12.01.2015 Views

Journal of Reliable Power - SEL

Journal of Reliable Power - SEL

Journal of Reliable Power - SEL

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

$12.00<br />

<strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong><br />

Volume 1 Number 1 July 2010


Schweitzer Engineering Laboratories, Inc.<br />

2350 NE Hopkins Court<br />

Pullman, WA 99163<br />

Copyright ©2010 Schweitzer Engineering Laboratories, Inc. All rights reserved.


<strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong> Volume 1 Number 1 July 2010<br />

2<br />

4<br />

15<br />

29<br />

41<br />

52<br />

65<br />

83<br />

86<br />

87<br />

Introduction<br />

Distance Relay Element Design<br />

E. O. Schweitzer, III and J. B. Roberts (1993)<br />

Impedance-Based Fault Location Experience<br />

K. Zimmerman and D. Costello (2004)<br />

Digital Communications for <strong>Power</strong> System Protection: Security, Availability, and Speed<br />

E. O. Schweitzer, III, K. Behrendt, and T. Lee (1998)<br />

Transmission Line Protection System for Increasing <strong>Power</strong> System Requirements<br />

A. Guzmán, J. Mooney, G. Benmouyal, and N. Fischer (2001)<br />

Lessons Learned Analyzing Transmission Faults<br />

D. Costello (2007)<br />

Adaptive Phase and Ground Quadrilateral Distance Elements<br />

F. Calero, A. Guzmán, and G. Benmouyal (2009)<br />

Line Protection Bibliography<br />

<strong>SEL</strong> University 2010 Course Schedule<br />

July through December<br />

Modern Solutions for Protection, Control, and Monitoring <strong>of</strong> Electric <strong>Power</strong> Systems<br />

Ordering Information<br />

Issue Editors<br />

Bogdan Kasztenny is a principal systems engineer in the Research and Development Division <strong>of</strong><br />

Schweitzer Engineering Laboratories, Inc. He has 20 years <strong>of</strong> experience in protection and control,<br />

including his ten-year academic career at Poland’s Wrocław University <strong>of</strong> Technology, Southern<br />

Illinois University, and Texas A&M University. He also has ten years <strong>of</strong> industrial experience with<br />

General Electric, where he developed, promoted, and supported many protection and control products.<br />

Bogdan is an IEEE Fellow, Senior Fulbright Fellow, Canadian member <strong>of</strong> CIGRE Study Committee<br />

B5, and an adjunct pr<strong>of</strong>essor at the University <strong>of</strong> Western Ontario. He has authored about 200 technical<br />

papers and holds 16 patents. He is active in the IEEE <strong>Power</strong> System Relaying Committee and is a<br />

registered pr<strong>of</strong>essional engineer in the province <strong>of</strong> Ontario.<br />

Karl Zimmerman is a senior power engineer with Schweitzer Engineering Laboratories, Inc.<br />

in Fairview Heights, Illinois. His work includes providing application and product support and<br />

technical training for protective relay users. He is an active member <strong>of</strong> the IEEE <strong>Power</strong> System<br />

Relaying Committee and chairman <strong>of</strong> the Working Group on Distance Element Response to<br />

Distorted Waveforms. Karl received his BSEE degree at the University <strong>of</strong> Illinois at Urbana-<br />

Champaign and has over 20 years <strong>of</strong> experience in system protection. He is a past speaker at many<br />

technical conferences and has authored over 20 papers and application guides on protective relaying.


Introduction<br />

With this inaugural issue, <strong>SEL</strong> introduces the<br />

<strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>. This regularly published<br />

journal will present a series <strong>of</strong> papers with lasting<br />

value that focuses on the fundamentals <strong>of</strong> power reliability.<br />

We believe that the concepts, case studies, technologies,<br />

and issues that the <strong>Journal</strong> covers will be applicable<br />

to power engineers working for utilities, industrial and<br />

commercial facilities, government, and academia.<br />

Each issue <strong>of</strong> the <strong>Journal</strong> will focus on a broad topic<br />

area or theme. The theme <strong>of</strong> this first issue, guest-edited<br />

by <strong>SEL</strong> Senior <strong>Power</strong> Engineer Karl Zimmerman and<br />

Principal Systems Engineer Bogdan Kasztenny, with the<br />

combined technical expertise <strong>of</strong> many <strong>SEL</strong> authors, is<br />

line distance protection, where <strong>SEL</strong> got its start. Future<br />

themes will include the application <strong>of</strong> digital communications<br />

in protection and control, distribution automation<br />

and protection (“smart grid”), communications for critical<br />

infrastructure, synchronous measurement and control,<br />

and advances in protection methods. The second issue,<br />

edited by <strong>SEL</strong> Senior Automation System Engineer Tim<br />

Tibbals, will focus on communications and protocols. We<br />

welcome your suggestions for topics as well.<br />

Several outstanding publications in the electric power<br />

industry today communicate important ideas, applications,<br />

successes, and trends through news articles, editorials,<br />

and success stories. Our goal is to complement<br />

these publications by sharing in-depth, conference-length<br />

papers in a journal format.<br />

The <strong>Journal</strong> is forward-looking. When we include an<br />

older paper, it will be a foundational classic, with concepts<br />

that are still applicable today. In many cases, these<br />

older papers will teach us new things as we read and apply<br />

them; the power system and the work we do have changed<br />

in ways that none <strong>of</strong> us could have anticipated.<br />

We hope that you find this first collection <strong>of</strong> papers<br />

centered around line protection relevant to you and the<br />

work you do to make electric power safer, more reliable,<br />

and more economical.<br />

***<br />

Microprocessor-based relay technology became possible<br />

more than 25 years ago, owing to a perfect<br />

combination <strong>of</strong> available, suitable microprocessors and<br />

advancements in industrial electronics more generally;<br />

the opportunity to provide new functions far beyond what<br />

was possible within the electromechanical and static technologies;<br />

remarkable size and cost reductions; and the<br />

imagination, skills, and perseverance <strong>of</strong> early inventors in<br />

this new field.<br />

Early designs, although limited by available processing<br />

power, integrated a wealth <strong>of</strong> new functions, ranging<br />

from sequential events recording and fault location to<br />

metering, multiple settings groups, and communications.<br />

Once digital relays proved reliable in the eyes <strong>of</strong> the<br />

power system protection community, and as the dramatic<br />

cost reduction compared to previous technologies became<br />

evident, the digital relay revolution only accelerated. It<br />

became increasingly obvious that analog technology, in<br />

addition to being more expensive, could not provide new<br />

functions, self-monitoring, reliability, innovative operating<br />

principles, and other facets <strong>of</strong> protection performance<br />

made possible through microprocessor-based technology.<br />

Early microprocessor-based protective relays brought<br />

innovation in line protection principles as well as digital<br />

implementation <strong>of</strong> the existing art <strong>of</strong> distance protection.<br />

Constrained by the available processing power and<br />

facing the disadvantage <strong>of</strong> a nonzero processing latency,<br />

designers strived for increased response time <strong>of</strong> their<br />

algorithms while optimizing the computational power<br />

required. Applying the cosine filter and shaping distance<br />

characteristics via “m calculations” are good examples<br />

<strong>of</strong> smart design that delivered good performance under<br />

the limitations <strong>of</strong> early microprocessor-based relay platforms.<br />

“Distance Relay Element Design,” written in 1993<br />

and included in this issue <strong>of</strong> the <strong>Journal</strong>, is a foundational<br />

paper for <strong>SEL</strong> microprocessor-based distance protection.<br />

One cannot overlook the role that fault location played<br />

in the industry’s adoption <strong>of</strong> the new technology. This<br />

useful and “safer”—compared with the full protection<br />

package—function opened the door for new technology<br />

(<strong>SEL</strong>-21 and <strong>SEL</strong>-121 Distance Relay and Fault Locator<br />

products). In the 1980s, early applications <strong>of</strong> relays<br />

working as fault locators demonstrated hardware and s<strong>of</strong>tware<br />

reliability and delivered solid, useful, and verifiable<br />

information to early adopters. “Impedance-Based Fault<br />

Location Experience” summarizes the almost 25 years <strong>of</strong><br />

advancements in fault location at the time it was written<br />

in 2004.<br />

2 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


Capable <strong>of</strong> running protection calculations with latencies<br />

low enough for line protection, microprocessor-based<br />

relays allowed designers to take full advantage <strong>of</strong> the flexibility<br />

<strong>of</strong> digital implementations. Building protection elements<br />

by calculating signature signals and freely applying<br />

comparators, logic, and timers, designers were no longer<br />

constrained by analog means; they could truly innovate<br />

by going back to first principles and using these principles<br />

to devise new and powerful protection elements. Positivesequence<br />

voltage memory to maintain distance element<br />

security during close-in faults, advancements in digital filtering,<br />

a negative-sequence impedance ground directional<br />

element, load-encroachment logic, and fault identification<br />

selection logic are among the many innovations introduced<br />

in the early 1990s as part <strong>of</strong> the <strong>SEL</strong>-321 Relay.<br />

Ability to communicate was an inherent advantage<br />

<strong>of</strong> microprocessor-based relays from the very beginning.<br />

Communication started with users accessing and manipulating<br />

settings, records, and online measurements but<br />

quickly progressed into serving data to SCADA systems<br />

and peer-to-peer devices. Invention <strong>of</strong> Mirrored Bits ®<br />

communications—fast, reliable, and dependable protection-grade<br />

communications—opened a new chapter <strong>of</strong><br />

digital teleprotection signaling for line protection. (See<br />

“Digital Communications for <strong>Power</strong> System Protection:<br />

Security, Availability, and Speed” in this issue.)<br />

The last decade brought considerable advances in<br />

distance protection. Higher sampling rates and increased<br />

processing power opened the door to sophisticated protection<br />

principles and enhancements. Improvements in security<br />

and speed <strong>of</strong> distance protection have been achieved<br />

through a combination <strong>of</strong> CVT transient detection under<br />

high SIR conditions and usage <strong>of</strong> incremental quantities<br />

while applying optimized, faster, short-window filters.<br />

Fault type identification logic now performs better under<br />

weak infeed conditions. Adaptive polarizing algorithms<br />

maintain the best possible choice <strong>of</strong> polarization while<br />

releasing the user from making trade<strong>of</strong>fs and running<br />

detailed engineering studies to aid setting selection. Protection<br />

issues related to series compensation <strong>of</strong> transmission<br />

lines have been addressed as well. Frequency tracking<br />

<strong>of</strong> modern relays allows applications under stressed<br />

system conditions when frequency excursions and rate <strong>of</strong><br />

change <strong>of</strong> frequency jeopardize some traditional concepts,<br />

such as memory polarization. “Transmission Line Protection<br />

System for Increasing <strong>Power</strong> System Requirements”<br />

gives an excellent overview <strong>of</strong> these advancements.<br />

The <strong>SEL</strong>-421 Protection, Automation, and Control<br />

System, developed in 2001, incorporates all <strong>of</strong> these<br />

enhancements while including a truly impressive set <strong>of</strong><br />

functions that complement its core protection functionality.<br />

Examples are high-resolution digital fault recording,<br />

sequential events recording, and synchrophasor measurements,<br />

all with submicrosecond accuracy, a wealth <strong>of</strong><br />

both peer-to-peer and SCADA communications protocols,<br />

metering functions, advanced programmable protection<br />

and automation logic, two CT sets <strong>of</strong> inputs for dualbreaker<br />

line terminals with associated protection, and<br />

automation functions. These are functions that we all take<br />

for granted today.<br />

Looking back, the ability to provide data for postevent<br />

analysis turned out to be one <strong>of</strong> the major benefits<br />

<strong>of</strong> microprocessor-based relays. “Lessons Learned Analyzing<br />

Transmission Line Faults” provides an overview <strong>of</strong><br />

how data produced by modern relays aid troubleshooting<br />

and improve our understanding <strong>of</strong> the power system.<br />

Innovations in line protection are ongoing. Examples<br />

include improved power swing detection based on the rate<br />

<strong>of</strong> change <strong>of</strong> the swing center voltage and better resistive<br />

coverage <strong>of</strong> distance functions through an adaptive quadrilateral<br />

characteristic. (See “Adaptive Phase and Ground<br />

Quadrilateral Distance Elements” in this issue.)<br />

The journey that started a quarter <strong>of</strong> a century ago<br />

continues. Today, we use precise timing to align and timetag<br />

measurements. We use high-speed communications<br />

to share data in real time to improve protection and automation<br />

functions. And, we use more powerful processors<br />

to efficiently run increasingly more sophisticated algorithms.<br />

The opportunity that technology created 25 years<br />

ago is still here. It is only the laws <strong>of</strong> physics and our own<br />

imaginations that bound us today.<br />

If you have comments or suggestions, please e-mail<br />

journal@selinc.com or share your thoughts with our editors<br />

by phone at +1.509.332.1890.<br />

Introduction | 3


Distance Relay Element Design<br />

Edmund O. Schweitzer, III and Jeff B. Roberts, Schweitzer Engineering Laboratories, Inc.<br />

I. INTRODUCTION<br />

All distance relays compare voltages and currents to create<br />

impedance-plane and directional characteristics. Electromechanical<br />

relays do so by developing torques. Most staticanalog<br />

implementations use coincidence-timing techniques.<br />

Numerical techniques are the newest way to implement<br />

distance and directional relay elements. These relays use<br />

torque-like products and other methods to accomplish their<br />

operating characteristics. How do these new techniques relate<br />

to the classical electromechanical and static phase-angle<br />

comparators<br />

This paper presents basic distance and directional element<br />

design. A large emphasis is placed on relating the newer<br />

digital and numerical methods to the established electromechanical<br />

and static-analog methods <strong>of</strong> designing relay<br />

elements.<br />

In addition, we discuss:<br />

• A new method for characterizing distance elements;<br />

i.e., equations for mapping points on a relay<br />

characteristic onto a single point on a number line.<br />

• How multi-input comparators can be viewed as a<br />

family <strong>of</strong> two-input comparators.<br />

• Which characteristics result from various<br />

combinations <strong>of</strong> comparator inputs.<br />

• Classical element-security problems and remedies.<br />

• A new negative-sequence directional element.<br />

• A different approach to the load-encroachment<br />

problem.<br />

Finally, we point out a problem with fault-type selection<br />

logic which uses the angle between the negative- and zerosequence<br />

currents. It can select the wrong phase for certain<br />

resistive line-line-ground faults. We present a solution to this<br />

problem which compares ground and phase fault-resistance<br />

estimates.<br />

II. PHASE ANGLE COMPARATORS<br />

Phase angle comparators test the angle between various<br />

voltage and current combinations to produce directional,<br />

reactance, mho, and other characteristics.<br />

This section describes three technologies frequently used to<br />

compare phasors in relays: induction cylinders, coincidence<br />

timers, and digital multiplication. It also presents a new<br />

method: mapping <strong>of</strong> a characteristic (e.g., a mho circle) onto a<br />

point on a number line.<br />

A. Induction Cylinder Phase Comparator<br />

Figure 1 is a sketch <strong>of</strong> an induction cylinder comparator.<br />

Assume currents A and B flow in the windings as shown. The<br />

cup tends to rotate in the direction <strong>of</strong> the rotating flux<br />

established by the currents. For example, if B leads A, the cup<br />

rotates clockwise to close the contacts. If A and B are in<br />

phase, the net torque is zero and the cup does not move. This<br />

is the only external information available from the relay;<br />

either the contacts are open or closed.<br />

The equation for the cup torque, T, is:<br />

T = k •│A│•│B│• sin Θ,<br />

where Θ is the angle between A and B. External circuitry and<br />

the coils themselves can be used to modify the torque equation<br />

to test other phase relationships.<br />

Figure 1:<br />

Induction Cylinder Comparator<br />

B. Digital Product Phase Comparator<br />

We can easily emulate the behavior <strong>of</strong> the induction cup<br />

element using a computer as part <strong>of</strong> a digital relay. Given<br />

phasors A and B, consider the following complex product:<br />

S = A • B*<br />

= (Ax + j • Ay) • (Bx – j • By)<br />

= Ax • Bx + Ay • By + j • (Ay • Bx – Ax • By)<br />

( * = complex conjugate)<br />

The angle <strong>of</strong> the product A • B* is the same as the angle <strong>of</strong><br />

A/B, and is the angle by which A leads B.<br />

Without loss <strong>of</strong> generality, assume our phase reference is<br />

B, and phasor A leads phasor B by angle Θ. In this frame <strong>of</strong><br />

reference,<br />

Bx = │B│ By = 0<br />

Ax = │A│• cos Θ Ay = │A│• sin Θ<br />

and,<br />

S = │A│•│B│• cos Θ + j•│A│•│B│• sin Θ<br />

Separate the real and imaginary parts <strong>of</strong> S = P + j • Q =<br />

A • B*:<br />

P = │A│•│B│• cos Θ<br />

Q = │A│•│B│• sin Θ<br />

4 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


Both P and Q are two-input phase angle comparators. The<br />

P-comparator has a maximum “torque” when A and B are in<br />

phase. The Q-comparator has maximum torque when the two<br />

inputs are in quadrature. The Q-comparator is essentially the<br />

same as the induction cylinder with current inputs.<br />

In digital relays, it is easy to save the torques (P, Q). We<br />

can use the sign <strong>of</strong> the result (analogous to the cup rotation<br />

direction) as well as the magnitudes for tests involving fault<br />

type, sensitivity, etc.<br />

C. Coincidence-Timing Two-Input Phase Comparator<br />

To test the phase angle between sinusoids a and b, we can<br />

first convert a and b to square waves to derive signals A and<br />

B. The time coincidence <strong>of</strong> A and B is shown in Figure 2.<br />

We need to test the angle between δV and V p . When the<br />

angle is 90°, the relationship between δV and V p is any point<br />

on the circle, as shown in Figure 3.<br />

IX<br />

r • Z • I<br />

V = r • Z • I – V<br />

V<br />

Vpol<br />

a<br />

A<br />

Characteristic<br />

Timer, T<br />

IR<br />

AND<br />

T<br />

Output y<br />

b<br />

B<br />

Figure 3: Mho Element Derivation<br />

Figure 2:<br />

Coincidence-Timing Two-Input Comparator Logic<br />

One advantage <strong>of</strong> this comparator is the characteristic timer<br />

setting T controls the phase angle <strong>of</strong> coincidence required to<br />

obtain an output y, so we can easily synthesize mho,<br />

lenticular, and tomato characteristics.<br />

D. Application <strong>of</strong> Digital Product Phase Comparator<br />

A mho element tests the angle between a line-dropcompensated<br />

voltage and a polarizing or reference voltage.<br />

Let δV = (r • Z • I – V), where δV is the line-drop<br />

compensated voltage<br />

Z = replica line impedance<br />

r = per-unit reach in terms <strong>of</strong> the replica impedance<br />

I = measured current<br />

V = measured voltage<br />

V p = polarizing voltage.<br />

Let us test δV and V p using a digital product comparator.<br />

Which one should we use—sine or cosine Since balance (or<br />

zero torque) is at 90°, we need to use the cosine comparator,<br />

because cos 90° = 0.<br />

Let P = Re[δV • Vp*]<br />

Then: P > 0 represents the area inside the circle <strong>of</strong> reach<br />

r • Z<br />

P = 0 represents the circle itself<br />

P < 0 represents the area outside the circle <strong>of</strong> reach<br />

r • Z<br />

(Re = real portion)<br />

E. Characteristic-Mapping Approach<br />

Traditionally, one comparator is required for each zone,<br />

and for each voltage and current input combination. We can<br />

achieve significant economy in processing with no loss in<br />

performance by mapping the points on any mho circle <strong>of</strong> reach<br />

r onto a unique point on a number line.<br />

Recall the mho comparator, P:<br />

P = Re[δV • V p *]<br />

= Re[(r • Z • I – V) • V p *]<br />

For any V, I, V p combination on a circle <strong>of</strong> reach r, P is<br />

zero. This condition <strong>of</strong> balance is:<br />

0 = Re[(r • Z • I – V) • V p *]<br />

Solving for r yields in an equation which is the reach <strong>of</strong> the<br />

mho circle corresponding to the condition <strong>of</strong> balance:<br />

*<br />

Re ( V • Vp<br />

)<br />

r =<br />

(1)<br />

*<br />

Re ⎡Z • I • V ⎤<br />

⎣ p ⎦<br />

Distance Relay Element Design | 5


Observations:<br />

1. Equation 1 maps all the points on any mho circle<br />

<strong>of</strong> reach r onto a single point on the number line. If<br />

we need four mho circles, we no longer require<br />

four comparators. Instead, we simply need four<br />

tests <strong>of</strong> the calculated r.<br />

For example, a Zone 1 mho circle test might test r<br />

against 0.85, which represents a reach <strong>of</strong> 85%.<br />

Figure 4 illustrates mapping <strong>of</strong> mho circles into<br />

points for a four zone relay.<br />

2. Because V could be zero, we cannot rely on the<br />

sign <strong>of</strong> r to reliably indicate direction. Fortunately,<br />

the denominator <strong>of</strong> the r-equation is a directional<br />

element because it tests the angle between a<br />

voltage and a current. The sign <strong>of</strong> the denominator<br />

reliably indicates fault direction.<br />

The answer is simply the intersection <strong>of</strong> three two-input<br />

comparator characteristics, using pairs (X,Y), (Y,Z), and<br />

(Z,X). This arrangement is shown in the bottom <strong>of</strong> Figure 5.<br />

Refer to the coincidence timing diagram shown in the<br />

middle <strong>of</strong> Figure 5. Assume signals X and Y are slightly less<br />

than 90° apart, so we barely produce an output if X and Y<br />

were the only two inputs. Input Z does not interfere with the<br />

output as long as its leading edge is between the leading edges<br />

<strong>of</strong> X and Y. (Assuming all pulses are 180° wide, we need not<br />

make the parallel argument about the trailing edges.)<br />

The first condition is X and Y overlap by at least 90°. The<br />

two-input comparator XY represents this condition. The<br />

second condition is that the leading edge <strong>of</strong> Z lies between the<br />

leading edges <strong>of</strong> X and Y. The ZX condition ensures that Z is<br />

within ±90° <strong>of</strong> X. The YZ condition ensures that Z is within<br />

