: MATLAB Simulink Fault Current Results


Quantity

MATLAB Simulink Fault Current Results


7967.1 A 6 89.9
7967.1 A 6 150.1
7967.1 A 6 30.06

Phase A Fault Current


Phase B Fault Current
Phase C Fault Current

C.2.1

Validation with MATLAB Simulink Model

To verify and validate DEW fault analysis, a model of Figure C.20 was built in MATLAB
Simulink. The model consists of a 13.8kV L-L source, three lines of impedance Zline = 1j and
three load buses with constant current loads. A bolted 3-Phase fault (Rf = 0)was applied and
yielded the results in Table C.4.

173

Appendix D

Supplemental FLA Simulation


Results for Stewart Street 12.47kV
Feeder
D.1

Pole P4622 Supplemental Recorded Results

174

Figure D.1: Worst Matched Locations for a fault at P4622 as calculated by the FLA.

175

Figure D.2: Other Likely Fault Locations for a fault at P4622 as calculated by the FLA.

176

Table D.1: Other Likely Fault Possibilities for P4622


Node
P4509
P4513
P4515
P4520
P4522
N7974
1L5936
N7976
L1254807
L1254808
L1254809
L1279017
P4498
N7977
P4241
L1254804
P4268

Rank

Upstream Device

3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19

FUSE654427653(65T)
FUSE654427653(65T)
FUSE654427653(65T)
FUSE654427653(65T)
FUSE654427653(65T)
FUSE654427566(25T)
FUSE654427566(25T)
FUSE654427566(25T)
FUSE1052482716(40T)
FUSE1052482716(40T)
FUSE1052482716(40T)
FUSE1146924279(25T)
FUSE654421586(25T)
FUSE1071439054(15T)
FUSE654427484(25T)
FUSE1052482746(200T)
FUSE1052482746(200T)

177

D.2

Pole P4266 Supplemental Recorded Results

Figure D.3: Worst Matched Locations for a fault at P4266 as calculated by the FLA.

178

Figure D.4: Other Likely Fault Locations for a fault at P4266 as calculated by the FLA.

179

Table D.2: Other Likely Fault Possibilities for P4266


Node
L1254806
L1254807
L1254808
L1254809
P4619
P4622
L1279017
N7974
1L5936
N7976
N7977
P4241
P4508
P4509
P4513
P4515
P4520
P4496
P4498

Rank

Upstream Device

4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22

FUSE1052482716(40T)
FUSE1052482716(40T)
FUSE1052482716(40T)
FUSE1052482716(40T)
FUSE654421506(65T)
FUSE654421506(65T)
FUSE1146924279(25T)
FUSE654427566(25T)
FUSE654427566(25T)
FUSE654427566(25T)
FUSE1071439054(15T
FUSE654427484(15T)
FUSE654427653(65T)
FUSE654427653(65T)
FUSE654427653(65T)
FUSE654427653(65T)
FUSE654427653(65T)
FUSE654421586(25T)
FUSE654421586(25T)

180

Vous aimerez peut-être aussi