±90° <strong>of</strong> Y.<br />

If Z leads X, then we lose the YZ output. If Z lags Y then<br />

we lose the ZX output. Therefore Z must be between X and Y,<br />

which is the same condition we noted for the three-input<br />

comparator.<br />

X<br />

Y<br />

Z<br />

90°<br />

0°<br />

Output T<br />

X<br />

Y<br />

Z<br />

X<br />

Y<br />

90°<br />

90°<br />

Figure 4:<br />

Each Mho Circle Maps Onto a Point on the M-Line<br />

Y<br />

Z<br />

90°<br />

90°<br />

T<br />

III. MULTIPLE-INPUT COMPARATORS ARE REALLY A<br />

FAMILY OF TWO-INPUT COMPARATORS<br />

Multi-input comparators are widely used in distance relays.<br />

These comparators can be easily understood by representing<br />

them as several two-input comparators.<br />

The top <strong>of</strong> Figure 5 shows a three-input comparator. If<br />

inputs X, Y, and Z overlap by at least 90°, then output T<br />

asserts.<br />

What characteristic does a multiple-input comparator<br />

provide<br />

Z<br />

X<br />

90°<br />

90°<br />

Figure 5: Coincidence-Timing Multi-Input Logic<br />

IV. A QUICK REVIEW OF MHO ELEMENT<br />

POLARIZING CHOICES<br />

Mho elements compare the angle between (Z • I – V) and<br />

V p . There are many choices for the polarizing voltage, V p .<br />

Table 1 reviews some <strong>of</strong> them.<br />

6 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


TABLE 1:<br />

MHO ELEMENT POLARIZING CHOICES<br />

Operating Polarizing Characteristic General Comments<br />

(Z • I bc – V bc ) V bc<br />

(self pol.)<br />

• No expansion<br />

• Unreliable for zero-voltage faults.<br />

• Directionally insecure for reverse bus faults<br />

during high load. Requires additional<br />

directional element.<br />

(Z • I bc – V bc ) –j • V a<br />

(cross pol.<br />

w/o memory)<br />

• Good expansion for φ–φ faults.<br />

• Unreliable for zero-voltage 3φ faults.<br />

• Reverse bus fault security problems during<br />

high load periods. Requires additional<br />

directional element.<br />

(Z • I bc – V bc ) –j • V a,mem<br />

(cross pol.<br />

w/ memory)<br />

(Z • I bc – V bc ) –j • V a1mem<br />

(pos.-seq.<br />

mem. pol.)<br />

[Z • (I) –V a ]<br />

I = I a + k • I r<br />

V a<br />

(self pol.)<br />

• Good expansion for phase faults.<br />

• <strong>Reliable</strong> operation for zero-voltage 3φ faults.<br />

• Rev. φ–φ bus fault security problems during<br />

high load periods. Requires additional<br />

directional element.<br />

• Single-pole trip applications require study for<br />

pole-open security.<br />

• Greatest characteristic expansion for φ–φ and<br />

3φ faults.<br />

• <strong>Reliable</strong> operation for zero-voltage 3φ faults<br />

until pol. memory expires.<br />

• Rev. φ–φ bus fault security problems during<br />

high load periods. Requires additional<br />

directional element.<br />

• Best single-pole trip security.<br />

• No expansion.<br />

• Unreliable for zero voltage single-lineground<br />

faults.<br />

• Requires directional element.<br />

[Z • (I) –V a ]<br />

I = I a + k • I r<br />

j • V bc<br />

(cross pol.)<br />

• Good expansion.<br />

• <strong>Reliable</strong> operation reliable for zero-voltage<br />

single-line-ground faults.<br />

• Requires directional element.<br />

• Single-pole trip applications require study for<br />

pole-open security.<br />

[Z • (I) –V a ]<br />

I = I a + k • I r<br />

V a1mem<br />

(pos.-seq.<br />

mem. pol.)<br />

• Greatest expansion.<br />

• <strong>Reliable</strong> operation for zero-voltage ground<br />

faults.<br />

• Best single-pole trip security.<br />

Distance Relay Element Design | 7


The positive-sequence memory-polarized elements are<br />

generally preferred. The benefits include:<br />

• The greatest amount <strong>of</strong> expansion for improved<br />

resistive coverage. These elements always expand<br />

back to the source.<br />

• Memory action for all fault types. This is very<br />

important for close-in 3φ faults.<br />

• A common polarizing reference for all six distancemeasuring<br />

loops. This is important for single-pole<br />

tripping, during a pole-open period.<br />

V. CREATING QUADRILATERAL GROUND DISTANCE<br />

CHARACTERISTICS<br />

The quadrilateral characteristic requires four tests:<br />

• Reactance test (top line)<br />

• Positive and negative resistance tests (sides)<br />

• Directional test (bottom)<br />

A. Ground Distance Reactance Comparator<br />

A reactance element tests the angle between the line-drop<br />

compensated voltage and the polarizing current.<br />

Let δV = (r • Z • I – V), where δV is the line-drop<br />

compensated voltage<br />

Z1 = replica positive-sequence line impedance<br />

Z0 = replica zero-sequence line impedance<br />

r = per-unit reach in terms <strong>of</strong> the replica impedance<br />

I = phase current plus the residual current compensated<br />

by k = (Z0 – Z1)/3 • Z1<br />

V = measured voltage<br />

I p = polarizing current.<br />

We need to test the angle between δV and I p *. When the<br />

angle is 0°, the impedance is on the line shown in Figure 6.<br />

poor choices for the polarizing reference, because they make<br />

the reactance element severely under- or overreach, depending<br />

on the flow <strong>of</strong> load current. Negative-sequence or residual<br />

currents are appropriate polarizing choices.<br />

In some non-homogeneous system applications, the tip<br />

produced by I p may be insufficient to prevent overreach. To<br />

compensate, we can introduce an angle bias, or tip, to the<br />

reactance characteristic, or else reduce the reach <strong>of</strong> the Zone 1<br />

element.<br />

B. Ground Fault Directional Element<br />

Directional elements for ground fault must operate at fault<br />

current levels well-below the magnitude <strong>of</strong> load currents.<br />

Negative- and zero-sequence currents and voltages are mainly<br />

due to faults, and therefore are good choices for directional<br />

elements.<br />

System unbalance and measurement errors ultimately limit<br />

the sensitivity <strong>of</strong> directional elements based on negative- or<br />

zero-sequence components.<br />

Let V seq = measured sequence voltage (V 2 or V 0 )<br />

I seq = measured sequence current (I 2 or I 0 )<br />

Z = impedance whose angle adjusts I seq<br />

When the angle between –V seq and Z • I seq is 0°, the<br />

directional comparator has maximum torque. A basic<br />

negative-sequence implementation <strong>of</strong> this concept is shown in<br />

Figure 7.<br />

Figure 6:<br />

Ground Distance Reactance Element Derivation<br />

Again, using a digital product comparator, test the angle<br />

between δV and I p . The correct comparator to use is the sine<br />

comparator, since the balance point is 0°.<br />

Let Q = Im[δV • I p *]; (Im = imaginary portion)<br />

Then: Q < 0 represents the area above the line with reach<br />

r • X<br />

Q = 0 represents the line itself<br />

Q > 0 represents the area below the line <strong>of</strong> reach<br />

r • X<br />

This element must measure line reactance without adverse<br />

affects from fault resistance or load flow. Phase currents are<br />

Figure 7: Negative-Sequence Directional Element<br />

The cosine comparator gives a maximum output when the<br />

angle between the two inputs is zero degrees.<br />

Let P = Re[V seq • (Z • I seq )*];<br />

Then:<br />

P < 0 represents the area above the zero-torque line<br />

(forward)<br />

P = 0 represents the zero torque line<br />

P > 0 represents the area below the zero-torque line<br />

(reverse)<br />

C. Resistance Tests<br />

Rather than using two separate comparators, we shall apply<br />

the digital-mapping method to calculate the apparent<br />

resistance and test it against left and right side resistance<br />

thresholds.<br />

8 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


For an Aφ-ground fault on the system in Figure 8, the Aφground<br />

voltage at Bus S is:<br />

V A = m • Z1L • (I AS + k 0 • I RS ) + R AF • I F (2)<br />

Where:<br />

V A = Aφ voltage measured at Bus S<br />

m = per-unit distance to the fault from Bus S<br />

R AF = Aφ fault resistance<br />

I F = total current flowing through R F<br />

I AS = Aφ current measured at Bus S<br />

I RS = residual current measured at Bus S (3I 0S )<br />

Figure 8:<br />

System One-Line Diagram With SLG Fault<br />

The goal is to extract R AF from Equation 2. We must<br />

eliminate the line-drop voltage term, m • Z1L • (I AS + k 0 • I RS ),<br />

save the imaginary components, and solve for R AF . The result<br />

is:<br />

R<br />

Im ⎡<br />

⎢<br />

V • Z1L• I<br />

=<br />

⎣<br />

( ( + k •I ))<br />

( ( + k •I ))<br />

A AS 0 RS<br />

AF *<br />

Im ⎡I F• Z1L• IAS 0 RS<br />

⎢⎣<br />

The denominator contains I F , which includes fault and load<br />

current from both ends <strong>of</strong> the line. However, only Bus S<br />

currents are available to the relay at Bus S. We need to<br />

approximate I F in terms <strong>of</strong> Bus S current components. The<br />

approximation must be minimally system and load dependent.<br />

This last requirement permits setting the resistive thresholds<br />

with less concern that the resistive boundaries might be<br />

crossed under balanced load-flow conditions.<br />

Let I F = 3/2 • (I 2S + I 0S ), where I 2S and I 0S are the Bus S<br />

negative- and zero-sequence currents respectively. This<br />

current combination has all the available fault information,<br />

except the positive-sequence current (I 1 ). We specifically<br />

ignore I 1 because it is heavily influenced by load flow.<br />

Then:<br />

R<br />

=<br />

⎤<br />

⎥⎦<br />

⎤<br />

⎥⎦<br />

*<br />

⎡<br />

A ( ( AS<br />

+ k<br />

0•I<br />

))<br />

⎤<br />

RS<br />

⎣⎢<br />

⎥⎦<br />

( + ) ( ( + ))<br />

Im V • Z1L• I<br />

AF 3 *<br />

Im ⎡ • I I • Z1L• I k •I<br />

2 2S 0S AS 0 RS<br />

⎢⎣<br />

With this substitution, the fault resistance estimate <strong>of</strong><br />

Equation 3 is independent <strong>of</strong> balanced load. The 3/2 scale<br />

factor accounts for the missing I 1 contribution and ensures R AF<br />

measures the true fault resistance on a radial system. Infeed<br />

from Bus R causes R AF to increase, because our substitution<br />

for I F does not include any measurement <strong>of</strong> current from<br />

Bus R. For example, if the impedances on either side <strong>of</strong> the<br />

fault are equal, R AF is half the actual fault resistance.<br />

This method provides an easy means <strong>of</strong> testing R AF for<br />

both the left and right sides <strong>of</strong> the quadrilateral element:<br />

calculate R and test the result against ±R thresholds for each<br />

zone.<br />

*<br />

⎤<br />

⎥⎦<br />

(3)<br />

For example, a Zone 1 resistive boundary might test R AF<br />

against ±2 Ω. If the result is 1 Ω, this satisfies the criteria set<br />

for the Zone 1 quadrilateral resistive checks.<br />

VI. MAINTAINING DIRECTIONAL SECURITY<br />

Directional security is paramount. At first glance, mho<br />

elements appear directional. However, some safeguards are<br />

required to ensure security.<br />

A. Ground Direction Security Concerns<br />

Reverse Ground Faults: The operating quantities for all<br />

ground distance elements include residual current. For<br />

example, the residual current produced by a reverse Aφ<br />

ground fault is also used in the phase-ground distance<br />

elements for B and C phases. The residual current can cause a<br />

forward-reaching Bφ or Cφ ground distance element to<br />

operate. We can avoid this problem by supervising the ground<br />

distance elements with a directional element, by a phaseselection<br />

comparator, or by introducing additional conditions<br />

in a multiple-input comparator.<br />

B. Selection <strong>of</strong> Directional Element Input Quantities<br />

Negative-sequence directional elements have notable<br />

advantages:<br />

• Insensitivity to zero-sequence mutual coupling.<br />

• There is generally more negative-sequence current<br />

than zero-sequence current for remote ground faults<br />

with high fault resistance. This allows higher<br />

sensitivity with reasonable and secure sensitivity<br />

thresholds.<br />

• Insensitivity to vt neutral shift, possibly caused by<br />

multiple grounds on the vt neutral.<br />

Perhaps the major disadvantage is negative-sequence<br />

elements are rendered useless when one or two poles <strong>of</strong> the<br />

breaker are open.<br />

C. Compensated Negative-Sequence Directional Element<br />

When the negative-sequence source behind the relay<br />

terminal is very strong, the negative-sequence voltage (V2) at<br />

the relay can be very low, especially for remote faults.<br />

To overcome low V2 magnitude, we can add a<br />

compensating quantity which boosts V2 by (α • ZL2 • I2).<br />

The constant α controls the amount <strong>of</strong> compensation.<br />

Equation 4 shows the torque equation for a compensated<br />

negative-sequence directional element.<br />

T32Q = Re[(V2 – α • ZL2 • I2) • (ZL2 • I2)*] (4)<br />

The term (α • ZL2 • I2) adds with V2 for forward faults,<br />

and subtracts for reverse faults. Setting α too large can make a<br />

reverse fault appear forward. This results when (α • ZL2 • I2)<br />

is greater but opposed to the measured V2 for reverse faults.<br />

Distance Relay Element Design | 9


D. Relationship <strong>of</strong> the Apparent Z2 to Fault Direction<br />

The sequence network for a ground fault at the relay bus is<br />

shown in Figure 9. The relay measures IS2 for forward faults,<br />

and –IR2 for reverse faults.<br />

ES<br />

Positive<br />

Sequence<br />

Negative<br />

Sequence<br />

Zero<br />

Sequence<br />

Figure 9:<br />

ZS1<br />

ES<br />

Relay<br />

ZL1<br />

ZS1 ZL1 ZR1<br />

ZR1<br />

IS2<br />

IR2<br />

V2<br />

ZS2 ZL2 ZR2<br />

ZS0 ZL0 ZR0<br />

RF<br />

Sequence Network for a Reverse Single-Line-Ground Fault<br />

From V2 and I2, calculate Z2:<br />

−V2<br />

Forward SLG Faults: Z2 = = − ZS2<br />

IS2<br />

−V2<br />

Reverse SLG Faults: Z2 = = ( ZL2 + ZR2)<br />

−IR2<br />

This relationship is shown in Figure 10 for a 90° system.<br />

S<br />

Reverse Fault<br />

ZS<br />

Reverse<br />

Forward<br />

RF<br />

Relay<br />

RF<br />

ZL<br />

Z2 Impedance Plane<br />

+X2<br />

Forward Fault<br />

ZS2<br />

ZR<br />

ZR2 + ZL2<br />

X2 = 0<br />

Figure 10: Measured Negative-Sequence Impedance Yields Direction<br />

For the system in Figure 10, the fault is forward if Z2 is<br />

negative, and reverse if Z2 is positive.<br />

ER<br />

R<br />

ER<br />

E. Negative-Sequence Directional Element Based on<br />

Calculating and Testing Z2<br />

The discussion above shows that calculated Z2 could be<br />

used to determine fault direction.<br />

Recall the compensated negative-sequence directional<br />

element equation, T32Q:<br />

T32Q = Re[(V2 – α • ZL2 • I2) • (ZL2 • I2)*]<br />

The forward/reverse balance condition for this element is<br />

zero torque. This is:<br />

0 = Re[(V2 – α • ZL2 • I2) • (ZL2 • I2)*]<br />

Let α = z2<br />

ZL2 = 1∠Θ where Θ is the angle <strong>of</strong> ZL2<br />

Substituting,<br />

0 = Re[(V2 – z2∠Θ I2) • (I2 • 1∠Θ)*]<br />

Solving for z2 results in an equation corresponding to the<br />

condition <strong>of</strong> zero-torque:<br />

Re[V2 • (I2 • I ∠Θ)*]<br />

z2 =<br />

Re[(I2 •1 ∠Θ) • (I2 •1 ∠Θ)*]<br />

Re[V2 • (I2 •1 ∠Θ)*]<br />

z2 =<br />

2<br />

I2<br />

Recall the (α • ZL2 • I2) term increases the amount <strong>of</strong> V2<br />

for directional calculations. This is equivalent to increasing<br />

the magnitude <strong>of</strong> the negative-sequence source behind the<br />

relay location. This same task is accomplished by increasing<br />

the forward z2 threshold.<br />

The criteria for declaring forward and reverse faults are<br />

then:<br />

If<br />

z2 < forward threshold, then the fault is forward<br />

z2 > reverse threshold, then the fault is reverse<br />

The forward threshold must be less than the reverse<br />

threshold to avoid any overlap.<br />

The z2 directional element has all <strong>of</strong> the benefits <strong>of</strong> both<br />

the traditional and compensated negative-sequence directional<br />

element. It also provides better visualization <strong>of</strong> how much<br />

compensation is secure and required. Set the for-ward and<br />

reverse impedance thresholds based upon the strongest source<br />

conditions.<br />

F. Phase-Phase Distance Element Security for Reverse<br />

Phase-Phase Faults<br />

Phase-phase distance elements use phase-phase currents.<br />

For example, a BC phase-phase distance element uses I BC or<br />

(I B – I C ). For a close-in reverse CA fault, the Cφ current can<br />

cause operation <strong>of</strong> the forward-reaching BC element. An easy<br />

way to avoid this risk is by supervising the phase-phase<br />

distance elements with the negative-sequence direction<br />

element just described. (The negative-sequence directional<br />

element is ignored for 3φ faults which pick up all three phasephase<br />

distance elements.)<br />

G. Phase-Phase Distance Element Security for Reverse<br />

Three-Phase Faults<br />

Phase distance elements require memory polarization to be<br />

secure and reliable for reverse three-phase (3φ) faults.<br />

10 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


The most onerous 3φ fault is one with the following<br />

qualifications:<br />

1. A small critical amount <strong>of</strong> fault resistance.<br />

2. Significant load flow into the bus from a weaker<br />

source.<br />

Three-phase faults are a concern to phase distance elements<br />

only after the memory expires. For bolted faults, each phase<br />

voltage is zero. Once the memory expires, the distance<br />

elements are disabled. However, with some resistance, the<br />

polarization voltage does not go to zero, and could move to an<br />

angle permitting tripping.<br />

Figure 11(a) illustrates a system with a reverse 3φ fault<br />

with fault resistance. Figure 11(b) shows the total Aφ fault<br />

current (I F,a ), the prefault memory voltage (V a,mem ), Aφ voltage<br />

and current seen by the relay (V af and I RLY,a respectively), and<br />

I RLY,a adjusted by the replica line impedance.<br />

Recall the denominator term <strong>of</strong> Equation 1 for a phase<br />

distance element. This denominator term is a directional<br />

element. It indicates how a phase distance element <strong>of</strong> infinite<br />

reach would perform for this same fault. If the angle between<br />

Z • I and V p is less than 90°, the phase distance element<br />

declares the fault forward. From Figure 11(b), the angle<br />

between I RLY,a • Z and V a is less than 90°. Thus, after the<br />

memory voltage (V a,mem ) becomes in step with V af , the<br />

directional security <strong>of</strong> the phase distance element is<br />

compromised.<br />

Figure 11(c) illustrates the same reverse 3φ fault with load<br />

flowing out from the relay terminal (from E S towards E R ). For<br />

this case, the phase distance relay is secure even after the<br />

memory voltage becomes in-step with the fault voltage.<br />

Solutions:<br />

1. Clear the 3φ fault before the memory expires.<br />

2. The apparent impedance for the reverse 3φ-G fault<br />

enters the tripping characteristic in the secondquadrant.<br />

Reducing the maximum torque angle<br />

reduces the amount <strong>of</strong> second-quadrant coverage.<br />

3. Require an increased torque magnitude output from<br />

the comparator (i.e., desensitize the element). As the<br />

apparent impedance enters the tripping characteristic<br />

very near the origin, forward direction 3φ faults<br />

produce greater torques than do the reverse 3φ faults.<br />

4. Add current <strong>of</strong>fset in the positive direction to the<br />

polarizing reference.<br />

5. Test the angle between the positive-sequence current<br />

and voltage (i.e., use a positive-sequence directional<br />

element). This angular test limits the three-phase fault<br />

coverage to a 180° impedance angle sector <strong>of</strong> –120° to<br />

60°.<br />

VII. LOAD ENCROACHMENT<br />

The impedance <strong>of</strong> heavy loads can actually be less than the<br />

impedance <strong>of</strong> some faults. Yet, the protection must be made<br />

selective enough to discriminate between load and fault<br />

conditions. Unbalance aids selectivity for all faults except<br />

three-phase faults.<br />

Figure 12 shows the load-encroachment characteristics in<br />

the impedance plane.<br />

Figure 11:<br />

Reverse 3φ Fault Conditions and Mho Element Performance<br />

Figure 12: Load Encroachment on Mho Distance Element Characteristics<br />

When power flows out, the load impedance is in the<br />

wedge-shaped load-impedance area to the right <strong>of</strong> the X-axis.<br />

When power flows in, the load impedance is in the left-hand<br />

load-impedance area.<br />

There is overlap (shaded solid) between the mho circle and<br />

the load areas. Should the load impedance lie in the shaded<br />

area, the impedance relay will detect the under-impedance<br />

condition and trip the heavily-loaded line. Such protection<br />

unnecessarily limits the load-carrying capability <strong>of</strong> the line.<br />

For better load rejection, the mho circle can be squeezed<br />

into a lenticular or elliptical shape. Unfortunately, this also<br />

reduces the fault coverage.<br />

Distance Relay Element Design | 11


Alternatively, we could use additional comparators to make<br />

blinders parallel to the transmission line characteristic, to limit<br />

the impedance-plane coverage, and exclude load from the<br />

tripping characteristic.<br />

Or, we could build quadrilateral characteristics, which boxout<br />

load.<br />

All traditional solutions have the same common approach:<br />

shape the operating characteristic <strong>of</strong> the relay to avoid load.<br />

The traditional solutions have two major disadvantages:<br />

1. Reducing the size <strong>of</strong> the relay characteristic<br />

desensitizes the relay to faults with resistance.<br />

Avoiding a small area <strong>of</strong> load encroachment <strong>of</strong>ten<br />

requires sacrificing much larger areas <strong>of</strong> fault<br />

coverage.<br />

2. From a user's point <strong>of</strong> view, the more complex shapes<br />

become hard to define, and the relays are harder to set.<br />

A new approach does not modify the relay characteristic<br />

shape directly. Instead, it defines the load regions in the<br />

impedance plane, and blocks operation <strong>of</strong> distance elements if<br />

the impedance is in either <strong>of</strong> the load regions.<br />

Figure 13 shows the new approach applied to a four-zone<br />

mho relay. The mho characteristics are conventional, and are<br />

not modified to exclude load.<br />

Figure 13:<br />

Improved Load-Encroachment Method<br />

There are two load regions shown (load-in and load-out).<br />

The relay calculates the complex positive-sequence impedance<br />

and tests it against the boundaries <strong>of</strong> these load regions. If the<br />

impedance is inside either region, then the relay concludes the<br />

impedance represents load, and blocks the mho elements. If<br />

the impedance is outside both load regions, the mho elements<br />

are permitted to operate.<br />

The advantages <strong>of</strong> the new approach include:<br />

1. Greater coverage for faults is possible because only<br />

the overlap between the load characteristic and the<br />

relay characteristic is blocked.<br />

2. The load-encroachment characteristics are easy to set,<br />

because they can be directly related to the maximum<br />

load conditions. Once we have maximum load-in and<br />

load-out conditions, and the range <strong>of</strong> power factors <strong>of</strong><br />

the load, the load-impedance areas can be found. No<br />

customization <strong>of</strong> relay characteristics is required to<br />

avoid load.<br />

3. Separate characteristics for load-in and load-out are<br />

easy to define.<br />

4. Since the characteristics for faults and loads are<br />

independent, there is less chance <strong>of</strong> setting errors.<br />

The logic applies to three-phase faults only. Loadencroachment<br />

blocking is not required or desired for<br />

unbalanced faults (e.g., AB, BC, CA, AG, BG, CG, ABG,<br />

BCG, CAG faults).<br />

VIII. FAULT-TYPE <strong>SEL</strong>ECTION CONSIDERATIONS<br />

For security, distance relay schemes must consider the<br />

behavior <strong>of</strong> the distance elements in all six fault loops (AG,<br />

BG, CG, AB, BC, and CA) under very broad and general<br />

system, load, and fault conditions.<br />

There are two major concerns:<br />

1. Ground distance elements can overreach for line-lineground<br />

(LLG) faults.<br />

2. Phase distance elements can operate for close-in lineground<br />

(LG) faults.<br />

The first concern is generally considered a problem in all<br />

applications. The second concern is a problem in single-poletrip<br />

schemes, and a targeting nuisance.<br />

How can we reliably prevent unwanted relay elements<br />

from interfering with the performance <strong>of</strong> the overall scheme<br />

If we know the fault is an AG fault, then we can block the<br />

AB and CA elements in order to avoid a three-pole trip for a<br />

LG fault.<br />

If we know the fault is a BCG fault, then we can block the<br />

BG and CG elements, avoiding possible overreach by the BG<br />

and CG elements.<br />

(The BG element tends to overreach for a BCG fault with<br />

resistance to ground. The CG element tends to overreach for a<br />

BCG fault with resistance between the phases.)<br />

A. Selection Using the Angle Between I 0 and I 2<br />

The angle between the negative-sequence current and the<br />

zero-sequence current is a frequently used and very useful<br />

indicator.<br />

Figure 14 shows the sequence networks for AG and BCG<br />

faults. The symmetrical component currents are referenced to<br />

phase A. That is, I 0 means I A0 , and I 2 means I A2 . For these two<br />

faults, the angle between I 2 and I 0 is zero degrees. Figure 15<br />

shows phasor diagrams for AG, BG, and CG faults. It also<br />

shows the phase relationships between I 0 and I 2 for the three<br />

faults.<br />

12 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


If the angle is near +120°, then the fault is CG or ABG.<br />

• Enable CG and AB elements only.<br />

• One approach is to assign 120° sectors to AG, BG and<br />

CG faults. For example, angles between ±60° belong<br />

to AG and BCG faults.<br />

B. Fault Resistance Can Affect This Scheme<br />

Figure 16 shows the effects <strong>of</strong> introducing fault resistance<br />

between the point where phases B and C are shorted together<br />

and ground. Rf appears in the zero-sequence net-work, and the<br />

angle <strong>of</strong> the I 0 leads the angle <strong>of</strong> I 2 , because the network zerosequence<br />

impedance angle is less than that <strong>of</strong> the negativesequence<br />

impedance angle.<br />

Figure 14:<br />

Resistance)<br />

I 0 and I 2 Relationship for AG and BCG Faults (Without Fault<br />

Figure 15:<br />

I 0 and I 2 Relationship for AG, BG, and CG Faults<br />

Given that the fault is a single-line-ground (SLG) fault, the<br />

angle between I 0 and I 2 in the fault is a very reliable indicator<br />

<strong>of</strong> the fault type: centered around zero degrees for AG, –120°<br />

for BG (I 2 lags I 0 by 120°), and +120° for CG (I 2 leads I 0 by<br />

120°). Fortunately, these angles do not change much over a<br />

broad range <strong>of</strong> system conditions.<br />

Thinking about Figures 14 and 15 at the same time, we<br />

conclude:<br />

If the angle is near zero, then the fault is AG or BCG.<br />

• If it is AG, then AB or CA could also pick up and<br />

three-pole trip for a single-line-ground fault.<br />

• If it is BCG, then the BG and CG ground distance<br />

elements may overreach.<br />

• THEREFORE: Enable AG, BC elements only.<br />

If the angle is near –120°, then the fault is BG or CAG.<br />

Enable BG and CA elements only.<br />

Figure 16: Effects <strong>of</strong> Increasing Rf for a BCG Fault<br />

When Rf is big enough, there may be confusion as to<br />

whether the fault is AG/BCG or BG/CAG, because the angle<br />

could be less than –60°, as shown in the bottom <strong>of</strong> the figure.<br />

Distance Relay Element Design | 13


1) Estimating and Comparing Fault Resistances Solves the<br />

Problem<br />

When the angle itself is not conclusive, e.g., when the<br />

angle is more than 30° from its expected value, we can<br />

compare phase and ground fault resistance estimates, and<br />

select the fault type associated with the minimum resistance.<br />

For example, and again referring to the bottom <strong>of</strong> Figure 16, a<br />

comparison <strong>of</strong> an estimate <strong>of</strong> Rbg against the minimum phaseto-phase<br />

fault resistance (in this case, Rbc), would reveal that<br />

Rbg is much larger than Rbc. Therefore, the logic concludes<br />

the fault mainly involves phases B and C, and the best<br />

measurement is made by the BC element.<br />

IX. SUMMARY<br />

Important points presented in this paper include:<br />

1. A review <strong>of</strong> several types <strong>of</strong> phase-angle comparators<br />

shows their similarities and differences.<br />

2. A multiple-input comparator can be viewed as a<br />

family <strong>of</strong> two-input comparators, which includes<br />

every pair <strong>of</strong> inputs.<br />

3. Writing the torque equation for a relay characteristic,<br />

setting it equal to zero, and solving for a scalar<br />

multiple <strong>of</strong> the element reach yields a result which<br />

maps points on the characteristic onto a single point<br />

on the number line. This makes for very efficient<br />

synthesis <strong>of</strong> a family <strong>of</strong> characteristics with different<br />

reaches.<br />

4. Positive-sequence memory potential is generally the<br />

most secure and reliable polarization method for mho<br />

characteristics.<br />

5. A negative-sequence element which calculates<br />

negative-sequence impedance, and tests it against<br />

thresholds is presented. It is easy to adapt to virtually<br />

any system conditions, because its two threshold<br />

settings relate to system impedances.<br />

6. We present ground-fault-resistance elements for<br />

quadrilateral characteristics, which estimate the fault<br />

resistance, and use the fault-current estimate<br />

1.5 • (I0 + I2). This estimate rejects the loadinfluenced<br />

positive-sequence current, but includes the<br />

rest <strong>of</strong> the fault current.<br />

7. A negative-sequence directional check is sufficient to<br />

overcome security problems with phase and ground<br />

distance elements for unbalanced faults.<br />

8. A positive-sequence directional element can be used<br />

to cut <strong>of</strong>f part <strong>of</strong> the mho characteristics in the second<br />

quadrant <strong>of</strong> the impedance plane, to enhance security<br />

for three-phase faults. This element would normally be<br />

set with a maximum torque angle much less than that<br />

<strong>of</strong> the mho circle.<br />

9. Instead <strong>of</strong> shaping impedance-plane tripping<br />

characteristics to avoid load, we introduced a load<br />

characteristic. Impedances must leave the load<br />

characteristic, before tripping is permitted for threephase<br />

faults. The load characteristic looks like a bowtie<br />

without the knot, and therefore neatly surrounds<br />

load areas. It minimizes the area <strong>of</strong> the impedance<br />

plane eliminated for load considerations.<br />

10. We point out a potential problem with fault-type<br />

selectors based on the angle between I0 and I2. They<br />

may err for certain resistive LLG faults. The solution<br />

is a comparison <strong>of</strong> estimates <strong>of</strong> LL and LG fault<br />

resistances.<br />

X. REFERENCES<br />

[1] R.J. Martilla, “Directional Characteristics <strong>of</strong> Distance Relay Mho<br />

Elements: Part II - Results.” IEEE Trans-actions on <strong>Power</strong> Apparatus<br />

and Systems, Vol. PAS-100, No.1 January 1981.<br />

[2] Edmund O. Schweitzer III, “New Developments in Distance Relay<br />

Polarization and Fault Type Selection,” 16th Annual Western Protective<br />

Relay Conference, October 24 - 26, 1989, Spokane, WA.<br />

[3] A.R. Van C. Warrington, “Protective Relays: Their Theory and<br />

Practice.” Chapman and Hall, 1969, Volume I and II.<br />

XI. BIOGRAPHIES<br />

Edmund O. Schweitzer, III received BSEE and MSEE degrees from Purdue<br />

University in 1968 and 1971, respectively. He earned a PhD at Washington<br />

State University (WSU) in 1977.His pr<strong>of</strong>essional experience includes<br />

electrical engineering work at Probe Systems in California and the National<br />

Security Agency in Maryland. He served as an assistant pr<strong>of</strong>essor at Ohio<br />

University and as an assistant and an associate pr<strong>of</strong>essor at WSU. Since 1983,<br />

he has directed the activities <strong>of</strong> Schweitzer Engineering Laboratories, Inc.<br />

(<strong>SEL</strong>), the company he founded in Pullman, Washington. <strong>SEL</strong> designs,<br />

manufactures, and markets digital protective relays for power system<br />

protection.<br />

Schweitzer started investigating digital relays during PhD studies at WSU in<br />

1976, which produced both his doctoral dissertation and <strong>SEL</strong>. His university<br />

research was supported by Bonneville <strong>Power</strong> Administration, Electric <strong>Power</strong><br />

Research Institute, and various utilities. Although the company has grown<br />

significantly, Schweitzer is still involved in the development <strong>of</strong> new relays<br />

and auxiliary equipment.<br />

He is a Fellow <strong>of</strong> the Institute <strong>of</strong> Electrical and Electronic Engineers (IEEE),<br />

is a member <strong>of</strong> Eta Kappa Nu and Tau Beta Pi, and has authored or coauthored<br />

30 technical papers.<br />

Jeff B. Roberts received his BSEE from Washington State University in<br />

1985. He worked for Pacific Gas and Electric Company as a relay protection<br />

engineer for over three years. In November, 1988 he joined Schweitzer<br />

Engineering Laboratories, Inc. as an Application Engineer. He now serves as<br />

Application Engineering Supervisor. He has delivered papers at the Western<br />

Protective Relay Conference, Texas A&M University and the Southern<br />

African Conference on <strong>Power</strong> System Protection. He holds one patent on the<br />

<strong>SEL</strong>-121B relay and has other patents pending.<br />

Copyright © <strong>SEL</strong> 1992, 1993<br />

(All rights reserved.)<br />

Printed in USA<br />

Rev. 2<br />

14 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


Impedance-Based Fault Location Experience<br />

Karl Zimmerman and David Costello, Schweitzer Engineering Laboratories, Inc.<br />

I. INTRODUCTION<br />

Accurate fault location reduces operating costs by avoiding<br />

lengthy and expensive patrols. Accurate fault location<br />

expedites repairs and restoration <strong>of</strong> lines, ultimately reducing<br />

revenue loss caused by outages.<br />

In this paper, we describe one- and two-ended impedancebased<br />

fault location experiences. We define terms associated<br />

with fault location, and describe several impedance-based<br />

methods <strong>of</strong> fault location (simple reactance, Takagi, zerosequence<br />

current with angle correction, and two-ended<br />

negative-sequence). We examine several system faults and<br />

analyze the performance <strong>of</strong> the fault locators given possible<br />

sources <strong>of</strong> error (short fault window, nonhomogeneous<br />

system, incorrect fault type selection, etc.).<br />

Finally, we show the laboratory testing results <strong>of</strong> a twoended<br />

method, where we automatically extracted a two-ended<br />

fault location estimate from a single end.<br />

II. FAULT LOCATION METHODS AND DEFINITIONS<br />

Several methods <strong>of</strong> estimating fault location are presently<br />

used in the field:<br />

• DFR and short circuit data match<br />

• Traveling wave methods<br />

• Impedance-based methods<br />

− One-ended methods without using source<br />

impedance data (simple reactance, Takagi)<br />

− One-ended methods using source impedance data<br />

• Two-ended methods<br />

In this paper, we focus on certain impedance-based fault<br />

location methods and provide results from actual system<br />

faults.<br />

III. NOTABLE IEEE DEFINITIONS<br />

IEEE PC37.114, “Draft Guide for Determining Fault<br />

Location on AC Transmission and Distribution Lines”[1] was<br />

recently balloted and is in the approval process. One <strong>of</strong> the<br />

important contributions <strong>of</strong> the guide is the definitions section.<br />

Here are a few notable definitions found in the guide:<br />

Fault location error: Percentage error in fault location<br />

estimate based on the total line length: e (error) = (instrument<br />

reading – exact distance to the fault) / total line length.<br />

For example, suppose a line is 100 miles long and the<br />

actual fault is 90 miles from the local terminal. If the local<br />

fault locator provides a fault location <strong>of</strong> 94 miles, the fault<br />

location error is (94–90)/100 = 4%. If the remote fault locator<br />

indicates 8 miles, the fault location error is (8–10)/100 = 2%.<br />

Homogeneous line: A transmission line where impedance<br />

is distributed uniformly on the whole length.<br />

Examples <strong>of</strong> this are lines that use the same conductor size<br />

and construction throughout. Lines that are nonhomogeneous<br />

can be a source <strong>of</strong> error for one- or two-ended impedancebased<br />

fault location methods.<br />

Homogeneous system: A transmission system where the<br />

local and remote source impedances have the same system<br />

angle as the line impedance. A homogeneous system is shown<br />

in Figure 1.<br />

Z<br />

Z<br />

Figure 1<br />

1S<br />

0S<br />

Z 1S<br />

= 2∠80°<br />

= 3 • Z<br />

1S<br />

Relay<br />

mZ 2L<br />

Z<br />

Z<br />

1L<br />

0L<br />

R F<br />

(1-m)Z 2L<br />

= 8∠80°<br />

= 3• Z<br />

Example <strong>of</strong> a Homogeneous System<br />

1L<br />

Relay<br />

∠Z<br />

∠Z<br />

1R<br />

0R<br />

Z 1R<br />

= ∠Z<br />

1S<br />

= ∠Z<br />

Nomograph: A graph that plots measured fault location<br />

versus actual fault location by compensating for known<br />

system errors.<br />

Figure 2 shows a 69 kV line with 12.47 kV underbuild.<br />

Figure 2<br />

Relay<br />

Location<br />

N<br />

69 kV Line Configuration Sketch<br />

N<br />

69 kV<br />

0S<br />

12.47 kV<br />

Load<br />

How to build a nomograph:<br />

1. Calculate line constants.<br />

2. Determine which faults require a nomograph.<br />

3. Using short circuit program, apply faults along the<br />

length <strong>of</strong> line (10%, 20%, etc.).<br />

4. Plug resultant voltage and current values into fault<br />

location algorithms.<br />

5. Plot a short circuit (actual) vs. calculated (relay)<br />

fault location.<br />

Impedance-Based Fault Location Experience | 15


Without Underbuild<br />

With 50 Miles Underbuild<br />

R1 7.50 Ω 7.50 Ω<br />

X1 22.757 Ω 22.757 Ω<br />

R0 21.327 Ω 15.488 Ω<br />

X0 134.16 Ω 88.75 Ω<br />

Figure 3 shows a completed nomograph.<br />

Figure 3<br />

Printout Miles<br />

69 kV Nomograph<br />

Partial, Total, and No Underbuild<br />

60<br />

55<br />

50<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 10 20 30 40 50 60<br />

Actual Miles<br />

10 Miles Underbuild No Underbuild 60 Miles Underbuild<br />

Completed Nomograph for 69 kV Line<br />

IV. IMPEDANCE-BASED FAULT LOCATION<br />

METHODS AND REQUIREMENTS<br />

Impedance-based methods require the following approach:<br />

1. Measure the voltage and current phasors.<br />

2. Extract the fundamental components.<br />

3. Determine the phasors and fault type.<br />

4. Apply impedance algorithm.<br />

One-ended impedance methods <strong>of</strong> fault location are a<br />

standard feature in most numerical relays. One-ended<br />

impedance methods use a simple algorithm, and<br />

communication channels and remote data are not required<br />

(except when a channel is required to bring the fault location<br />

estimate to an operator).<br />

Two-ended methods can be more accurate but require data<br />

from both terminals. Data must be captured from both ends<br />

before an algorithm can be applied.<br />

The most popular impedance-based fault location methods<br />

are discussed in this paper:<br />

• Simple reactance method (one-ended)<br />

• Takagi method (one-ended)<br />

• Modified Takagi method that corrects for source<br />

impedance angle differences (one-ended)<br />

• Two-ended negative-sequence method<br />

One-ended impedance-based fault locators calculate the<br />

fault location from the apparent impedance seen by looking<br />

into the line from one end. An example system one-line is<br />

shown in Figure 4. To locate all fault types, the phase-toground<br />

voltages and currents in each phase must be measured.<br />

(If only line-to-line voltages are available, it is possible to<br />

locate phase-to-phase faults; if the zero-sequence source<br />

impedance, Z 0 , is known, we can estimate the location for<br />

phase-to-ground faults).<br />

If the fault resistance is assumed to be zero, we can use one<br />

<strong>of</strong> the impedance calculations in Table 1 to estimate the fault<br />

location.<br />

Fault Type<br />

TABLE 1<br />

SIMPLE IMPEDANCE EQUATIONS<br />

Positive-Sequence Impedance<br />

Equation (mZ 1L =)<br />

A–ground Va ( Ia + k • 3 • I0<br />

)<br />

B–ground Vb ( Ib + k • 3 • I0<br />

)<br />

C–ground V ( I + k • 3 • I )<br />

a–b or a–b–g<br />

b–c or b–c–g<br />

c–a or c–a–g<br />

c c 0<br />

a–b–c Any <strong>of</strong> the following: Vab I<br />

ab, Vbc I<br />

bc, Vca I<br />

ca<br />

Where:<br />

k is (Z 0L – Z 1L ) / 3Z 1L ,<br />

Z 0L is the zero-sequence line impedance,<br />

Z 1L is the positive-sequence line impedance,<br />

m is the per unit distance to fault (for example:<br />

distance to fault in kilometers divided by the total<br />

line length in kilometers),<br />

is the zero-sequence current.<br />

I 0<br />

S<br />

Z S<br />

Figure 4<br />

VS<br />

VS<br />

I S<br />

m<br />

mZ L<br />

Z L<br />

V<br />

V<br />

V<br />

I S<br />

I R<br />

VR<br />

I F<br />

R F<br />

ab<br />

bc<br />

ca<br />

I<br />

I<br />

I<br />

1-m<br />

ab<br />

bc<br />

ca<br />

(1-m)Z L<br />

One-Line Diagram and Circuit Representation <strong>of</strong> Line Fault<br />

The challenges for accuracy <strong>of</strong> one-ended fault location are<br />

well known and are described in several sources [1] [2] [3]<br />

[4]. To summarize, the following conditions can cause errors<br />

for one-ended impedance-based fault location methods:<br />

• Combined effect <strong>of</strong> fault resistance and load<br />

• Zero-sequence mutual coupling<br />

• Zero-sequence modeling errors<br />

• System nonhomogeneity<br />

• System infeeds<br />

− Remote or third terminal infeed<br />

− Tapped load with zero-sequence source<br />

• Inaccurate relay measurement, instrument transformer<br />

or line parameters.<br />

I R<br />

VR<br />

R<br />

Z R<br />

16 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


A. Simple Reactance Method<br />

From Figure 4, the voltage drop from the S end <strong>of</strong> the line<br />

is:<br />

VS = m • Z1L • IS + R F • IF<br />

(1)<br />

For an A-phase to ground fault, Vs = V a−g and I S = Ia + k • 3• I0<br />

The goal is to minimize the effect <strong>of</strong> the R F • IF<br />

term.<br />

The simple reactance method divides all terms by I S (I<br />

measured at the fault locator) and ignores the (R F •I F / I S ) term.<br />

To do this, save the imaginary part, and solve for m:<br />

( )<br />

Im V /I = Im(m • Z ) = m • X<br />

S S 1L 1L<br />

m =<br />

I<br />

m<br />

⎛ VS<br />

⎞<br />

⎜<br />

⎝ I ⎟<br />

⎠<br />

X<br />

1L<br />

Error is 0 if ∠ I = ∠ I or R = 0<br />

S<br />

S F F<br />

B. Takagi Method—One-Ended Impedance Method With No<br />

Source Data<br />

The Takagi method requires prefault and fault data. It<br />

improves upon the simple reactance method [2] by reducing<br />

the effect <strong>of</strong> load flow and minimizing the effect <strong>of</strong> fault<br />

resistance.<br />

VS = m • Z 1L • I + R F • IF<br />

(3)<br />

Use Superposition current (I sup ) to find a term in phase with<br />

I F :<br />

Isup<br />

= I − Ipre<br />

I = Fault Current<br />

(4)<br />

I = Pre-fault Current<br />

pre<br />

Voltage drop from Bus S:<br />

VS = m • Z 1L • IS + R F • IF<br />

Multiply both sides <strong>of</strong> equation (1) by the complex<br />

conjugate <strong>of</strong> I sup (I sup* ) and save the imaginary part. Then,<br />

solve for m:<br />

I m[VS • I sup* ] = m • I m (Z1L • IS • I sup* ) + R F • I m (IF • I sup* )<br />

I m (VS • I sup* )<br />

(5)<br />

m = I (Z • I • I )<br />

m 1L S sup*<br />

The key to the success <strong>of</strong> the Takagi method is that the<br />

angle <strong>of</strong> I S is the same as the angle <strong>of</strong> I F . For an ideal<br />

homogeneous system, these angles are identical. As the angle<br />

between I S and I F increases, the error in the fault location<br />

estimate increases.<br />

(2)<br />

C. Modified Takagi—Zero-Sequence Current Method with<br />

Angle Correction<br />

Another method (modified Takagi) uses zero-sequence<br />

current (3 • I 0S ) for ground faults instead <strong>of</strong> the superposition<br />

current. Therefore, this method requires no prefault data.<br />

Modified Takagi also allows for angle correction. If the<br />

user knows the system source impedances, the zero-sequence<br />

current can be adjusted by angle T to improve the fault<br />

location estimate for a given line.<br />

* − jT<br />

( ( ) )<br />

−<br />

( )<br />

Im V S • 3• I 0S • e<br />

m = (6)<br />

I Z • I • 3• I • e<br />

* jT<br />

( )<br />

m 1L S 0S<br />

The angle T selected will be valid for one fault location<br />

along the line. Figure 5 shows how to calculate T.<br />

Z 0S<br />

Figure 5<br />

known)<br />

3•I RS<br />

mZ 0L<br />

T<br />

3• I RS<br />

I F<br />

I F<br />

(1-m)Z 0L<br />

Zero-Sequence Current Angle Correction (if source impedances are<br />

IF<br />

Z0S + Z0L + Z0R<br />

= = A∠T<br />

3• I (1− m) • Z + Z<br />

RS 0L 0R<br />

D. Two-Ended Negative-Sequence Impedance Method<br />

A relatively new method, introduced in 1999, uses<br />

negative-sequence quantities from all line terminals for the<br />

location <strong>of</strong> unbalanced faults. By using negative-sequence<br />

quantities, we negate the effect <strong>of</strong> prefault load and fault<br />

resistance, zero-sequence mutual impedance, and zerosequence<br />

infeed from transmission line taps. Precise fault type<br />

selection is not necessary. Data alignment is not required<br />

because the algorithm employed at each line end uses the<br />

following quantities from the remote terminal (which do not<br />

require phase alignment).<br />

• Magnitude <strong>of</strong> negative-sequence current, I2<br />

• Calculated negative-sequence source impedance,<br />

Z ∠θ<br />

°<br />

2<br />

2<br />

Z 0R<br />

(7)<br />

Impedance-Based Fault Location Experience | 17


An observation from Figure 6 is that the negative-sequence<br />

is the same when viewed from all ends <strong>of</strong><br />

fault voltage ( V2F<br />

)<br />

the protected line.<br />

Source<br />

S<br />

Z 1S<br />

Z 2S<br />

Z 0S<br />

I 1S<br />

Relay S<br />

I 2S<br />

Relay S<br />

I 0S<br />

Relay S<br />

mZ 1L<br />

mZ 2L<br />

mZ 0L<br />

V 2F<br />

+<br />

(1-m)Z 1L<br />

(1-m)Z 2L<br />

(1-m)Z 0L<br />

I 1R<br />

Relay R<br />

I 2R<br />

Relay R<br />

I 0R<br />

Relay R<br />

3RF<br />

Source<br />

R<br />

Z 1R<br />

Z 2R<br />

Z 0R<br />

I TOTAL<br />

Figure 6 Connection <strong>of</strong> Sequence Networks for a Single Line-to-Ground<br />

Fault at m<br />

At Relay S:<br />

At Relay R:<br />

( )<br />

V = –I • Z + m • Z<br />

(8)<br />

2F 2S 2S 2L<br />

( )<br />

V = − I • Z + (1− m) • Z<br />

(9)<br />

2F 2R 2R 2L<br />

Eliminate V 2F from Equations 8 and 9 and rearrange the<br />

resulting expression as follows:<br />

( Z2S<br />

+ m • Z2L<br />

)<br />

( 2R + ( − ) 2L )<br />

I2R<br />

= I 2S •<br />

Z 1 m • Z<br />

(10)<br />

To avoid alignment <strong>of</strong> Relay S and R data sets, take the<br />

magnitude <strong>of</strong> both sides <strong>of</strong> Equation 10 as follows:<br />

( Z2S<br />

+ m • Z2L<br />

)<br />

( 2R + ( − ) 2L )<br />

I2R<br />

= I 2S •<br />

Z 1 m • Z<br />

Equation 11 is then simplified to Equation 12 below.<br />

I<br />

2R<br />

=<br />

( I 2S • Z2S ) + m •( I 2S • Z2L<br />

)<br />

( Z + Z ) − m •( Z )<br />

2R 2L 2L<br />

(11)<br />

(12)<br />

To further simply Equation 12, define the following<br />

variables:<br />

I • Z = a + jb<br />

2R<br />

2S<br />

2S<br />

2S<br />

I • Z = c + jd<br />

2L<br />

Z + Z = e + jf<br />

2L<br />

Z = g + jh<br />

2L<br />

Substituting these variables into Equation 12 produces:<br />

I<br />

2R<br />

=<br />

( a + jb) + m •( c + jd)<br />

( e + jf ) − m •( g + jh)<br />

(13)<br />

Taking the square <strong>of</strong> both terms <strong>of</strong> Equation 13, expanding<br />

and rearranging terms produces a quadratic equation <strong>of</strong> the<br />

form:<br />

2<br />

A • m + B• m + C = 0<br />

(14)<br />

Equation 14 is solved for m using a quadratic solution. The<br />

coefficients <strong>of</strong> Equation 14 are given below.<br />

( ) ( )<br />

2 2 2 2 2<br />

2R<br />

A = I • g + h − c + d<br />

( ) ( )<br />

2<br />

2R<br />

2 2 2 2 2<br />

2R<br />

B = − 2 • I • e • g + f • h − 2 • a • c + b • d<br />

( ) ( )<br />

C = I • e + f − a + b<br />

(15)<br />

V. DISTRIBUTION SYSTEMS<br />

Fault location for distribution feeders uses the same basic<br />

principles as for transmission lines, but presents a great<br />

challenge for substation fault locators because <strong>of</strong> the diverse<br />

topology <strong>of</strong> the distribution system: laterals, spurs, and singlephase<br />

taps. On important feeders, some utilities model the line<br />

parameters to achieve a more precise fault location.<br />

One utility models a feeder using an Excel ® spreadsheet to<br />

show the line parameters. The spreadsheet includes node<br />

numbers, wire size, distance from the source, positive- and<br />

zero-sequence impedances, and fault currents. Figure 7 is an<br />

actual model <strong>of</strong> a feeder that is 5.47 miles long.<br />

Figure 7<br />

Spreadsheet Model (See Appendix for Enlarged View)<br />

A graphical representation <strong>of</strong> the feeder is shown in<br />

Figure 8.<br />

1<br />

Figure 8<br />

(distance in miles)<br />

2<br />

3<br />

4<br />

5<br />

Actual Fault<br />

(3 miles)<br />

Distribution Feeder Topology<br />

11<br />

6 7 8 9 10<br />

(1.41) (2.29)<br />

(3.29)<br />

12<br />

(3.06)<br />

21<br />

(5.39)<br />

13<br />

(3.27)<br />

14 15 16 17 18<br />

(3.66) (3.99)<br />

23<br />

(5.47)<br />

(2.56)<br />

20 19<br />

22<br />

(4.47)<br />

18 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


Figure 9 shows a screen capture <strong>of</strong> event report data from<br />

an actual fault.<br />

Figure 9<br />

Distribution Feeder Fault Event Screen Capture<br />

The event report indicates that a B-phase-to-ground fault<br />

occurred 3.02 miles from the station. From the feeder<br />

topology, there are two possible locations for the fault. As it<br />

turns out, line crews found a fast growing skinny tree growing<br />

close to the line, approximately three miles from the<br />

substation on the main line near Node 12.<br />

VI. USE OF FAULT INDICATORS<br />

Fault Indicators can be applied on lines to help locate<br />

faults. If a fault occurs beyond the location <strong>of</strong> the fault<br />

indicator, line crews observe an LED or flashing light,<br />

indicating that fault current was sensed. Reset can be done<br />

manually, electrostatically, or through a timer, depending on<br />

the design. Figure 10 is an example <strong>of</strong> how fault indicators<br />

can be placed to assist line crews in finding the fault location.<br />

VII. TRANSMISSION FAULT LOCATION EXAMPLES<br />

A. Example 1: 345 kV Automatic Spraying System<br />

Background: Repeated phase-phase faults had occurred on<br />

a 345 kV line. There was no inclement weather or lightning in<br />

the area. The number <strong>of</strong> faults caused voltage issues and a<br />

negative-sequence overcurrent element went into an alarm<br />

state at a regional nuclear plant. Nuclear plant personnel were<br />

concerned about the possibility <strong>of</strong> tripping the unit <strong>of</strong>f line.<br />

The line data for the transmission line:<br />

• Circuit 345-LINE is a 28.16 mile long, 345 kV line<br />

between terminals G and H.<br />

• 345-LINE—“We had dispatched linemen to the area<br />

based on fault location. The lineman was patrolling<br />

the line in the area when he observed an automatic<br />

spraying system operating very near the line. He went<br />

to the property owner’s home and learned that the<br />

automatic system runs along a track and sprays<br />

liquefied manure onto the open fields. The landowner<br />

checked the mechanism that controls the sprinkler and<br />

found that it had failed, causing the sprinkler to run<br />

under the line.”<br />

Figure 11 shows a one-line and event report screen<br />

captures.<br />

G<br />

C-A<br />

17.28 mi<br />

Reported<br />

17.8 mi from G<br />

10.36 mi from H<br />

Actual<br />

345-Line 28.16 mi<br />

C-A<br />

10.00 mi<br />

Reported<br />

H<br />

Relay<br />

Figure 10<br />

Example Location for Fault Indicators on Distribution Feeder<br />

Figure 11<br />

Example 1: One-Line and Event Report Screen Captures<br />

Impedance-Based Fault Location Experience | 19


As a way <strong>of</strong> confirming that the data was correct, we ran<br />

the two-ended negative-sequence impedance algorithm using<br />

the event reports from each end, as shown in Figure 12.<br />

A<br />

A-B-G<br />

-143.2 mi<br />

Reported<br />

6.8 mi from A<br />

Actual<br />

No event<br />

available<br />

B<br />

1.8<br />

115-Line 48.4 mi<br />

1.5<br />

m v<br />

ms v<br />

.633<br />

1<br />

0.5<br />

Calculated<br />

Two-Ended FL<br />

Actual FL<br />

0.1<br />

0 2 4 6 8<br />

0 v<br />

8<br />

RS<br />

Figure 12 Example 1: Mathcad Screen Capture—Actual Fault Location vs.<br />

Two-Ended Estimate<br />

The actual fault occurred where the sprinkler system was<br />

found, between 17.8 and 17.9 miles (m = .633) from<br />

Terminal G (based on patrol map tower locations).<br />

Conclusions for Example 1:<br />

1. The one-ended (17.28 and 10.00 miles, respectively)<br />

and two-ended (17.46 miles) fault locations yielded<br />

good results. All fault location estimates were within<br />

2%.<br />

2. Two-ended negative-sequence impedance-based<br />

method corroborates one-ended method.<br />

B. Example 2: Incorrect Fault Location Due to Incorrect<br />

Fault Type Identified<br />

Background: On this 115 kV Line, a relay tripped for an<br />

apparent fault. Targets indicated an A B-G fault that tripped<br />

on Zone 2. Upon analysis <strong>of</strong> the event report data, both ground<br />

directional overcurrent and ground distance elements tripped.<br />

All <strong>of</strong> the data indicated that the relay elements functioned<br />

properly. However, the relay produced a fault location<br />

estimate <strong>of</strong> –143 miles. This spawned an investigation to find<br />

the correct fault location.<br />

Figure 13 shows a one-line and event report screen<br />

captures.<br />

Figure 13<br />

Example 2: One-Line and Event Report Screen Captures<br />

By inspecting the event data directly, we saw a depressed<br />

C-phase to ground voltage and relatively low fault current<br />

(under 400 A primary) on all three phases. Based on this, we<br />

suspected that the fault was C-phase to ground with a weak<br />

source behind the relay and that the relay selected the wrong<br />

fault type.<br />

The relay used in this application is for three-pole trip<br />

applications. Its fault identification logic compares the angular<br />

relationship between I 0 and I 2 [5]. For this fault, I 2 leads I 0 by<br />

about 120 degrees (using A-phase as reference,), as shown in<br />

Figure 14. This indicates that the fault is either C-phase to<br />

ground, or A-B-ground.<br />

Figure 14<br />

Symmetrical Components from Fault—I2 leads I0 by 120 degrees<br />

20 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


The relay then performs “torque” calculations to determine<br />

which fault appears to be “closer” to the relay to decide<br />

between the two loops (C-G or A-B-G). For this case, the<br />

“torque” calculation showed the A-B-G as being closest. Thus,<br />

the relay selected the wrong loop.<br />

Improved relay designs (intended for single-pole or threepole<br />

applications) have better fault-type selection. These<br />

designs measure the fault resistance between phases and<br />

phase-to-ground. These would have correctly identified the<br />

fault type.<br />

Still, this example demonstrates the need for correct fault<br />

type selection. Knowing that the fault was C-G or A-B-G, we<br />

calculated the actual fault location to be 6.8 miles for the C-<br />

phase-to-ground fault (calculations from the event report data<br />

using the one-ended modified Takagi method):<br />

Calculated Fault Locations:<br />

C-G: 6.8 A-B-G: –137<br />

Conclusions for Example 2:<br />

1. Fault was C-phase-to-ground, one-ended location<br />

6.8 miles. (later confirmed from field reports).<br />

2. Weak source conditions challenge the one-ended<br />

fault locators for two reasons: nonhomogeneous<br />

system is more likely; fault type selection more<br />

difficult.<br />

3. Superior fault type selection would have provided<br />

the correct fault type and fault location.<br />

4. Two-ended negative-sequence impedance fault<br />

location would have correctly selected fault<br />

location for faults all along the line. (fault type<br />

selection not needed)<br />

C. Example 3: 345 kV Line Failed Insulator<br />

The line data for the transmission line:<br />

• Circuit 345-LINE is a 39.26 mile long, double-circuit<br />

345 kV line between terminals E and F.<br />

• 345-LINE circuit fault data:<br />

− 345-LINE E SUB <strong>SEL</strong>-311C 08-19-03.txt<br />

− 345-LINE F SUB <strong>SEL</strong>-311C 08-19-03.txt<br />

• 345-LINE – “A failed insulator strut was found on<br />

structure # 483, at about 6 miles from the F<br />

termination. Total line length is 39.26 miles.” Note –<br />

the line length setting in the relays is 39.30 miles.<br />

Figure 15 shows a one-line and event report screen<br />

captures.<br />

IA IB IC<br />

VA VB VC<br />

Figure 15<br />

5000<br />

2500<br />

0<br />

-2500<br />

-5000<br />

200<br />

100<br />

0<br />

-100<br />

-200<br />

E<br />

15000<br />

10000<br />

5000<br />

0<br />

-5000<br />

-10000<br />

-15000<br />

200<br />

IA IB IC<br />

VA VB VC<br />

100<br />

B-G<br />

28.55 mi<br />

Reported<br />

33.26 mi from E<br />

6.0 mi from F<br />

Actual<br />

345-Line 39.26 mi<br />

B-G<br />

6.29 mi<br />

Reported<br />

Example 3: One-Line and Event Report Screen Captures<br />

In Figure 16, the horizontal axis is the number <strong>of</strong> cycles<br />

and the vertical axis is the two-ended fault location averaged<br />

over several cycles. The actual fault location (33.26 miles) is<br />

about 0.85 per unit from the E terminal (depicted by a<br />

horizontal line on the graph). Note that the calculated fault<br />

locations are close to but do not exactly match the actual.<br />

1.5<br />

1.5<br />

m v<br />

ms v<br />

.85<br />

0<br />

-100<br />

-200<br />

1<br />

0.5<br />

IA IB IC VA VB VC<br />

2.5 5.0 7.5 10.0 12.5 15.0<br />

Cycles<br />

IA IB IC VA VB VC<br />

2.5 5.0 7.5 10.0 12.5 15.0<br />

Cycles<br />

Actual FL<br />

Calculated<br />

Two-Ended FL<br />

F<br />

0.0<br />

0<br />

0 2 4 6 8<br />

0 v<br />

RS<br />

8<br />

Figure 16 Example 3: Mathcad Screen Capture—Actual Fault Location<br />

versus Two-Ended Estimate<br />

Conclusions for Example 3:<br />

1. The system was slightly nonhomogeneous, which<br />

contributed to the poor local one-ended fault<br />

location estimate and the remote end being more<br />

accurate.<br />

Impedance-Based Fault Location Experience | 21


2. The fault was a fast clearing fault (approximately<br />

two cycles). The fault was interrupted just as the<br />

relay filtering (one-cycle cosine filter) had<br />

processed the data. As a result, fault location<br />

results are based on data less accurate than that <strong>of</strong> a<br />

fault present for a longer time window.<br />

3. The only event reports collected were 4-sample per<br />

cycle event reports. Thus, our analysis was limited<br />

because <strong>of</strong> the limited number <strong>of</strong> data points.<br />

When possible, it is better to collect event data<br />

with more data points (16 or more samples-percycle).<br />

4. Two-ended negative-sequence impedance fault<br />

location provided the best estimate, mainly because<br />

it mitigated any effects <strong>of</strong> the nonhomogeneous<br />

system and smoothed out the short data window by<br />

averaging the fault location estimates over several<br />

samples.<br />

D. Example 4 - 345 kV Line – Fire Under Line Conductors<br />

Background: There was a fire on a long 345 kV line in a<br />

wooded area. The fire caused several faults to occur, on<br />

different phases. Several reclose attempts were momentarily<br />

successful, until the still burning fire caused other phases to<br />

flash over creating another fault.<br />

345-LINE – “Line crews found a fire burning under a<br />

transmission line approximately 90.7 miles from the G<br />

substation. Total line length is 160.63 miles.”<br />

G<br />

A-G<br />

93.64 mi<br />

Reported<br />

90.7 mi from G<br />

69.93 mi from H<br />

Actual<br />

345-Line 160.63 mi<br />

A-G<br />

67.01 mi<br />

Reported<br />

H<br />

m v<br />

1.0<br />

ms v<br />

.564<br />

0.0<br />

1<br />

0.5<br />

Actual FL<br />

0<br />

0 2 4 6 8 10<br />

0 v<br />

Calculated<br />

Two-Ended FL<br />

Figure 18 Example 4: MathCad Screen Capture—Actual Fault Location vs<br />

Two-Ended Estimate<br />

Conclusions for Example 4:<br />

In Figure 18, the horizontal axis is the number <strong>of</strong> cycles<br />

and the vertical axis is the two-ended fault location averaged<br />

over several cycles. The actual fault location (90.7 miles) is<br />

0.564 per unit from the G terminal (depicted by a horizontal<br />

line on the graph). Even though the one-ended method<br />

produced good results, the two-ended negative-sequence<br />

impedance method is superior and allows operators to get<br />

much closer to the actual fault location.<br />

VIII. LAB TESTS TO OBTAIN TWO-ENDED FAULT<br />

LOCATION FROM ONE END<br />

The examples from the previous sections show how we can<br />

use the two-ended negative-sequence impedance method to<br />

get fault location data for operators. However, it takes time to<br />

collect the event reports and analyze the data. Is there any way<br />

we can get the two-ended data faster<br />

Many relays have the capability to send and receive the<br />

status <strong>of</strong> up to eight digital elements. In some newer designs,<br />

if less than 8 bits are used, we can use the unassigned bits to<br />

send additional information, such as remote time<br />

synchronization, virtual terminal sessions, and analog data.<br />

Via relay settings, we can send either measured or<br />

calculated analog quantities over a communication channel.<br />

Remembering that we need to exchange negative-sequence<br />

impedance and current information, we made an effort in the<br />

lab to calculate two-ended fault location from a single end<br />

with some promising results.<br />

Once analog values are sent, the remote relay receives the<br />

analog quantities. The received analog quantities can be used<br />

directly in logic or math equations and viewed using a<br />

s<strong>of</strong>tware command.<br />

The relay receiving the remote data then processes the<br />

multi-ended fault location algorithm. To do this, the relay uses<br />

internal mathematical capabilities, such as trigonometric<br />

functions, multiplication, division, addition, subtraction, and<br />

square root, to solve the quadratic equation for the fault<br />

location.<br />

RS<br />

10<br />

Figure 17<br />

Example 4: One-Line and Event Report Screen Captures<br />

22 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


IX. LAB TEST SETUP<br />

Several line faults with known fault locations were used as<br />

test cases. Local and remote relays exchange and use fault<br />

data for the purpose <strong>of</strong> implementing the two-ended negativesequence<br />

impedance fault location algorithm.<br />

We created a time-aligned COMTRADE file from the two<br />

line-end event reports for each fault. This fault simulation file<br />

is replayed into two relays to simulate the faults as seen by the<br />

relays in the field.<br />

The relays measure phase currents and voltages, and<br />

calculate 3I2 and ZS2. The magnitude <strong>of</strong> 3I2, and the<br />

magnitude and angle <strong>of</strong> Z2S from the remote line terminal is<br />

needed by the local relay to calculate a two-ended fault<br />

location. Therefore, we have to first save the 3I2 and Z2<br />

values at an appropriate time during the fault, and then<br />

communicate those values to the remote line terminal for the<br />

purpose <strong>of</strong> the fault location calculation.<br />

The data from the two ends <strong>of</strong> the line does not have to be<br />

time-aligned. However, we do need to select a point during the<br />

fault when values have settled to a steady state. We arbitrarily<br />

chose a point 1.5 cycles after fault detection for our data<br />

capture. At that point in time, the local relays will lock and<br />

hold the present fault value <strong>of</strong> 3I2 and Z2S. These values are<br />

sent to the remote line terminal. The remote line terminal will<br />

then perform a two-ended fault algorithm.<br />

Once physical test connections are verified and data scaled<br />

properly, the fault is played back to the relays. We compared<br />

the actual event report data to the played back data to verify<br />

accuracy, as shown in Figure 19.<br />

1_IA 1_IB 1_IC<br />

1_VA 1_VB 1_VC<br />

2500<br />

0<br />

-2500<br />

-5000<br />

50<br />

0<br />

-50<br />

-100<br />

10000<br />

1_IA 1_IB 1_IC 1_VA 1_VB 1_VC 2_IA 2_IB 2_IC 2_VA 2_VB 2_VC<br />

.066667 sec<br />

1_IA(A) 1_IB(A) 1_IC(A)<br />

1_VA(kV) 1_VB(kV) 1_VC(kV)<br />

2_IA(A) 2_IB(A) 2_IC(A)<br />

2_VA(kV) 2_VB(kV) 2_VC(kV)<br />

2500<br />

0<br />

-2500<br />

50<br />

0<br />

-50<br />

10000<br />

0<br />

-10000<br />

50<br />

0<br />

-50<br />

Figure 19<br />

Data<br />

1_IA(A) 1_IB(A) 1_IC(A) 1_VA(kV) 1_VB(kV) 1_VC(kV)<br />

2_IA(A) 2_IB(A) 2_IC(A) 2_VA(kV) 2_VB(kV) 2_VC(kV)<br />

.075 sec<br />

30.650 30.675 30.700 30.725 30.750 30.775 30.800 30.825 30.850<br />

Event Time (Sec) 15:13:30.717666<br />

Screen Capture <strong>of</strong> Original Event Data and COMTRADE Event<br />

X. TEST RESULTS AND ANALYSIS<br />

Table 2 shows the results from the laboratory tests.<br />

The One-Ended Estimate is the fault location taken directly<br />

from the relays.<br />

The Two-Ended Mathcad Point Estimate is a fault location<br />

based on the two-ended negative-sequence impedance<br />

method, where two-ended event report data is manually<br />

entered from one point in time (selected by the user).<br />

The Two-Ended Mathcad Average Estimate is based on the<br />

two-ended negative-sequence impedance method that<br />

averages the fault location over several samples. These results<br />

are produced using four-samples-per-cycle event reports.<br />

The Two-Ended Relay Estimate is the two-ended negativesequence<br />

impedance fault location automatically estimated by<br />

a relay using local and remote data captured 1.25 or 1.5 cycles<br />

after fault inception. (This estimate is based on the Two-<br />

Ended Mathcad Point Estimate Method.) A detailed listing <strong>of</strong><br />

the logic settings and calculated analog results are in the<br />

Appendix.<br />

2_IA 2_IB 2_IC<br />

2_VA 2_VB 2_VC<br />

0<br />

-10000<br />

100<br />

0<br />

-100<br />

2.50 2.55 2.60 2.65 2.70<br />

Event Time (Sec) 18:53:02.576166<br />

XI. CONCLUSIONS<br />

1. One-ended impedance-based fault location still<br />

produces very good results in most cases.<br />

2. If event data is available from both ends <strong>of</strong> the line,<br />

two-ended impedance fault location can improve fault<br />

location estimate.<br />

3. Off-line analytical tools are available to find the best<br />

fault location estimates.<br />

TABLE 2<br />

RESULTS FROM 345 KV LINE FAULT LOCATION LAB TESTS<br />

One-Ended Estimate<br />

Two-Ended Estimate<br />

Case Study Actual Location (Local) (Remote) Mathcad Point Mathcad Average Relay<br />

345-LINE Example 2 90.7 miles 93.64 miles 67.01 91.2 miles 90.7miles 91.25 miles<br />

Case Study Actual Location (Local) (Remote) Mathcad Point Mathcad Average Relay<br />

Impedance-Based Fault Location Experience | 23


4. Improve results by collecting events with the highest<br />

sampling rate and by using the average <strong>of</strong> several fault<br />

location estimate samples instead <strong>of</strong> a single point<br />

estimate.<br />

5. Short events present a challenge. More analysis is<br />

<strong>of</strong>ten required to get a more accurate fault estimate<br />

because <strong>of</strong> the short data window. The longer an event<br />

lasts, the better the fault location estimate.<br />

6. Technology is available to automatically calculate a<br />

two-ended fault location from a single end. Lab testing<br />

confirmed the viability <strong>of</strong> the technology. Testing<br />

indicates that accuracy is good on stable, longer<br />

lasting events.<br />

7. Developments are needed to make automatic<br />

collection <strong>of</strong> two-ended fault location applicable for<br />

all lines and faults.<br />

XIII. BIOGRAPHIES<br />

David Costello graduated from Texas A&M University in 1991 with a BSEE.<br />

He worked as a System Protection Engineer for Central and Southwest, and<br />

served on the System Protection Task Force for the ERCOT. In 1996, David<br />

joined Schweitzer Engineering Laboratories, where he has served as a Field<br />

Application Engineer and Regional Service Manager. He presently holds the<br />

title <strong>of</strong> Senior Application Engineer and works in Boerne, Texas. He is a<br />

member <strong>of</strong> IEEE, and the Planning Committee for the Conference for<br />

Protective Relay Engineers at Texas A&M University.<br />

Karl Zimmerman is a Senior Application Engineer with Schweitzer<br />

Engineering Labs in Belleville, Illinois. His work includes providing<br />

application support and technical training for protective relays. He is an active<br />

member <strong>of</strong> the IEEE <strong>Power</strong> System Relaying Committee and is the Chairman<br />

<strong>of</strong> Working Group D-2 on fault locating. Karl received his BSEE degree at the<br />

University <strong>of</strong> Illinois at Urbana-Champaign and has over 20 years <strong>of</strong><br />

experience in the area <strong>of</strong> system protection. He is a past speaker at many<br />

technical conferences and has authored several papers and application guides<br />

on protective relaying.<br />

XII. REFERENCES<br />

[1] IEEE Standard PC37.114, “Draft Guide For Determining Fault Location<br />

on AC Transmission and Distribution Lines,” 2004.<br />

[2] T. Takagi, Y. Yamakoshi, M. Yamaura, R. Kondou, and T. Matsushima,<br />

“Development <strong>of</strong> a New Type Fault Locator Using the One-Terminal<br />

Voltage and Current Data,” IEEE Transactions on <strong>Power</strong> Apparatus and<br />

Systems, Vol. PAS-101, No. 8, August 1982, pp. 2892-2898.<br />

[3] Edmund O. Schweitzer, III, “A Review <strong>of</strong> Impedance-Based Fault<br />

Locating Experience,” Proceedings <strong>of</strong> the 15th Annual Western<br />

Protective Relay Conference, Spokane, WA, October 24-27, 1988.<br />

[4] D.A. Tziouvaras, J.B. Roberts, G. Benmouyal, “New Multi-Ended Fault<br />

Location Design For Two- or Three-Terminal Lines,” presented at<br />

CIGRE Conference, 1999, http://www.selinc.com/techpprs/6089.pdf.<br />

24 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


APPENDIX<br />

The following are the relay logic settings to perform two-ended fault location from one end after receiving data from remote<br />

terminal. We based these settings on the two-ended negative-sequence impedance fault location method described in this paper.<br />

=>>SHO L<br />

Protection 1<br />

1: #<br />

2: # THESE SETTINGS ARE USED FOR TWO-ENDED FAULT LOCATION<br />

3: #<br />

4: # USE DEFINITE-TIME O/C DELAYS IN CONJUNCTION WITH<br />

5: # A COND TIMER TO MARK 1.5 CYCLES INTO FAULT<br />

6: #<br />

7: PCT10PU := 0.000000<br />

8: PCT10DO := 0.875000 #1.0 CYCLE DELAY INCLUDING PROCESSING TIME<br />

9: PCT10IN := R_TRIG 67P1T OR R_TRIG 67G1T<br />

10: #<br />

11: # USE A PROCESSING-INTERVAL WIDE PULSE TO MARK THE FAULT DATA<br />

12: #<br />

13: PSV02 := F_TRIG PCT10Q # 2 MSEC PULSE 1.5 CYCLES AFTER EVENT TRIGGER<br />

14: PSV03 := NOT PSV02 # THIS WILL BE ONE ALL TIMES EXCEPT DURING FAULT<br />

15: #<br />

16: # MEMORIZE FAULT DATA I2S MAG & ANG, Z2S MAG AND ANG, V2S MAG AND ANG<br />

17: # READ MEASURED VALUE TIMES PULSE BINARY ONE DURING FAULT, PLUS ZERO<br />

18: # NEXT TIME THRU, READ ZERO PLUS PREVIOUS STORED VALUE TIMES ONE<br />

19: #<br />

20: PMV01 := L3I2FIM * PSV02 + PMV01 * PSV03<br />

21: # PMV01 STORES THE LINE 3I2 MAGNITUDE AT 1.5 CYCLES AFTER EVENT TRIGGER<br />

22: # UNTIL A NEW EVENT OCCURS<br />

23: #<br />

24: PMV02 := L3I2FIA * PSV02 + PMV02 * PSV03<br />

25: # PMV02 STORES THE LINE 3I2 ANGLE AT 1.5 CYCLES AFTER EVENT TRIGGER<br />

26: # UNTIL A NEW EVENT OCCURS<br />

27: #<br />

28: PMV03 := 3V2FIM * PSV02 + PMV03 * PSV03<br />

29: # PMV03 STORES THE LOCAL 3V2 MAGNITUDE AT 1.5 CYCLES AFTER EVENT<br />

30: # TRIGGER UNTIL A NEW EVENT OCCURS<br />

31: #<br />

32: PMV04 := 3V2FIA * PSV02 + PMV04 * PSV03<br />

33: # PMV04 STORES THE LOCAL 3V2 ANGLE AT 1.5 CYCLES AFTER EVENT<br />

34: # TRIGGER UNTIL A NEW EVENT OCCURS<br />

35: #<br />

36: PMV05 := (3V2FIM / (L3I2FIM + 0.001000)) * PSV02 + PMV05 * PSV03<br />

37: # PMV05 STORES THE NEG SEQ SOURCE IMPEDANCE MAGNITUDE AT 1.5 CYCLES<br />

38: # AFTER EVENT TRIGGER UNTIL A NEW EVENT OCCURS<br />

39: #<br />

40: PMV06 := (3V2FIA - L3I2FIA) * PSV02 + PMV06 * PSV03<br />

41: # PMV06 STORES THE NEG SEQ SOURCE IMPEDANCE ANGLE AT 1.5 CYCLES AFTER<br />

42: # EVENT TRIGGER UNTIL A NEW EVENT OCCURS<br />

43: #<br />

44: # ANALOG MIRRORED BIT VALUES ARE 16-BIT SIGNED INTEGERS<br />

45: # SO WE MUST SCALE APPROPRIATELY BEFORE SENDING TO RETAIN ACCURACY<br />

46: #<br />

47: PMV07 := PMV01 * 100.000000 # SCALE 3I2 MAG BY MULTIPLYING BY 100<br />

48: PMV08 := PMV05 * 100.000000 # SCALE Z2S MAG BY MULTIPLYING BY 100<br />

49: PMV09 := PMV06 * 10.000000 # SCALE Z2S ANGLE BY MULTIPLYING BY 10<br />

50: #<br />

51: # SEND PMV07, PMV08, AND PMV09 AS MIRRORED BIT ANALOGS<br />

52: # AND REMEMBER TO DIVIDE BY SCALING VALUE AT OTHER END<br />

53: #<br />

54: PMV10 := 0.970000 # ENTER Z1MAG FROM RELAY SETTINGS HERE<br />

55: PMV11 := 79.000000 # ENTER Z1ANG FROM RELAY SETTINGS HERE<br />

56: PMV12 := 2.170000 # ENTER LINE LENGTH FROM RELAY SETTINGS HERE<br />

57: #<br />

58: # SCALE RECEIVED MIRRORED BIT ANALOG VALUES FROM REMOTE RELAY<br />

59: #<br />

60: PMV13 := MB1A / 100.000000 # REMOTE 3I2 MAG RECEIVED THRU MB A, SCALED<br />

61: PMV14 := MB2A / 100.000000 # REMOTE Z2S MAG RECEIVED THRU MB A, SCALED<br />

62: PMV15 := MB3A / 10.000000 # REMOTE Z2S ANGLE RECEIVED THRU MB A, SCALED<br />

63: #<br />

64: # CORRECTION MADE - CONVERT TO PRIMARY VALUES<br />

65: #<br />

66: # NEXT WE SOLVE THE QUADRATIC EQUATION AND DETERMINE FAULT LOCATION<br />

67: # REFER TO ROBERTS, TZIOUVARAS, BENMOUYAL "NEW MULTI-ENDED FAULT LOC"<br />

68: # REFER TO MOXLEY, WOODWARD "IMPROVE SUBSTATION CONTROL AND PROTECTION"<br />

69: # REFER TO ZIMMERMAN "TWO-ENDED FAULT LOCATION" MATHCAD FILE<br />

70: #<br />

71: PMV16 := (PMV03 * 600.000000 / 3.000000) * COS(PMV04) # A'<br />

72: PMV16 := (PMV03 * 600.000000 / 3.000000) * COS(PMV04) # A'<br />

73: PMV17 := (PMV03 * 600.000000 / 3.000000) * SIN(PMV04) # B'<br />

74: #<br />

75: PMV18 := (PMV01 * 400.000000 / 3.000000) * (PMV10 * 600.000000 / \<br />

400.000000)<br />

Impedance-Based Fault Location Experience | 25


76: PMV18 := PMV18 * COS(PMV02 + PMV11) # C'<br />

77: #<br />

78: PMV19 := (PMV01 * 400.000000 / 3.000000) * (PMV10 * 600.000000 / \<br />

400.000000)<br />

79: PMV19 := PMV19 * SIN(PMV02 + PMV11) # D'<br />

80: #<br />

81: PMV20 := (PMV14 * 600.000000 / 400.000000) * COS(PMV15)<br />

82: PMV20 := PMV20 + (PMV10 * 600.000000 / 400.000000) * COS(PMV11) # E'<br />

83: #<br />

84: PMV21 := (PMV14 * 600.000000 / 400.000000) * SIN(PMV15)<br />

85: PMV21 := PMV21 + (PMV10 * 600.000000 / 400.000000) * SIN(PMV11) # F'<br />

86: #<br />

87: PMV22 := (PMV10 * 600.000000 / 400.000000) * COS(PMV11) # G'<br />

88: #<br />

89: PMV23 := (PMV10 * 600.000000 / 400.000000) * SIN(PMV11) # H'<br />

90: #<br />

91: PMV24 := (PMV13 * 400.000000 / 3.000000) * (PMV13 * 400.000000 / \<br />

3.000000) * (PMV22 * PMV22 + PMV23 * PMV23)<br />

92: PMV24 := PMV24 - (PMV18 * PMV18 + PMV19 * PMV19)<br />

93: # PREVIOUS LINE IS A<br />

94: #<br />

95: PMV25 := -2.000000 * (PMV13 * 400.000000 / 3.000000) * (PMV13 * \<br />

400.000000 / 3.000000)<br />

96: PMV25 := PMV25 * (PMV20 * PMV22 + PMV21 * PMV23)<br />

97: PMV37 := PMV25 - 2.000000 * (PMV16 * PMV18 + PMV17 * PMV19)<br />

98: # PREVIOUS LINE IS B - NOTE EQUATION STARTS WITH "-2"<br />

99: # CORRECTING A TYPOGRAPHICAL ERROR IN REFERENCES ABOVE<br />

100: #<br />

101: PMV26 := (PMV13 * 400.000000 / 3.000000) * (PMV13 * 400.000000 / \<br />

3.000000) * (PMV20 * PMV20 + PMV21 * PMV21)<br />

102: PMV26 := PMV26 - (PMV16 * PMV16 + PMV17 * PMV17)<br />

103: # PREVIOUS LINE IS C<br />

104: #<br />

105: PMV27 := ABS(PMV37 * PMV37 - 4.000000 * PMV24 * PMV26)<br />

106: PMV27 := SQRT(PMV27)<br />

107: PMV27 := PMV27 - PMV37<br />

108: PMV28 := PMV27 * PMV12 / (2.000000 * (PMV24 + 0.001000))<br />

109: # PREVIOUS LINE IS M1 FAULT LOC ESTIMATE<br />

110: #<br />

111: PMV29 := -PMV37<br />

112: PMV30 := (PMV37 * PMV37 - 4.000000 * PMV24 * PMV26)<br />

113: PMV30 := SQRT(PMV30)<br />

114: PMV31 := -PMV30<br />

115: PMV32 := PMV29 + PMV31<br />

116: PMV33 := PMV32 * PMV12 / (2.000000 * (PMV24 + 0.001000))<br />

117: # PREVIOUS LINE IS M2 FAULT LOC ESTIMATE<br />

118: #<br />

119: # ONE ESTIMATE WILL BE "REASONABLE", WITHIN THE LINE, TRASH OTHER<br />

120: PMV34 := ABS(PMV28)<br />

121: PSV04 := PMV34 > PMV12<br />

122: PSV05 := NOT PSV04 # THIS WILL BE A LOGICAL ONE IF M1 IS OK<br />

123: PMV35 := ABS(PMV33)<br />

124: PSV06 := PMV35 > PMV12<br />

125: PSV07 := NOT PSV06 # THIS WILL BE A LOGICAL ONE IF M2 IS OK<br />

126: #<br />

127: PMV36 := PSV05 * PMV34 + PSV07 * PMV35 # THIS IS THE FAULT LOC ESTIMATE<br />

128: PMV64 := PMV36 # MOVE FAULT LOC TO AREA VISIBLE TO "METER PMV"<br />

129: #<br />

26 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


The following shows the analog math results from the relay logic. PMV64 is the two-ended fault location calculated from the<br />

relay.<br />

=>>MET PMV A<br />

4XX RELAY Date: 09/10/2004 Time: 17:49:07.559<br />

SUB C Serial Number: 2004104018<br />

Protection Analog Quantities<br />

PMV01 = 9.051<br />

PMV02 = -64.952<br />

PMV03 = 14.069<br />

PMV04 = -175.743<br />

PMV05 = 1.554<br />

PMV06 = -110.791<br />

PMV07 = 905.150<br />

PMV08 = 155.413<br />

PMV09 = -1107.905<br />

PMV10 = 3.960<br />

PMV11 = 84.000<br />

PMV12 = 160.630<br />

PMV13 = 11.590<br />

PMV14 = 1.080<br />

PMV15 = 255.700<br />

PMV16 = -14029.924<br />

PMV17 = -1044.396<br />

PMV18 = 33881.355<br />

PMV19 = 11697.921<br />

PMV20 = 3.679<br />

PMV21 = 72.294<br />

PMV22 = 10.348<br />

PMV23 = 98.458<br />

PMV24 = 8.217E+08<br />

PMV25 = -3.076E+09<br />

PMV26 = 9.283E+08<br />

PMV27 = 3.268E+09<br />

PMV28 = 319.442<br />

PMV29 = 2.101E+09<br />

PMV30 = 1.167E+09<br />

PMV31 = -1.167E+09<br />

PMV32 = 9.336E+08<br />

PMV33 = 91.249<br />

PMV34 = 319.442<br />

PMV35 = 91.249<br />

PMV36 = 91.249<br />

PMV37 = -2.101E+09<br />

PMV38 = 0.000<br />

PMV39 = 0.000<br />

PMV40 = 3000.000<br />

PMV41 = 120.000<br />

PMV42 = 0.000<br />

PMV43 = 0.000<br />

PMV44 = 0.000<br />

PMV45 = 0.000<br />

PMV46 = 0.000<br />

PMV47 = 0.000<br />

PMV48 = 0.000<br />

PMV49 = 0.000<br />

PMV50 = 0.000<br />

PMV51 = 0.000<br />

PMV52 = 0.000<br />

PMV53 = 0.000<br />

PMV54 = 0.000<br />

PMV55 = 0.000<br />

PMV56 = 0.000<br />

PMV57 = 0.000<br />

PMV58 = 0.000<br />

PMV59 = 0.000<br />

PMV60 = 0.000<br />

PMV61 = 0.000<br />

PMV62 = 0.000<br />

PMV63 = 0.000<br />

PMV64 = 91.249<br />

Impedance-Based Fault Location Experience | 27


28 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong><br />

Copyright © <strong>SEL</strong> 2004<br />

(All rights reserved)<br />

20041004 TP6180


Digital Communications for <strong>Power</strong> System<br />

Protection: Security, Availability, and Speed<br />

Edmund O. Schweitzer III, Ken Behrendt, and Tony Lee, Schweitzer Engineering Laboratories, Inc.<br />

I. INTRODUCTION<br />

New channels and digital techniques in communications<br />

provide opportunities to advance the speed, security,<br />

dependability, and sensitivity <strong>of</strong> protection—while<br />

simultaneously reducing the costs associated with using<br />

communications. Lower communication costs mean more<br />

opportunities to benefit from pilot protection. The net result is<br />

a higher quality <strong>of</strong> electric power delivered for each dollar<br />

invested.<br />

A classical pilot communication scheme is shown in<br />

Figure 1, and a direct digital-to-digital scheme is shown in<br />

Figure 2. Clearly, the direct digital system is simpler. In this<br />

paper, we will show that direct digital communications<br />

economically provide several bits in each direction—and these<br />

extra bits lead to simpler, more sensitive, and more flexible<br />

schemes.<br />

Figure 1<br />

Figure 2<br />

Frequency-Shift Audio Keying Over Voice Channel<br />

Direct Digital Signaling Over Asynchronous Channel<br />

II. WHAT IS THE CAPACITY OF A CHANNEL<br />

TO COMMUNICATE<br />

How much information can be sent, theoretically and<br />

practically, through a given channel while still maintaining<br />

acceptable reliability<br />

In 1948, Claude E. Shannon [1] mathematically formalized<br />

a theoretical limit. His theory was that information can be<br />

reliably transmitted over a noisy channel if the data<br />

transmission rate is sufficiently low. If the relative noise<br />

increases, the maximum reliable transmission rate decreases.<br />

For instance, a channel with bandwidth W and received noise<br />

power N, can transmit information at rate C with arbitrarily<br />

high dependability, as long as the average signal power P<br />

satisfies:<br />

⎛ P ⎞<br />

C = W log2<br />

⎜ + 1⎟<br />

⎝ N ⎠<br />

The ratio P/N is the signal to noise ratio (SNR). The base<br />

<strong>of</strong> the logarithm depends on how we measure C. When C is<br />

measured in bits per second, the logarithm is base two. In<br />

general, the logarithm base is the number <strong>of</strong> symbols in the<br />

alphabet to be transmitted.<br />

Quieter channels lend themselves to faster data<br />

transmission. Conversely, faster data transmission requires a<br />

quieter channel.<br />

If SNR = P/N = 1, then the Shannon limit is C = W<br />

(bits/second), i.e., the channel capacity for reliable<br />

transmission is the channel bandwidth.<br />

If the SNR = 20dB = 100, then C = 6.7 W (bits/second).<br />

We can see that increasing the SNR gives us an opportunity to<br />

reliably transmit more data, faster.<br />

For example, contrast the channel requirements for two<br />

different transmission rates. A frequency shift keyed (FSK)<br />

audio tone transmission over a voice channel might carry a<br />

single bit <strong>of</strong> information, such as a permissive trip signal.<br />

Suppose we require transmission to occur reliably in 20 ms, so<br />

the required data transmission rate is 1 bit / 20 ms =<br />

50 bits/second. Also assume the receiver filter has a<br />

bandwidth <strong>of</strong> 300 Hz. According to Shannon, the received<br />

signal, after filtering, must have an SNR <strong>of</strong> greater than about<br />

0.1. Before filtering, the SNR on a 3 kHz channel must be<br />

greater than about 0.01, assuming white noise.<br />

A 9600 bits/second data stream transmitted over a 3 kHz<br />

channel requires an SNR <strong>of</strong> at least about 10, according to<br />

Shannon's work. Therefore, the voice grade channel must be<br />

about 10/0.01 or 1000 times quieter (assuming the same<br />

transmit power and modulation techniques) to reliably transfer<br />

data at 9600 baud.<br />

In practice, it is difficult to approach Shannon's limits. The<br />

two examples cited above, when implemented using present<br />

technology, both need about ten times better SNR than<br />

Shannon’s limit.<br />

Both Shannon’s prediction and practical experience show<br />

that when we have a better channel, we can send more<br />

information per unit bandwidth. The quality <strong>of</strong> many digital<br />

channels is excellent, and opens the door to new digital<br />

techniques in protection.<br />

The communications engineer uses encoding and<br />

modulation techniques to approach Shannon’s limit, and to<br />

balance data rate with reliability within the context <strong>of</strong> a given<br />

application.<br />

Digital Communications for <strong>Power</strong> System Protection: Security, Availability, and Speed | 29


III. ENCODING EIGHT BITS FOR DIGITAL TRANSMISSION<br />

We foresee a great future for sharing a handful <strong>of</strong> bits<br />

directly from one relay to another, over an array <strong>of</strong> digital<br />

channels <strong>of</strong> moderate capacity. Pilot protection, control,<br />

adaptive relaying, monitoring, and breaker-failure are some<br />

examples.<br />

Our starting point was eight bits <strong>of</strong> information in a<br />

message with enough redundancy to meet protection-security<br />

requirements, yet efficient enough to be useful at data rates<br />

from several kilobaud and up.<br />

A. Security<br />

Communications are secure when the receiving end<br />

reliably detects whether the received information differs from<br />

the transmitted information.<br />

The standard IEC 834-1 [2] contains recommendations for<br />

blocking, permissive, and direct tripping pilot schemes, in<br />

terms <strong>of</strong> their susceptibility to noise bursts. The short table<br />

below gives the expected minimum number <strong>of</strong> noise bursts<br />

required to produce an undesirable output.<br />

Scheme Type<br />

Security<br />

(bursts/undetected error)<br />

Blocking 10 4<br />

Permissive Tripping 10 7<br />

Direct Tripping 10 8<br />

To help detect noise bursts, we can add some redundant<br />

information to the transmitted message. Shannon gives a<br />

formal definition <strong>of</strong> redundancy:<br />

Redundancy = Total Bits Transmitted – Information<br />

For example, if we transmit a total <strong>of</strong> 10 bits, and there are<br />

eight bits <strong>of</strong> information, then the redundancy is 10 – 8 =<br />

2 bits.<br />

Redundancy is necessary, but not sufficient for security.<br />

One <strong>of</strong> the objectives <strong>of</strong> encoding is to make each <strong>of</strong> the<br />

distinct messages (e.g., for eight bits there are 256 different<br />

messages) as different as possible from the rest. The<br />

quantitative measure <strong>of</strong> this difference is the Hamming<br />

distance. It is defined as the minimum number <strong>of</strong> bits that<br />

could be corrupted in one distinct message, which would<br />

result in a different distinct valid message.<br />

The simplest form <strong>of</strong> redundancy that increases Hamming<br />

distance is repetition. Consider a single bit <strong>of</strong> information<br />

(permissive trip for instance) that must be received by the<br />

remote relay from the local relay. If the local relay transmits<br />

only that bit, the remote relay cannot detect whether the bit<br />

has been corrupted. The remote relay receives a one or a zero,<br />

and has no indication if the received value is the same as the<br />

transmitted value.<br />

Now assume that the local relay transmits the bit <strong>of</strong><br />

information twice. The receiving relay compares the two bits.<br />

If they are the same, the receiving relay assumes they are<br />

correct and accepts the bit <strong>of</strong> information. But, if they differ,<br />

the receiving relay discards the information.<br />

When we inject noise bursts onto the channel, we generate<br />

messages almost at random, because the bit error rate is so<br />

high. There are two valid messages (00 and 11) and two<br />

invalid messages (10 and 01). Therefore, we expect that a<br />

randomly generated message will be accepted by the receiving<br />

relay half the time, or once per two noise bursts. Since that is a<br />

long way from 10 7 , add more redundancy.<br />

If the transmitting relay adds another bit <strong>of</strong> redundancy,<br />

then there are still only two valid messages (000 and 111), but<br />

there are now six invalid messages (001, 010, 011, 100 101,<br />

and 110). The receiving relay will accept a randomly<br />

generated message two out <strong>of</strong> eight times, or one out <strong>of</strong> four,<br />

on average.<br />

Every time we add a bit <strong>of</strong> redundancy, we cut the<br />

probability <strong>of</strong> accepting a randomly generated message by<br />

half. Thus, the expected number <strong>of</strong> randomly generated<br />

messages is 2 n per undetected error, where n is the number <strong>of</strong><br />

redundant bits.<br />

We need log 2 (10 7 ), or about 23.3, bits <strong>of</strong> redundancy to get<br />

10 7 security. Our 1 bit <strong>of</strong> information, plus 24 redundant bits,<br />

yields a 25-bit message.<br />

This method does not make very good use <strong>of</strong> our channel<br />

however, because we must transmit 25 bits to securely<br />

communicate a single bit <strong>of</strong> information. Now that we have<br />

the required security, we can add more information bits with<br />

no loss <strong>of</strong> redundancy.<br />

Suppose we decide to repeat eight bits <strong>of</strong> information three<br />

times, and add some channel framing bits for 36 bits total.<br />

Four <strong>of</strong> these channel framing bits do not count as redundant<br />

bits, so there are 36 – 4 – 8 = 24 redundant bits. We have<br />

increased our information transmission capability by a factor<br />

<strong>of</strong> eight over the original single bit, decreased the rate <strong>of</strong><br />

transmission by 1 – 36/25 = 44%, and maintained 2 24 = 1.7 x<br />

10 7 security to randomly generated messages.<br />

B. Dependability<br />

Suppose we want to transmit eight data bits with 10 7<br />

security as described above. We have shown that the message<br />

must consist <strong>of</strong> at least the eight data bits plus 24 redundant<br />

bits, plus some non-redundant channel framing bits (36 bits<br />

total).<br />

Assume that channel errors occur at some average random<br />

bit error rate (BER). The receiver rejects an entire message <strong>of</strong><br />

36 bits even if just one bit is in error. In addition, after the<br />

receiver detects a bit error, it will probably require that two or<br />

more consecutive messages are received without error before<br />

using the received information. The probability that a message<br />

will not be accepted is about 2 • k • BER, where k is the<br />

number <strong>of</strong> bits in the message. This approximation holds for<br />

2 • k • BER < 0.1. Therefore, we expect the average<br />

unavailability to be about 72 times the channel bit error rate.<br />

By increasing the redundancy from one to 24 bits we have<br />

increased security from 2 to 10 7 . Simultaneously, we increased<br />

unavailability by a factor <strong>of</strong> 72. If we started with a channel<br />

with BER <strong>of</strong> 10 -6 the unavailability is now 72 x 10 -6 . This<br />

demonstrates a trade-<strong>of</strong>f between security and dependability:<br />

30 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


increasing security by a factor <strong>of</strong> 10 6 decreased dependability<br />

by a factor <strong>of</strong> 72.<br />

C. Speed<br />

Again consider the 36-bit message developed earlier.<br />

Compared to the two-bit message with a single redundant bit,<br />

it takes 18 times as long to transmit/receive. At<br />

9600 bits/second, it takes about 3.6 ms to transmit/receive the<br />

36-bit message, compared to 0.2 ms for the 2 bit message.<br />

Therefore, increasing security decreases speed somewhat.<br />

D. Adaptability<br />

The 36-bit message developed above gives 10 7 security,<br />

and, when coupled with the proper digital channel, still yields<br />

high speed and excellent dependability. Suppose we use one<br />

<strong>of</strong> the eight data bits for a block-trip signal, and another <strong>of</strong> the<br />

bits for a remote-control direct-trip signal. The security<br />

afforded by the 36-bit protocol is sufficient for blocking<br />

schemes. However, IEC 834-1 recommends ten times better<br />

security for direct tripping. We want to increase the security <strong>of</strong><br />

the direct trip signal without affecting the speed or availability<br />

<strong>of</strong> the blocking signal.<br />

This is easily accomplished with a pickup security counter<br />

on the direct trip bit. For example, a count <strong>of</strong> two requires<br />

reception <strong>of</strong> two successive 36-bit messages with the direct<br />

trip bit set before updating the direct trip bit in the receiving<br />

relay. If we return to the test prescribed by IEC 834-1, we<br />

would still expect to inject 10,000,000 bursts <strong>of</strong> noise, on<br />

average, to get one corrupted message that is incorrectly<br />

accepted by the receiving relay. However, to perturb the<br />

direct-trip bit qualified by a two-count security counter, the<br />

very next message must also be acceptable, and must have the<br />

same direct trip value. This also happens about one in<br />

10,000,000 times. So the probability <strong>of</strong> a false trip in response<br />

to noise bursts is about (10 7 ) 2 or 10 14 . This is six orders <strong>of</strong><br />

magnitude more secure than the IEC 834-1 recommendation<br />

for direct tripping.<br />

Remember we must trade <strong>of</strong>f speed and/or availability to<br />

gain security. Here, the direct trip signal is delayed by one<br />

additional message-time, and the unavailability is roughly<br />

doubled.<br />

E. Practical Implementation<br />

We implemented the 36-bit code described above. The 36-<br />

bit message is transmitted at 19,200 kbits/s in 36/19,200 =<br />

1.875 ms. Allowing for 2 ms <strong>of</strong> latency, plus 2 ms for<br />

processing time in the receiving relay, gives a total <strong>of</strong> about<br />

6 ms from the time the transmitting relay makes a decision to<br />

when the receiving relay has makes a decision influenced by<br />

the transmitting relay.<br />

The delays for a tone set between two relays are much<br />

longer:<br />

2 ms output contact + 12 ms tone set +<br />

2 ms latency + 2 ms processing = 18 ms.<br />

Thus, the direct digital communications gives us eight<br />

times the data with one-third the delay, at far less cost and<br />

complexity.<br />

To test the protocol security, we injected 200 ms long<br />

white noise bursts onto a direct copper connection between<br />

relays. We set the transmitting relay to transmit a known set <strong>of</strong><br />

eight bits, and we set the receiving relay to trigger an event<br />

report upon reception <strong>of</strong> anything but that known pattern.<br />

The receiving relay triggered the first event report after<br />

7 million noise bursts. We terminated the test after 20 million<br />

noise bursts (and nearly 50 days) with still only one<br />

undetected error.<br />

F. Applicability<br />

Since the protocol described above is a simple serial bit<br />

stream, it is compatible with many channels and many types<br />

<strong>of</strong> data communications equipment.<br />

G. Channel Performance Monitoring<br />

Digital communications provide opportunities for<br />

performance monitoring, so the quality can be assessed, and<br />

problems can be quickly detected and remedied.<br />

One channel monitor tallies the time the received data are<br />

corrupted or absent, and normalizes that time to the total<br />

elapsed time. This directly measures the unavailability <strong>of</strong> the<br />

communications. Although unavailability is a useful long-term<br />

measurement, it hides long but infrequent channel<br />

disturbances. For example, suppose a channel monitor is set to<br />

alarm when the unavailability exceeds 500 x 10 -6 . If that<br />

channel is error-free for one year and then the channel is<br />

completely lost, the unavailability monitor will not alarm until<br />

four hours later.<br />

A second monitor can be used to alarm when the channel is<br />

not available for a certain continuous time, say one second.<br />

The unavailability alarm responds to gradual degradations<br />

in bit-error rate. The duration alarm responds more quickly to<br />

outright channel failures.<br />

A sample report from such a monitor follows. It reports the<br />

256 most-recent errors, the average unavailability for the time<br />

<strong>of</strong> the report, and the longest-duration channel outage.<br />

Digital Communications for <strong>Power</strong> System Protection: Security, Availability, and Speed | 31


Summary for Channel A<br />

For 06/19/98 15:43:48.887 to 07/30/98 10:13:11.925<br />

Total failures 14 Last error Re-sync<br />

Relay disabled 1<br />

Data error 4 Longest failure 0 00:00:41.352<br />

Re-sync 4<br />

Underrun 1 Unavailability 0.000015<br />

Overrun 0<br />

Parity error 3<br />

Framing error 1<br />

START START END END<br />

# DATE TIME DATE TIME DURATION EVENT<br />

1 07/10/98 11:19:14.769 07/10/98 11:19:24.419 00:00:09.650 Re-sync<br />

2 07/09/98 11:48:13.572 07/09/98 11:48:14.126 00:00:00.554 Underrun<br />

3 07/09/98 11:48:12.710 07/09/98 11:48:13.481 00:00:00.770 Re-sync<br />

4 07/09/98 10:38:32.062 07/09/98 10:39:13.414 00:00:41.352 Parity error<br />

5 07/07/98 09:33:35.389 07/07/98 09:33:35.419 00:00:00.029 Re-sync<br />

6 07/07/98 09:21:44.183 07/07/98 09:21:44.229 00:00:00.045 Parity error<br />

7 07/07/98 09:21:44.087 07/07/98 09:21:44.154 00:00:00.066 Data error<br />

8 07/07/98 09:21:36.077 07/07/98 09:21:36.127 00:00:00.049 Data error<br />

9 07/07/98 09:21:33.727 07/07/98 09:21:33.777 00:00:00.050 Data error<br />

10 06/29/98 09:19:12.075 06/29/98 09:19:12.120 00:00:00.044 Framing error<br />

11 06/26/98 15:04:28.653 06/26/98 15:04:28.701 00:00:00.047 Data error<br />

12 06/26/98 15:01:40.209 06/26/98 15:01:40.243 00:00:00.033 Re-sync<br />

13 06/26/98 15:00:27.803 06/26/98 15:00:27.845 00:00:00.041 Parity error<br />

14 06/19/98 15:43:48.887 06/19/98 15:43:48.887 00:00:00.000 Relay disabled<br />

=><br />

This specific channel is a digital leased line running at<br />

56 kbaud. The relay detected 14 total errors that resulted in an<br />

average unavailability <strong>of</strong> 15 x 10 -6 . The longest channel<br />

outage was 41.352 seconds.<br />

Notice that the errors are grouped in clumps. The report<br />

was cleared on 6/19, and the circuit experienced no errors<br />

until 6/26. On 6/26 there were three errors in four minutes.<br />

The circuit was then perfect for three days, until a single error<br />

occurred on 6/29. There were no more errors for over a week,<br />

then there were four errors in eleven seconds, followed by<br />

another error twelve minutes later. After two days without<br />

error, there was a span <strong>of</strong> 41 seconds where the relay did not<br />

receive an acceptable message. This underscores the<br />

additional value <strong>of</strong> a continuous outage monitor, because the<br />

unavailability monitor would not have alarmed for this outage<br />

unless it was set as low as 30 x 10 -6 . On the next day another<br />

extended outage <strong>of</strong> over nine seconds occurred. The channel<br />

was error-free for the next 20 days, from 7/10 to 7/30, when<br />

the report was downloaded from the relay.<br />

Later, we will discuss error-seconds per day as a measure<br />

<strong>of</strong> quality <strong>of</strong> the leased lines. Performance monitoring<br />

provides the quantitative feedback needed to maintain and<br />

improve the quality <strong>of</strong> communications. Relay event reporting<br />

provides an additional perspective on the performance <strong>of</strong><br />

communications for every event, because the communicated<br />

bits are reported as additional I/O points. For example,<br />

communications disruptions during faults are easily observed,<br />

should they occur.<br />

IV. CHANNELS<br />

This section compares some communications channels that<br />

might be used in pilot and control schemes.<br />

A. Dedicated Fiber<br />

Perhaps the ultimate digital channel in terms <strong>of</strong><br />

dependability, security, speed, and simplicity is dedicated<br />

fiber optics. Low-cost fiber-optic modems make dedicated<br />

fiber channels even more attractive. Often, the modems can be<br />

powered by the relay, eliminating the cost and loss <strong>of</strong><br />

availability involved in separate power sources. Some<br />

modems also plug directly onto the digital relay [3], which<br />

eliminates a metallic cable. Eliminating the cable and the<br />

external power source removes “antennas” for possible EMI<br />

susceptibility.<br />

When the communications path is short, the cost <strong>of</strong> the<br />

fiber is not very significant. On longer paths, multiplexers<br />

may be considered, to increase the amount <strong>of</strong> data<br />

communicated over a fiber-pair. However, the relatively small<br />

incremental cost <strong>of</strong> adding and using one fiber-pair for<br />

protection alone is probably justified by the increase in<br />

simplicity and availability that a dedicated fiber scheme<br />

<strong>of</strong>fers. Furthermore, the small incremental cost is partially<br />

<strong>of</strong>fset by the very low cost <strong>of</strong> the simple direct-connected<br />

fiber-optic modems.<br />

Bit errors are extremely rare on most fiber-optic links. If<br />

the link is available, then it is near-perfect, because the fiber<br />

medium is unaffected by the RFI, EMI, ground-potential rise,<br />

weather, and so on.<br />

32 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


The receiver amplifier is the major source <strong>of</strong> noise in the<br />

system—and that noise source is very small compared to the<br />

large signals used in simple, practical systems. The received<br />

signal strength is the transmit power minus attenuation.<br />

Attenuation in decibels is proportional to fiber length plus<br />

some loss for each connection or splice. A system designer<br />

usually fixes the transmit power at some level, specifies a<br />

power margin, and then guarantees some allowable fiber<br />

length at some maximum bit-error-rate or BER. If we use less<br />

fiber or fewer connections, there is more signal power at the<br />

receiver, and the BER decreases.<br />

Suppose the system designer chose a maximum allowable<br />

BER <strong>of</strong> 10 -9 at some maximum allowable fiber length.<br />

Figure 3 was adapted from [4]. It shows that decreasing the<br />

fiber length by 40% in such a system (to 0.6 per unit),<br />

decreased the BER from 10 -9 to about 10 -23 , a decrease <strong>of</strong> 14<br />

orders <strong>of</strong> magnitude! Random bit errors cease to be an<br />

important source <strong>of</strong> unavailability.<br />

Unavailability then becomes dominated by other factors<br />

such as fiber breaks, misapplications, etc. Therefore, welldesigned<br />

fiber-optic communications systems will result in<br />

long periods <strong>of</strong> error-free performance separated by complete<br />

outages caused by human factors or equipment failure.<br />

BER<br />

Figure 3<br />

Bit Error Rate vs. Normalized Fiber Length<br />

10 -10<br />

10 -15<br />

10<br />

-25<br />

10<br />

0.5 0.6 0.7 0.8 0.9 1 1.1 1.2<br />

10 -5 Fiber Length (1.0 at BER=1.0E-9)<br />

Conservative Designs Yield Near-Zero BER on Fiber Links<br />

Does this mean we can ignore communications security<br />

No. Consider what happens to a fiber-optic receiver when a<br />

user disconnects one end <strong>of</strong> the fiber while the link is in<br />

service. The disconnection process is slow compared to most<br />

fiber-optic data transfer rates. As the user draws the fiber<br />

away from the receiver, attenuation increases and ambient<br />

light begins to flood the receiver. This causes the bit error rate<br />

to increase until the received bit stream is essentially all noise.<br />

The receiving device must recognize this noise and reject the<br />

corrupted data, or a misoperation may result. The security<br />

built into the 36-bit message described earlier is sufficient to<br />

ensure less than a 10 -7 chance that disconnecting the fiber<br />

could cause an undetected error. A security counter <strong>of</strong> just two<br />

virtually eliminates the risk, even for direct tripping.<br />

Can direct fiber channels be affected by faults There is<br />

some risk that the physical event which breaks the fiber could<br />

cause the fault, such as a tower collapse or static wire failure<br />

where the fibers are in the static wire. Tornadoes, ice loading,<br />

or an airplane collision are also possibilities. Even then, there<br />

is some chance the message will get through to permit the<br />

scheme to work, after the fault occurs and before the channel<br />

is destroyed.<br />

When a dedicated fiber is closely associated with the power<br />

line right-<strong>of</strong>-way, the probability that an external fault will<br />

cause a communications disturbance is negligible.<br />

B. Multiplexed Fiber<br />

Fiber-optic multiplexers combine many relatively slow<br />

digital and analog channels into one wideband light signal.<br />

The multiplexer, therefore, makes efficient use <strong>of</strong> bandwidth<br />

in the fiber. A direct digital connection between the relay and<br />

the multiplexer is more reliable and economical than<br />

interfacing through conventional relay contacts, then a tone<br />

set, and into an analog channel on the multiplexer. The<br />

multiplexer adds a level <strong>of</strong> complexity, which can be avoided<br />

by the simple dedicated fiber approach discussed earlier.<br />

C. Fiber-Optic Networks<br />

Wide-area networks, such as SONET, move large<br />

quantities <strong>of</strong> data at high speed. Many such networks consist<br />

<strong>of</strong> self-healing rings.<br />

Since the self-heal time is long compared to expected<br />

protective relay tripping times, we must still be concerned<br />

with correlation between faults and communications<br />

problems.<br />

There is a trade<strong>of</strong>f between long-term availability and<br />

short-term dependability. The ring self-heals so that<br />

communications are rarely totally lost. However, a failure<br />

anywhere in the network results in a short communications<br />

loss. <strong>Power</strong> Networking [5] describes a cascaded ring<br />

topology that reduces the exposure to these short interruptions.<br />

While the ring is self-healing, the terminal equipment is<br />

generally not. Thus the terminal equipment, and possibly other<br />

points, must be considered as possible single points <strong>of</strong><br />

failure—even though we have a self-healing ring.<br />

D. Multiplexed Microwave<br />

Microwave systems have gone digital, too—opening new<br />

opportunities for direct relay-to-relay communications. (Later<br />

we describe a low-delay modem, which can be used to transfer<br />

digital information through analog microwave channels, with<br />

the quality required for pilot protection.)<br />

Microwave equipment failures include multiplexers, radio<br />

gear, antenna pointing errors, cabling, etc. Microwave<br />

communications are fairly immune to power system faults. In<br />

general, the likelihood <strong>of</strong> a communication failure for an<br />

internal fault is not much different from the likelihood <strong>of</strong> a<br />

failure for an external fault.<br />

E. Narrow-Band UHF Radio<br />

Dedicated radios have been used for pilot channels.<br />

Reference 1 describes how a 960 MHz radio link was used in<br />

a POTT scheme. The radio was purchased with a single on<strong>of</strong>f-keyed<br />

tone interface between the radio and the relay<br />

Digital Communications for <strong>Power</strong> System Protection: Security, Availability, and Speed | 33


contact I/O. The security <strong>of</strong> this scheme comes from the<br />

inherent security <strong>of</strong> POTT schemes.<br />

Narrow-band digital radios permit the use <strong>of</strong> direct digital<br />

communications. The performance would be similar to that <strong>of</strong><br />

the microwave system given earlier, but may be more reliable<br />

than a channel in a microwave system, because <strong>of</strong> lower<br />

complexity.<br />

Radio channels are relatively immune to interference from<br />

faults. One possible source <strong>of</strong> interference is the power wiring<br />

to the radio. Radio channels might also be disrupted by<br />

antenna-pointing errors and severe weather.<br />

Radio and fiber channels can be highly complementary.<br />

Mechanical damage that might disrupt a fiber channel is<br />

generally unlikely to interfere with a radio channel, and vice<br />

versa. However, it is possible to conceive <strong>of</strong> events that would<br />

destroy both communications channels and cause a fault. For<br />

example, suppose a radio tower collapses during an<br />

earthquake or windstorm, and falls through the static wire<br />

with the fiber in it, and then causes a line fault!<br />

F. Spread-Spectrum Radio<br />

Spread-spectrum techniques have been broadly applied in<br />

radar systems to increase the energy in a radar pulse, while<br />

maintaining and enhancing target resolution. Spread-spectrum<br />

communications are used in military applications for the<br />

advantages <strong>of</strong> communications security, interference<br />

immunity, low probability <strong>of</strong> detection, and difficulty in<br />

jamming. Commercial uses <strong>of</strong> spread-spectrum radio have<br />

been growing, ever since the Federal Communications<br />

Commission permitted license-free operation under certain<br />

conditions. For power system protection, the advantages <strong>of</strong><br />

spread-spectrum radio channels are immunity to interference,<br />

freedom from licensing requirements, and low cost.<br />

Signals may be spread in the frequency domain by several<br />

methods. For example, frequency hopping, either slow or fast<br />

compared to the information rate, spreads the signal over the<br />

spectrum covered by dozens <strong>of</strong> discrete frequencies occupied<br />

sequentially in time, in a pseudo-random sequence. Directsequence<br />

spread spectrum systems, on the other hand,<br />

multiply the information bit stream by a much faster pseudorandom<br />

binary sequence. The bandwidth is expanded by the<br />

fast rate <strong>of</strong> the pseudo-random sequence.<br />

The processes <strong>of</strong> spreading, despreading, synchronization,<br />

and forward error correction (FEC) take some time, and,<br />

depending on the scheme, may be too slow for teleprotection.<br />

Most presently available radios also rely on the data terminal<br />

equipment (DTE) to negotiate a half-duplex channel.<br />

However, at least one model automatically switches its halfduplex<br />

channel rapidly enough to simulate full-duplex<br />

transmission at speeds as high as 19,200 baud. This same<br />

radio performs no FEC, and so has a very reasonable roundtrip<br />

delay <strong>of</strong> about 2.5 power system cycles. The cost, power<br />

requirements, and performance are promising for applications<br />

all the way down to distribution voltages [7].<br />

G. Digital Telephone Circuits<br />

Telephone companies <strong>of</strong>fer leased digital lines for several<br />

hundred dollars per month, and these can be used for pilot<br />

protection schemes. A CSU/DSU interfaces the protective<br />

relay to the leased line. It receives timing information from the<br />

telephone company equipment via the leased line, and passes<br />

that timing information on to the relay (for synchronous data)<br />

or synchronizes the asynchronous data stream from the relay<br />

(for asynchronous data). It also converts the serial data<br />

received from the relay to the proper electrical levels and<br />

format.<br />

The digital data are not modulated on the twisted pair in the<br />

traditional sense; they remain binary (actually ternary) while<br />

on the leased line. Such communications are characterized by<br />

long periods between short bursts <strong>of</strong> errors. For example, the<br />

standard AT&T Technical Reference TR 62310 [8] defines<br />

acceptable performance <strong>of</strong> a 56kbps using the concept <strong>of</strong> an<br />

error-second (ES) and a severe-error-second (SES). At<br />

56 kbps, an ES has between one and 56 bit-errors. A SES has<br />

more than 56 bit-errors. That standard allows 20 error-seconds<br />

and six severe-error-seconds per day.<br />

If we assume each ES results in 1 second <strong>of</strong> protection<br />

scheme unavailability, then communications-assisted<br />

protection might not be available for 26 seconds per day. This<br />

is a very extreme upper limit.<br />

Like fiber-optic links, digital leased lines are distancesensitive.<br />

Far from the distance limit, the actual performance<br />

is significantly better than the worst-case prediction above.<br />

Never accept anything close to the worst-case scenario<br />

depicted above. We have experience with digital leased lines<br />

that produced less than one error every three days. Earlier in<br />

this paper, we presented data from a digital line that was<br />

unavailable for approximately 1 minute during a 40 day<br />

period. This is 17 times better than the worst-case scenario.<br />

Another factor affecting error rates on digital leased lines is<br />

transmit power. In AT&T Data Communications TR 62310<br />

transmit power is restricted at 9,600 and 12,400 bits per<br />

second to 1/4 <strong>of</strong> the power allowed at other rates. At least one<br />

CSU/DSU manufacturer takes advantage <strong>of</strong> the higher<br />

allowed transmit power at data rates other than 9,600 and<br />

12,400 bits per second. As with fiber optics, increasing the<br />

transmit power by a factor <strong>of</strong> four on a twisted pair can have a<br />

pr<strong>of</strong>ound impact on random bit errors. Assuming a leased line<br />

takes advantage <strong>of</strong> the higher permitted power, then rates<br />

other than 9,600 are greatly preferred.<br />

Any leased line must be galvanically isolated between the<br />

substation and the central <strong>of</strong>fice. This isolation prevents<br />

damage and danger when ground faults produce high voltages<br />

between the substation ground and the telephone exchange<br />

[9]. However, isolation does not guarantee that the leased line<br />

will remain operational during the fault. Ground potential rise<br />

or noise coupled from the faulted power line to the twisted<br />

pair can produce enough noise on the circuit to cause bit errors<br />

or a complete loss <strong>of</strong> signal. The analysis required to<br />

determine if a circuit will remain operational would be<br />

difficult.<br />

34 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


Noise from faults or other sources that might not corrupt<br />

signaling by audio tones over a given twisted pair still might<br />

interfere with fast digital signaling over that same channel.<br />

(Recall that faster signaling requires a higher SNR, given the<br />

same bandwidth.)<br />

H. Analog Voice-Grade Channels<br />

Figure 4 shows a low-delay modem interface between the<br />

direct digital data from the protective relay, and a voice<br />

channel. The voice channel may be analog microwave, a<br />

leased circuit from the telephone company, a dedicated<br />

twisted pair, or something similar. Analog channels on<br />

microwave should not suffer the same degradation during<br />

power system faults as analog channels on a twisted pair or<br />

some other conductor. The challenge becomes modulating the<br />

9600 bit per second data stream so it will be compatible with<br />

the 300 to 3,000 Hz audio band channel.<br />

Relay<br />

U<br />

A<br />

R<br />

T<br />

Figure 4<br />

Asynchronous<br />

Serial Bit<br />

Stream<br />

<strong>Power</strong><br />

Low-Delay Modem<br />

Quadrature<br />

Amplitude<br />

Modulated<br />

Direct Digital Signaling Over Voice Channel<br />

Voice Channel<br />

Computer modems generally are unsuitable for protection.<br />

They are optimized for throughput, at the expense <strong>of</strong> delay.<br />

Therefore a special low-delay modem was developed. It also<br />

takes a very short time to adapt to changes in the channel<br />

(retraining), compared to computer modems [10].<br />

V. PILOT PROTECTION APPLICATIONS<br />

Direct underreaching transfer trip schemes require<br />

extremely secure communications, because there is no local<br />

confirmation that a fault exists. DUTT is very simple. The<br />

digital communications described in this paper provide much<br />

greater security than that required by IEC 834-1 for direct<br />

tripping, when the protection scheme uses a security count <strong>of</strong><br />

two or greater. It is inherently secure from current reversals.<br />

Permissive underreaching transfer trip does not require as<br />

secure a level <strong>of</strong> communications, because the received<br />

underreaching element is qualified by a local overreaching<br />

element. The scheme is also quite simple, and requires no<br />

current-reversal logic. Sensitivity is very similar to DUTT.<br />

Permissive overreaching transfer trip schemes provide<br />

greater sensitivity, and have the same channel dependency as<br />

PUTT. In most cases, POTT schemes must be protected<br />

against current reversals. POTT schemes handle weak<br />

terminals, when the schemes include echo logic. If internal<br />

faults cause channel failures, then POTT schemes may not<br />

operate.<br />

Directional comparison blocking schemes provide very<br />

similar speed and sensitivity to POTT with echo logic, yet<br />

DCB schemes do not require echo logic. DCB schemes must<br />

also be protected against current reversals. If external faults<br />

cause channel failures, then DCB schemes will overtrip. The<br />

security and practicality <strong>of</strong> DCB schemes depend on known<br />

and reasonable upper limits on element pickup times and<br />

channel delays.<br />

Directional comparison unblocking schemes attempt to<br />

give the best <strong>of</strong> POTT and DCB. DCUB schemes only make<br />

sense when we can definitely associate a much greater<br />

likelihood <strong>of</strong> channel failures with internal faults, than with<br />

external faults. For example, DCUB might be sensible for a<br />

power-line carrier channel, or an optical-fiber in the shield<br />

wire <strong>of</strong> the protected line.<br />

Consider some pilot-scheme possibilities, given different<br />

channels.<br />

A. Fiber-Optic Ring<br />

A good approach is POTT, with weak-infeed and openbreaker<br />

echo. DCB should not be used because: security<br />

depends directly on availability <strong>of</strong> the communications, we<br />

cannot associate channel failures with internal faults, and<br />

communication delays may depend on routing. DCUB should<br />

not be used, because we cannot associate channel failures with<br />

internal faults.<br />

B. Leased Digital Line<br />

As with the ring, it is not generally possible to associate<br />

channel failures with either internal or external faults, and<br />

delays may be variable. Therefore DCB and DCUB should be<br />

avoided.<br />

C. Dedicated Fiber Optics<br />

If the fiber and the power line share the same path, then<br />

DCUB might be used to gain some availability, with little loss<br />

in security. This is because channel failures simultaneous with<br />

faults might reasonably be associated with internal faults.<br />

Sending multiple bits in each direction opens up some new<br />

possibilities in pilot protection.<br />

D. Cross-Country Faults<br />

Consider a double-circuit line from S to R, as shown in<br />

Figure 5. An AG fault on Line 1 simultaneously exists with a<br />

BG fault on Line 2. In a single-pole-tripping scheme, the<br />

desired action is for Line 1 to trip phase A, and Line 2 to trip<br />

phase B, so service is essentially uninterrupted between S and<br />

R. If the relays at R communicate the observed fault types,<br />

then the relays at S can trip single-pole and avoid the<br />

undesired three-pole trips for this situation. The fault type is<br />

easily communicated on three bits, or on just two, with some<br />

encoding.<br />

Figure 5 Communicate Fault Type for Secure Single-Pole Tripping<br />

Digital Communications for <strong>Power</strong> System Protection: Security, Availability, and Speed | 35


E. Remote-End Open Keying<br />

Given a line from S to R, assume the breaker at R is open.<br />

A POTT scheme needs to inform S that R is open.<br />

Traditionally, there are two ways. One is for S to send<br />

permission to R and for R to “echo” the permission back to S<br />

based on the open-breaker status. This status is sensed by<br />

monitoring an auxiliary contact at R. Breaker S trips after two<br />

communications delays, plus possibly an echo time-delay in<br />

terminal R. Given 16 ms for the protective relay element at S,<br />

12 ms for each communications delay, and no echo time-delay<br />

in R, we have a total time <strong>of</strong> 40 ms from fault inception until S<br />

trips. (Another timer is used to limit the echo duration to a few<br />

cycles, so the channel does not lock up.)<br />

A second method is for terminal R to send S a standing<br />

permission whenever the breaker at R is open. Both<br />

communications delays are eliminated, so tripping can occur<br />

in just relay-element time, e.g., 16 ms. With conventional<br />

channels, this method makes a compromise with security<br />

because guard-before-trip cannot be used. However, with<br />

direct digital communications, the security is built into the<br />

message so no loss <strong>of</strong> security occurs with hard-keying.<br />

An alternative is shown in Figure 6, where the remote end<br />

transmits its breaker status and breaker control commands to<br />

the local end. Overreaching elements trip the local breaker, as<br />

long as the remote end is open. The philosophy here is for the<br />

remote end to send the local end the state <strong>of</strong> the breaker so<br />

that the local end can directly observe it and use it as desired,<br />

in this scheme and possibly others. If the local end receives<br />

notice <strong>of</strong> close commands from the remote end, then this<br />

scheme can be briefly delayed to avoid risk <strong>of</strong> misoperation by<br />

very sensitive elements due to pole scatter. Other advantages<br />

are that this scheme is never encumbered by current-reversal<br />

timers, and it is possible to use different overreaching<br />

elements when the breaker is open than when it is closed.<br />

Figure 7 Avoid Lockup With Simple POTT Scheme<br />

G. Control and Balance <strong>of</strong> Speed, Sensitivity, and Security<br />

With multiple bits <strong>of</strong> information to transfer, we can<br />

consider schemes that simultaneously coexist and provide the<br />

advantages <strong>of</strong> each, while minimizing the risk <strong>of</strong> their<br />

individual disadvantages. Figure 8 and Figure 9 illustrate the<br />

concept for ground faults. The design uses an individual bit<br />

(channel) for each relay element that is desired at the remote<br />

end—instead <strong>of</strong> combining them into one. Then each terminal<br />

uses the set <strong>of</strong> local and remote elements to make its<br />

decisions, as desired or required by the immediate operating<br />

conditions.<br />

1<br />

2<br />

Remote<br />

Elements<br />

Zone 1<br />

Zone 1.5<br />

Local<br />

Elements<br />

Zone 1<br />

C<br />

0<br />

C Security Counts<br />

Zone 1.5<br />

Direct Trip<br />

Direct Transfer Trip<br />

PUTT or Short-POTT<br />

3<br />

Zone 2<br />

Zone 2<br />

POTT<br />

4<br />

Zone 3<br />

0<br />

T<br />

Figure 6<br />

Line Protection for Open-Remote-End<br />

Zone 3<br />

0<br />

T<br />

F. Simplify Pilot Logic<br />

When a single bit is available, say in a POTT scheme, a<br />

timer and additional logic are necessary to avoid latching the<br />

channels on during echo. The logic is simpler if we elect to<br />

use individual bits to transmit the status <strong>of</strong> individual relay<br />

elements, and <strong>of</strong> the breaker. Each terminal builds up its trip<br />

outputs from the locally-observed and remotely-reported relay<br />

elements without the need for feedback paths that lead to<br />

channel lock-up or other surprises. Figure 7a shows how a<br />

traditional POTT scheme avoids channel lock-up with extra<br />

logic and a timer. Figure 7b shows a simpler scheme, which<br />

does not have the risk <strong>of</strong> lock-up because there is no feedback<br />

path.<br />

5<br />

6<br />

50N<br />

ROK<br />

50N<br />

Figure 8 Balance Sensitivity, Speed, and Security<br />

T<br />

T<br />

0<br />

0<br />

S Not Blocked<br />

R Not Blocked<br />

36 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


80<br />

60<br />

40<br />

20<br />

R F<br />

100<br />

3 3<br />

S<br />

2<br />

5 6<br />

4<br />

1.5 1.5<br />

4 4<br />

1 3<br />

1<br />

2<br />

1<br />

2<br />

Figure 9 Fault Resistance Regions Covered by the Schemes in Figure 8<br />

The standard IEC 834-1 suggests that the user should<br />

decide the states <strong>of</strong> communicated bits, given a loss <strong>of</strong><br />

communications. For example, during a loss <strong>of</strong><br />

communications, should the received bit maintain (hold) its<br />

previous state, or default to a 1, or to a 0 As we apply bits in<br />

various schemes below, we will also determine the desired<br />

action <strong>of</strong> the bits given communication loss.<br />

A local Zone 1 element, such as a ground quadrilateral<br />

element, trips the breaker directly. Non-pilot tripping<br />

covers the area <strong>of</strong> overlap <strong>of</strong> the zones 1, marked by the<br />

circled 1 in the house-shaped characteristic in Figure 9.<br />

We must rely on the channel, or wait for zone timers<br />

for all other faults. Appendix A gives some guidelines<br />

for setting the resistance and reactance reaches, taking<br />

into consideration some <strong>of</strong> the angle errors that might<br />

be expected.<br />

The remote Zone 1 element state is communicated on<br />

one bit and passes through a local security counter to<br />

provide a Direct Transfer Trip. The additional regions<br />

covered are marked by the circled 2s. Given a channel<br />

failure, the received bit should default to a ‘0’. (A<br />

default <strong>of</strong> ‘1’ would trip the remote breaker for<br />

communications failures; a default <strong>of</strong> ‘hold’ defeats the<br />

receive security counter.)<br />

PUTT and short-POTT schemes follow:<br />

In a PUTT scheme (not shown in Figure 8), the local<br />

Zone 2 element would be set to reach beyond the<br />

remote end. The remote Zone 1 element is an<br />

underreaching element. The scheme is fast and<br />

simple—and is not encumbered by current-reversal<br />

timers or logic, which might delay tripping for faults<br />

that evolve to the healthy line.<br />

A short-POTT scheme can also be considered here, and<br />

is illustrated in the figures. The Zone 1.5 elements at<br />

2<br />

R F<br />

R<br />

each terminal would be set to cover the entire line, but<br />

would have their reaches limited sufficiently to avoid<br />

current-reversals on parallel paths. This improves faultresistance<br />

coverage and speed over PUTT, but does not<br />

involve current-reversal timing.<br />

Elements for a short-POTT scheme are labeled 1.5, and<br />

cover the additional area marked with the circled 3.<br />

The received bits should default to ‘0’ for<br />

communications failures. There is no benefit to<br />

holding—only risk.<br />

A POTT scheme based on directional overcurrent<br />

elements, labeled Zone 2, can provide some more<br />

sensitivity, but must be guarded against current<br />

reversals. Reverse-looking elements (Zone 3 directional<br />

overcurrent elements) and timers provide blocking for a<br />

short time after current reverses. The delay that would<br />

result if a fault evolved from one line to the other is not<br />

important, because <strong>of</strong> the coverage by the PUTT or<br />

short-POTT schemes. A timer could be added after the<br />

AND-gate, to gain some security against system<br />

unbalances produced by switching, for example. The<br />

additional coverage from the POTT scheme is labeled<br />

with the circled 4.<br />

Received bits for Zone 2 should default to ‘0’ on loss <strong>of</strong><br />

channel. A ‘hold’ or ‘1’ would significantly risk<br />

misoperation, with little or no benefit in dependability.<br />

The Zone 3 received bits should default to a ‘0,’ if<br />

communications are lost. A default to a ‘1’ would<br />

defeat the protection for T seconds for every bit error.<br />

A high-resistance fault close to S could be in region 5,<br />

which is not covered by the POTT scheme. After some<br />

time delay, and if no reverse elements pick up at the<br />

remote end, we can trip for such a fault, with the<br />

sensitivity we are accustomed to with DCB schemes.<br />

However, this scheme enjoys much greater security<br />

than DCB.<br />

Conventional blocking schemes assume the fault is<br />

internal if the blocking message is not received. (The<br />

truth might be that the channel or relay equipment<br />

failed to deliver the message to block. Put another way,<br />

in traditional DCB schemes, the block signal means<br />

“reverse.” No block signal means “forward,” “not<br />

detected,” or “bad channel.”)<br />

Here, the local terminal knows with near certainty the<br />

states <strong>of</strong> the remote forward and reverse elements, or it<br />

knows the channel is down. There is no confusing a lost<br />

channel with an internal fault. The fact that we are<br />

receiving messages from the remote end reassures us<br />

that the remote relay is functioning, and ready to<br />

produce blocking signals when appropriate.<br />

In addition, the security can be enhanced somewhat by<br />

an undercurrent element, 50N. This element would<br />

block tripping, if enough residual current is produced<br />

by an open CT, for instance.<br />

Digital Communications for <strong>Power</strong> System Protection: Security, Availability, and Speed | 37


The received-OK signal (ROK) ensures this part <strong>of</strong><br />

the logic is active only if the remote end is<br />

successfully communicating, and therefore in a<br />

position to block the local end should an external<br />

fault occur.<br />

If the source behind R is very weak, then Zone 2 at R<br />

might not be able to see all the way back to S. This is<br />

the only scheme described here that is capable <strong>of</strong><br />

detecting the fault.<br />

A high-resistance fault near R can be cleared in a<br />

complementary way, as compared to above. The<br />

ROK signal is not necessary, assuming the Zone-2-<br />

received signal defaults to a zero on communications<br />

loss.<br />

Protection speeds can be impressive, as the following quick<br />

look at a POTT scheme reveals. Consider a direct digital link<br />

at 19,200, and allow 6 ms for the message and processing<br />

time. Assume the overreaching elements operate in under<br />

12 ms, and the relays use instantaneous trip contacts, such as<br />

transistors. Both terminals trip their breakers in less than one<br />

cycle. With a 1.5-cycle breaker, the fault is cleared in under<br />

2.5 cycles, at both ends. Given such fast tripping, and also<br />

given that the channel could be used for breaker failure, we<br />

should investigate shorter breaker-failure times and faster<br />

time-step backup.<br />

In summary, the direct tripping Zone 1 elements cover very<br />

little <strong>of</strong> the line when the channel is not available. The<br />

coverage is limited to the small region <strong>of</strong> overlap near the<br />

middle <strong>of</strong> the line. The direct transfer trip path depends on<br />

secure measurements from one end, but involves a short<br />

security delay <strong>of</strong> about 4 ms. Because it depends on<br />

information from the remote end alone, it can trip the local<br />

breaker even if there is a problem at the local terminal such as<br />

a loss-<strong>of</strong>-potential (LOP) condition. The short-POTT path<br />

approximately doubles the fault resistance coverage as<br />

compared to the DT and DTT paths, but could be disabled by<br />

a LOP condition at either end. The POTT scheme adds<br />

sensitivity and speed (which might be sacrificed with an extra<br />

timer, if temporary unbalances could pick up the sensitive<br />

overreaching elements). Again, both ends must determine the<br />

fault to be internal. The last two schemes cover faults in the<br />

“bow-tie” regions labeled with the circled 5 and 6. Some time<br />

delay is required to wait for the possible block from the other<br />

end. Overcurrent elements can block these two regions, to<br />

ensure they are only active for low-level currents, thereby<br />

reducing the risk <strong>of</strong> misoperation should a current transformer<br />

fail at either end.<br />

It should be noted that the schemes above are presented as<br />

a concept, to show how communicating multiple bits <strong>of</strong><br />

information end-to-end can produce an adaptive and balanced<br />

protection scheme.<br />

VI. CONCLUSIONS<br />

1. The high quality <strong>of</strong> many digital communications<br />

channels permits more information to be sent in less<br />

time, as predicted by Shannon and demonstrated in<br />

practice.<br />

2. Direct-digital communications between relays can be<br />

designed with the security, speed, dependability, and<br />

adaptability needed for blocking, permissive, and<br />

direct-tripping applications—as well as for control.<br />

3. Eight bits can be securely communicated every 2 ms,<br />

with a worst-case end-to-end delay <strong>of</strong> 6 ms, including<br />

processing latency, over a 19,200 bit/second channel.<br />

4. Extremely secure direct transfer tripping can occur in<br />

less than 6 ms, when a two-message sequence is<br />

employed over a 19,200 bit/second channel.<br />

5. Direct communications using a simple serial<br />

asynchronous bit stream ensures compatibility with a<br />

wide variety <strong>of</strong> communications channels, systems,<br />

and test equipment.<br />

6. Channel performance monitoring, including sequence<br />

<strong>of</strong> events, outage duration, and unavailability provides<br />

the measurements <strong>of</strong> performance required to maintain<br />

and improve communications—without periodic<br />

testing. System operation is a continuous test <strong>of</strong> the<br />

channel.<br />

7. Relay event reports closely relate the performance <strong>of</strong><br />

the protection and the communications, during faults<br />

and other events.<br />

8. Fiber-optic networks and links, digital microwave<br />

systems, and point-to-point radio (narrow-band and<br />

spread-spectrum) are excellent channels to consider<br />

for direct digital-to-digital applications. The<br />

combination <strong>of</strong> lower-cost channels with direct relayto-relay<br />

communications opens up applications at<br />

lower voltages, including distribution feeders.<br />

9. Although metallic circuits have demonstrated<br />

satisfactory performance in the field, we must consider<br />

ground potential rise, isolation, and induced<br />

interference during faults. Channel and event<br />

monitoring can quickly point out difficulties, should<br />

they appear, and lead us to their resolution.<br />

10. High-quality analog channels, such as analog<br />

microwave, can be used for digital communications at<br />

protection speeds with the help <strong>of</strong> a limited-delay<br />

modem.<br />

11. As always, the channel characteristics, including<br />

signal routing, must be considered in selecting and<br />

designing protection schemes.<br />

12. Because channels fail in different ways, using<br />

different channel types can provide redundancy<br />

against failures produced by faults.<br />

13. Communicating eight bits end-to-end opens<br />

opportunities for new protection schemes, and for<br />

combining traditional schemes for enhanced<br />

performance.<br />

14. Schemes can be simpler, and some problems, such as<br />

channel lockup, can be avoided or solved more simply<br />

when more than one bit is available end-to-end.<br />

38 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


15. A POTT scheme implemented over a<br />

19,200 bit/second channel can clear at both ends in<br />

less than one cycle, plus breaker times.<br />

VII. REFERENCES<br />

[1] Shannon, C.E., 1948, “A Mathematical Theory <strong>of</strong> Communication,” The<br />

Bell System Technical <strong>Journal</strong>, Volume XXVII, July.<br />

[2] IEC 834-1, Performance and Testing <strong>of</strong> Teleprotection Equipment <strong>of</strong><br />

<strong>Power</strong> Systems. Part 1: Command Systems, May 1996.<br />

[3] OSD 136L, Optical Systems Design Pty Ltd 7/1 Vuko Place,<br />

Warriewood 2102, NSW Australia; http://www.optsysdesign.com<br />

[4] Haus, Robert J., 1980, Fiber Optics Communications Design Handbook<br />

pg 158. Inglewood Cliffs, New Jersey: Prentice Hall.<br />

[5] Gardner, Terry N., 1998, “<strong>Power</strong> Networking”, Transmission &<br />

Distribution World, February.<br />

[6] Kahler, J. Kraig, “Installation <strong>of</strong> 930-960 MHz Low Density Point-To-<br />

Point Radios and Solid State Relays for Primary Transmission Relay<br />

Protection on 69 kV Transmission Lines,” 21st Annual Western<br />

Protective Relay Conference, Spokane, Washington, October 18–20,<br />

1994.<br />

[7] Roberts, J. and Zimmerman, K, “Trip and Restore Distribution Circuits<br />

at Transmission Speeds,” 25th Annual Western Protective Relay<br />

Conference, Spokane, Washington, October 13–15, 1998.<br />

[8] AT&T Technical Reference TR 63210, “DS0 Digital Local Channel,<br />

Description and Interface Specification,” August 1993.<br />

[9] IEEE Std 487-1992, IEEE Recommended Practice for the Protection <strong>of</strong><br />

Wire-Line Communication Facilities Serving Electric <strong>Power</strong> Stations,<br />

November 4, 1992.<br />

[10] MBT 9600, Pulsar Technologies, Inc., 4050 NW 121st Ave, Coral<br />

Springs, FL 33065; http://www.pulsartech.com.<br />

[11] Schweitzer, E.O. III and Roberts, J, “Distance Relay Element Design,”<br />

19th Annual Western Protective Relay Conference, Spokane,<br />

Washington, October 20–22, 1992.<br />

VIII. APPENDIX A: QUADRILATERAL REACTIVE REACH<br />

VERSUS RESISTIVE REACH SETTING GUIDELINE<br />

A. Quadrilateral Element Review<br />

To pick up a forward-reaching zone <strong>of</strong> quadrilateral ground<br />

distance protection, the relay must determine that the fault<br />

presented to the relay passes the following four measurement<br />

test criteria:<br />

• Reactance < set reactance (top line)<br />

• Apparent fault resistance (RF) < positive-resistance<br />

(right-side) blinder<br />

• RF > negative-resistance (left-side) blinder<br />

• Fault direction is forward as measured by a negative- or<br />

zero-sequence polarized ground directional element<br />

Equations 1 and 2 repeat the equations shown in [11] for<br />

the Zone 1 A-Phase reactance and resistance tests.<br />

R<br />

X<br />

AF<br />

AG1<br />

jT<br />

∗<br />

( A<br />

⋅( R<br />

⋅ ) )<br />

jT<br />

( ) ( )<br />

Im V I e<br />

=<br />

Im Z I k I I e<br />

∗<br />

( 1L<br />

⋅<br />

A<br />

+<br />

0<br />

⋅<br />

R<br />

⋅<br />

R<br />

⋅ )<br />

∗<br />

( A<br />

⋅( 1L<br />

⋅ ( A<br />

+<br />

0<br />

⋅<br />

R ))<br />

)<br />

Im V Z I k I<br />

=<br />

⎛ 3<br />

Im ⎜ ⋅ I<br />

⎝ 2<br />

+ I ⋅ Z ⋅ I + k ⋅ I<br />

( A2 0 ) ( 1L ( A 0 R ))<br />

∗<br />

⎞<br />

⎟<br />

⎠<br />

(1)<br />

(2)<br />

Where:<br />

Im = Imaginary portion<br />

V A = A-Phase voltage, [V]<br />

I A = A-Phase current, [A]<br />

I R = Residual Current (I A + I B + I C ), [A]<br />

Z 1L = Positive-sequence replica line impedance, [Ω]<br />

Z 0L = Zero-sequence replica line impedance, [Ω]<br />

k 0 = (Z 0L - Z 1L )/(3• Z 1L ), [unitless]<br />

I A2 = Neg.-sequence current (I A + a 2 •I B + a•I C ), [A]<br />

I 0 = Zero-sequence current (I R /3), [A]<br />

T = Nonhomogeneous system factor, [degrees]<br />

* = Complex conjugate operator<br />

B. Calculating Reactance Reach As a Function <strong>of</strong> Resistive<br />

Reach<br />

The elements described by Equations (1) and (2) are phase<br />

angle comparators. For the reactance element described by (1),<br />

when the angle between the polarizing quantity (I R ) and the<br />

line drop compensated voltage (Z 1L •(I A + k 0 •I R ) – V) is 0°, the<br />

impedance is on the reactance element boundary. This element<br />

must measure line reactance without under- or overreaching<br />

from the effects <strong>of</strong> load flow or fault resistance. Hence, the<br />

element must use an appropriate polarizing current: negativeand<br />

zero-sequence currents are suitable choices. In some<br />

nonhomogeneous systems, the tip produced by the polarizing<br />

current may be insufficient to prevent overreach. To<br />

compensate for this nonhomogeneity, we introduce polarizing<br />

current angle bias (tip), or reduce the reach <strong>of</strong> the Zone 1<br />

element.<br />

Reducing the Zone 1 reach restricts that portion <strong>of</strong> the line<br />

protected by overlapping instantaneous Zone 1 protection.<br />

This overlapping “zone” is realized for low-resistance faults.<br />

As we show next, a large resistive reach can limit the<br />

reactance element reach when we consider instrumentation<br />

angle errors. If the quadrilateral ground distance elements are<br />

the only Zone 1 protection, then we strike a balance between<br />

overlapping zone for mid-line faults, and large resistive<br />

coverage by one terminal for close-in faults.<br />

Specifically, the instrumentation angle errors we consider<br />

are those due to current transformers (CTs), voltage<br />

transformers (VTs), and the measuring relay. For our example,<br />

the values <strong>of</strong> these angles are: CT = 1°, VT = 2°, Relay<br />

Measurement = 0.2°<br />

Digital Communications for <strong>Power</strong> System Protection: Security, Availability, and Speed | 39


Let us consider Relay R shown in Figure A.1.<br />

Figure A.1 System-Single Line and First Quadrant <strong>of</strong> the Quadrilateral<br />

Distance Characteristic at Source S Terminal<br />

For a ground fault outside <strong>of</strong> the protected zone with a<br />

reach m [XAG1 <strong>of</strong> Equation (1)], what is the maximum secure<br />

reactive reach for a given resistive reach coverage [RAG1 <strong>of</strong><br />

Equation (2)]<br />

XL<br />

m⋅<br />

XL<br />

From Figure A.1: tan θ 1 = and tan θ2<br />

=<br />

R<br />

R<br />

Solve for m:<br />

m ⋅ X<br />

R<br />

∴<br />

L<br />

= tan<br />

( θ − ε)<br />

L<br />

1<br />

=<br />

⎛<br />

tan⎜<br />

tan<br />

⎝<br />

⎛<br />

⎜<br />

⎝<br />

−1<br />

R ⎛<br />

−1<br />

⎛ XL<br />

⎞ ⎞<br />

m = ⋅ tan tan ε<br />

X ⎜ ⎜ −<br />

⎝ ⎝ R<br />

⎟<br />

⎠ ⎟<br />

⎠<br />

X<br />

R<br />

L<br />

⎞ ⎞<br />

⎟ − ε⎟<br />

⎠ ⎠<br />

For R >> X L , tan -1 (X L /R) and tan(X L /R) ≅ X L /R. (Note: this<br />

approximation nets an error less than 5% for R/X L > 2.5). If<br />

we assume the protected system is homogeneous (i.e., the only<br />

angular errors we must account for are those <strong>of</strong> the CT, VT,<br />

and relay), ε = 3°≅ 1/20 radians. Given these simplifications:<br />

R ⎛⎛<br />

XL<br />

⎞ ⎞<br />

m = ε<br />

X ⎜⎜<br />

−<br />

L ⎝⎝<br />

R<br />

⎟<br />

⎠ ⎟<br />

⎠ = R ⋅ε<br />

1− =<br />

X<br />

1 R<br />

−<br />

(3)<br />

X 20<br />

L<br />

L<br />

⋅<br />

Equation 3 shows us that the lower the resistive reach, the<br />

greater the permissible reactance reach. Figure A.2 shows a<br />

graph <strong>of</strong> allowable resistive to reactive reach ratio for ε = 1/20<br />

radians (3°). The dashed line in this figure shows an example<br />

where an R/X L ratio = 8 (for a 1 ohm line and an 8-ohm<br />

resistive reach) permits setting m = 0.6 per-unit <strong>of</strong> the line.<br />

Figure A.2<br />

ε ≠ 0)<br />

Increase Reactance Reach By Decreasing Resistive Reach (for<br />

Copyright © <strong>SEL</strong> 1998, 2004<br />

(All rights reserved)<br />

Printed in USA<br />

20040126<br />

40 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


Transmission Line Protection System for<br />

Increasing <strong>Power</strong> System Requirements<br />

Armando Guzmán, Joe Mooney, Gabriel Benmouyal, Normann Fischer,<br />

Schweitzer Engineering Laboratories, Inc.<br />

Abstract—This paper describes a protective relay for fast and<br />

reliable transmission line protection that combines elements that<br />

respond only to transient conditions with elements that respond<br />

to transient and steady state conditions. In this paper, we also<br />

present an algorithm that prevents Zone 1 distance element<br />

overreach in series-compensated line applications and show how<br />

to prevent corruption <strong>of</strong> the distance element polarization during<br />

pole-open conditions. We also introduce an efficient frequency<br />

estimation logic for single-pole-tripping (SPT) applications with<br />

line-side potentials. This logic prevents distance element<br />

misoperation during a system frequency excursion when one pole<br />

is open. We also discuss an algorithm and logic to prevent singlepole<br />

reclosing while the fault is present, avoiding additional<br />

power system damage and minimizing system disturbance.<br />

Applying these algorithms and logics results in a protective<br />

system suitable for increasing power system requirements such<br />

as heavy loading, SPT, series line compensation, and shunt line<br />

compensation.<br />

I. INTRODUCTION<br />

Right-<strong>of</strong>-way restrictions and limitations on building new<br />

transmission lines necessitate optimization <strong>of</strong> transmission<br />

networks. This optimization imposes challenges on distance<br />

relay-based transmission line protection. Network<br />

optimization increases transmission line loading and requires<br />

fast fault clearing times because <strong>of</strong> reduced stability margins.<br />

In many cases, series compensation, SPT, or the<br />

combination <strong>of</strong> both is necessary to optimize transmission<br />

network investment. Series compensation generates<br />

subharmonics that can cause distance element overreach. SPT<br />

adds complexity to the ability <strong>of</strong> the distance element to track<br />

the power system frequency during single-pole open (SPO)<br />

conditions when the line protection uses line-side potentials.<br />

Shunt compensation can corrupt the distance element<br />

polarization because <strong>of</strong> the presence <strong>of</strong> transient voltages<br />

during three-pole open conditions when the line protection<br />

uses line-side potentials.<br />

SPT applications without arc extinction methods can<br />

jeopardize power system operation if the fault condition has<br />

not disappeared before the reclosing attempt; that is, the<br />

breaker closes under fault condition, making the power system<br />

prone to instability.<br />

Combining elements that respond only to transient<br />

conditions with elements that respond to transient and steady<br />

state conditions results in dependable, high-speed protection.<br />

Dedicated logic for series compensation applications adds<br />

security to the distance elements in the presence <strong>of</strong><br />

subharmonics. Flexible polarizing quantities and frequency<br />

tracking algorithms adapt to different breaker and system<br />

operating conditions. Secondary arc extinction detection<br />

optimizes the single-pole open interval and minimizes power<br />

system damage in SPT applications.<br />

In this paper, we present solutions to the above issues that<br />

result in improved transmission line protection suitable for the<br />

most challenging power system requirements.<br />

II. RELIABLE HIGH-SPEED TRIPPING<br />

A significant reduction in operating time is one trend in<br />

recently developed digital transmission line relays. A number<br />

<strong>of</strong> new techniques allow secure sub-cycle tripping <strong>of</strong><br />

transmission line distance relays. One <strong>of</strong> the first proposed<br />

methods uses variable-length data-window filtering with<br />

adaptive zone reach [1]. Developments described in Reference<br />

[2] introduced the concept <strong>of</strong> multiple data-window filters (a<br />

total <strong>of</strong> four) with corresponding fixed reach (the smaller the<br />

data-window, the smaller the reach). Line protective relays not<br />

only need to pick up fast for incipient faults but also to<br />

provide proper and fast fault type selection in SPT<br />

applications.<br />

A. High-Speed Distance Element Operating Principles<br />

Based on the principle <strong>of</strong> multiple data-window filters<br />

mentioned above [2], we developed the concept <strong>of</strong> the dualfilter<br />

scheme. This scheme combines voltage and current data<br />

from half-cycle and one-cycle windows (Figure 1) to obtain<br />

Zone 1 distance element detection and achieve fast tripping<br />

times.<br />

V, I<br />

One-cycle<br />

Filter<br />

Half-cycle<br />

Filter<br />

One-cycle<br />

mho Calc.<br />

Half-cycle<br />

mho Calc.<br />

Zone 1<br />

Reach<br />

Reduced<br />

Zone 1<br />

Reach<br />

_<br />

+<br />

_<br />

+<br />

Zone 1<br />

Detection<br />

Figure 1 Concept <strong>of</strong> a Zone 1 Distance Element Using Dual-Filter Scheme<br />

The mho calculation in Figure 1 uses operating (S OP ) and<br />

polarizing (S POL ) vector quantities defined as follows to<br />

implement mho distance element calculations [3]:<br />

S OP = r ⋅ ZL1 ⋅ IR − VR<br />

(1)<br />

S POL = V POL<br />

(2)<br />

Where:<br />

V R = line voltage <strong>of</strong> the corresponding impedance loop<br />

I R = line current <strong>of</strong> the corresponding impedance loop<br />

Z L1 = positive-sequence line impedance<br />

r = per-unit mho element reach<br />

V POL = polarizing voltage<br />

Transmission Line Protection System for Increasing <strong>Power</strong> System Requirements | 41


V R and I R are the relay voltage and current phasors<br />

particular to an impedance loop (six loops are necessary to<br />

detect all faults), and V POL is the polarizing voltage, consisting<br />

<strong>of</strong> the memorized positive-sequence phasor [3]. A mho<br />

element with reach r detects the fault when the scalar product<br />

between the two vectors is positive (i.e., the angle difference<br />

between S OP and S POL is less than 90 degrees). We can<br />

represent this condition mathematically as:<br />

( )<br />

*<br />

Re al ⎡ r ZL1 IR VR V ⎤<br />

⎣<br />

⋅ ⋅ − ⋅<br />

POL ⎦<br />

≥ 0 (3)<br />

In this expression, “Real” stands for “real part <strong>of</strong>” and “*”<br />

for “complex conjugate <strong>of</strong>.”<br />

For a forward fault, this expression becomes equivalent to<br />

the reach r being greater than a distance m computed in<br />

Equation 4 [3]:<br />

*<br />

Re al ⎡VR<br />

V ⎤<br />

⎣<br />

⋅<br />

POL ⎦<br />

r ⋅ ZL1<br />

≥ m⋅ ZL1<br />

=<br />

Re al ⎡<br />

⎣( 1∠θ<br />

) ⋅ I ⋅V<br />

*<br />

L1 R POL<br />

Where:<br />

∠θ L1 = phasor with unity magnitude and an angle equal<br />

to the positive-sequence line impedance angle<br />

m = per-unit distance to fault<br />

As Figure 1 shows, we use two sets <strong>of</strong> filtering systems to<br />

achieve speed in fault detection: one fast (data window <strong>of</strong> one<br />

half-cycle) and one conventional (data window <strong>of</strong> one-cycle)<br />

to compute the line voltage and current phasors. For each<br />

loop, we implement two mho-type detectors, each <strong>of</strong> which<br />

uses the fast or the conventional phasors. We achieve the final<br />

detection logic simply by “ORing” the outputs from the two<br />

mho-type detectors. We apply this principle for Zone 1<br />

detection and for zones (normally 2 and 3) used in<br />

communications-assisted schemes (POTT, DCB, etc.). In the<br />

case <strong>of</strong> Zone 1, the half-cycle detector has a reach less than<br />

the one-cycle detector reach. For Zones 2 and 3, the reach<br />

remains the same. We do not apply the principle in timedelayed<br />

stepped distance schemes because high-speed<br />

detection <strong>of</strong>fers no benefit with the delayed outputs <strong>of</strong> these<br />

schemes. For each zone and the corresponding six impedance<br />

loops, using the two sets <strong>of</strong> phasors to implement two<br />

calculations for the distance m leads necessarily to a faster<br />

result with the phasors derived from the half-cycle window<br />

than with phasors derived from the one-cycle window.<br />

Mho-type detectors are inherently directional but are not<br />

useful for fault type selection because multiple elements pick<br />

up for single-phase-to-ground faults. Furthermore, phasors<br />

derived from the half-cycle data window are less stable than<br />

the ones derived with a one-cycle data window. Therefore, in<br />

order to achieve security for high-speed tripping, we<br />

supplement the fast mho-type fault detectors with an<br />

algorithm called High-Speed Directional and Fault Type<br />

Selection, HSD-FTS [4][5]. This algorithm calculates three<br />

incremental torques (Equation 5) to determine the faulted<br />

phases and the fault direction.<br />

⎤<br />

⎦<br />

(4)<br />

( θ )<br />

Δ T = Re al ⎡ΔV ⎣<br />

⋅ 1∠ ⋅ΔI<br />

Δ T = Re al ⎡ΔV ⎣<br />

⋅ 1∠ ⋅ ΔI<br />

Δ T = Re al ⎡ΔV ⎣<br />

⋅ 1∠ ⋅ ΔI<br />

AB AB L1 AB<br />

( θ )<br />

BC BC L1 BC<br />

( θ )<br />

CA CA L1 CA<br />

Where:<br />

ΔV AB = two-cycle window incremental A-phase-to-Bphase<br />

voltage<br />

ΔI AB = two-cycle window incremental A-phase-to-Bphase<br />

current<br />

The relay uses the sign <strong>of</strong> the torques to establish direction<br />

and the relative values <strong>of</strong> these torques to select fault type.<br />

The high-speed directional element, HSD, provides a<br />

combined directional and fault type selection signal for a total<br />

<strong>of</strong> seven outputs in each direction; Table 1 shows the elements<br />

for forward faults (equivalent elements are derived for reverse<br />

faults). As shown in Table 1, we use a logical “AND” to<br />

combine each fast mho element with the corresponding HSD-<br />

FTS output. The result <strong>of</strong> this combination has two objectives:<br />

• Provide a single output when a fault detection occurs,<br />

allowing then reliable fault type selection for SPT<br />

applications.<br />

• Provide fast and reliable fault detection, allowing fast<br />

tripping times.<br />

TABLE 1<br />

COMBINING THE HSD-FTS AND FAST MHO ELEMENT LOGIC<br />

HSD-FTS<br />

Output<br />

Fault Type Selection<br />

*<br />

*<br />

⎤<br />

⎦<br />

⎤<br />

⎦<br />

⎤<br />

⎦<br />

*<br />

“ANDed” With<br />

HSD-AGF Forward A-phase-to-ground MHO-AGH<br />

HSD-BGF Forward B-phase-to-ground MHO-BGH<br />

HSD-CGF Forward C-phase-to-ground MHO-CGH<br />

HSD-ABF Forward phases A-B MHO-ABH<br />

HSD-BCF Forward phases B-C MHO-BCH<br />

HSD-CAF Forward phases C-A MHO-CAH<br />

HSD-ABCF Forward phases A-B-C MHO-ABH, MHO-<br />

BCH, MHO-CAH<br />

As shown in Figure 2 in the case <strong>of</strong> the A-phase-to-ground<br />

impedance loop, we simply “OR” signals from the one-cycle<br />

and half-cycle data-window fault detectors to get the final<br />

loop logic signal. We apply similar logic for the five other<br />

impedance loops. Note that we also supplement the<br />

conventional one-cycle mho element with a directional<br />

element [6] derived from the one-cycle data window and with<br />

a fault type selection logic [3].<br />

The result <strong>of</strong> this technique is that reliable high-speed fault<br />

detection occurs before conventional one-cycle fault detection.<br />

The amount <strong>of</strong> time advance varies and depends primarily<br />

upon the network (particularly the source-impedance ratio,<br />

SIR) and the fault location. Extensive testing revealed that<br />

fault detection occurs in practically all cases within a subcycle<br />

time frame.<br />

(5)<br />

42 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


One-Cycle Directional Element<br />

Figure 2<br />

A-Phase One-Cycle<br />

Fault Selection<br />

A-Phase-to-Ground<br />

One-Cycle Mho Element<br />

High-Speed Directional<br />

HSD-AGF<br />

A-Phase-to-Ground<br />

Half-Cycle Mho Element<br />

A-Phase-to-Ground<br />

Fault Detector-MAG<br />

A-Phase Impedance Loop Half- and One-Cycle Combination Logic<br />

B. High-Speed Distance Element Performance<br />

As an example, Figure 3 shows the two calculations <strong>of</strong> the<br />

distance m for a long line (in a 50 Hz system), Zone 2 A-<br />

phase-to-ground fault occurring at 0.1 s in a system with an<br />

SIR = 0.2. In this case, the half-cycle signal detects the fault at<br />

0.111 s and the one-cycle signal detects the fault at 0.123 s.<br />

The half-cycle element detects the fault 12 ms faster than the<br />

one-cycle element.<br />

m calculation (pu)<br />

2<br />

1.5<br />

1<br />

0.5<br />

Figure 3<br />

m half-cycle<br />

detection<br />

Zone 2 reach = r<br />

m one-cycle<br />

detection<br />

Fault<br />

inception<br />

0<br />

0.05 0.1 0.15 0.2<br />

Time (Seconds)<br />

m Calculations for a Zone 2 A-Phase-to-Ground Fault<br />

For the same example as the Zone 2 A-phase-to-ground<br />

fault shown in Figure 3, the timing <strong>of</strong> the different half-cycle<br />

and one-cycle signals is shown in Figure 4. The HSD-FTS<br />

signal (HSD-AGF in this case) occurs first (this signal has a<br />

duration <strong>of</strong> two cycles because it is based on superimposed<br />

quantities). After a few milliseconds, the half-cycle mho<br />

element (MHO-AGH) follows. More than a half-cycle later,<br />

we have the one-cycle mho (MHO-AGF) detection.<br />

Because the mho-detector signals are “ORed”, the time<br />

occurrence <strong>of</strong> the final fault detection MHO-AG corresponds<br />

to the fast detection MHO-AGH. In this case, we achieve an<br />

improvement, typical in practically all cases, <strong>of</strong> about one<br />

half-cycle. Two cycles after the fault inception, the detection<br />

depends only upon the one-cycle derived signals.<br />

HSD-AGF<br />

MHO-AGH<br />

MHO-AGF<br />

MHO-AG<br />

0 0.05 0.1 0.15 0.2<br />

Time (Seconds)<br />

Figure 4<br />

A-Phase-to-Ground Zone 2 Fault Signals<br />

III. SERIES COMPENSATION OVERREACHING<br />

PROBLEMS AND SOLUTION<br />

Series capacitors applied on transmission systems improve<br />

system stability and increase power transfer capability. The<br />

application <strong>of</strong> a series capacitor reduces the inductive<br />

reactance <strong>of</strong> the given transmission line, making the line<br />

appear electrically shorter. Although series capacitors may<br />

improve power system operation, using these capacitors<br />

results in a challenging problem for impedance-based line<br />

protection.<br />

Adding series capacitors on a transmission line causes<br />

subharmonic transients to occur following faults or switching<br />

<strong>of</strong> the series capacitor. These subharmonics can cause<br />

underreaching Zone 1 distance elements to overreach for<br />

external faults. There are other problems associated with<br />

subsynchronous resonance in generators. In this paper, we<br />

focus on problems associated with distance relays.<br />

All series capacitors come equipped with protective<br />

elements that reduce or eliminate over-voltages across the<br />

capacitor. The protection may be as simple as a spark gap set<br />

to flashover at a given voltage or as elaborate as metal-oxide<br />

varistors (MOV) using complex energy monitoring schemes.<br />

In any case, operation <strong>of</strong> the series capacitor protection<br />

elements can either remove the series capacitor completely or<br />

change capacitive reactance in a nonlinear fashion.<br />

The simplest series capacitor protection scheme removes<br />

the series capacitor when the series capacitor voltage exceeds<br />

a set threshold. The use <strong>of</strong> a spark gap protection scheme can<br />

simplify use <strong>of</strong> underreaching distance relays on seriescompensated<br />

lines. Firing <strong>of</strong> the spark gap for external faults<br />

may prevent Zone 1 overreach, so it is then possible to ignore<br />

the series capacitor. In most applications, however, the spark<br />

gap firing voltage threshold is high enough that the spark gap<br />

does not fire for external faults.<br />

MOVs present an interesting challenge because this type <strong>of</strong><br />

protection scheme does not fully remove the series capacitor.<br />

In fact, the capacitive reactance can be very nonlinear. We can<br />

use an iterative model [16] to approximate the effective<br />

reactance <strong>of</strong> the MOV-protected bank. However, this model<br />

does not provide insight about the transient response. Some<br />

MOV-protected capacitor banks include energy monitoring <strong>of</strong><br />

the MOVs. Bypassing <strong>of</strong> the capacitor bank occurs when the<br />

energy level exceeds a specified threshold. This can work to<br />

our advantage when we use impedance-based schemes.<br />

In either case, we need careful analysis and study before<br />

applying any impedance-based protection scheme on seriescompensated<br />

lines.<br />

A. Distance Relay Overreaching Problems<br />

The series connection <strong>of</strong> the capacitor, the transmission<br />

line, and the system source create a resonant RLC circuit. The<br />

natural frequency <strong>of</strong> the circuit is a function <strong>of</strong> the level <strong>of</strong><br />

compensation and the equivalent power system source. The<br />

level <strong>of</strong> compensation can change according to the switching<br />

in and out <strong>of</strong> series capacitor “segments.” The source<br />

impedance can change because <strong>of</strong> switching operations<br />

external to the protected line section.<br />

Transmission Line Protection System for Increasing <strong>Power</strong> System Requirements | 43


Figure 5 illustrates a transmission line with a 50 percent<br />

series-compensated system (e.g., the series capacitor reactance<br />

equals 50 percent <strong>of</strong> the positive-sequence line reactance). For<br />

the fault location shown, the underreaching distance element<br />

at the remote terminal (Station S) should not operate.<br />

Intuitively, we would expect that setting the reach to<br />

80 percent <strong>of</strong> the compensated impedance (Z L1 – jX C ) would<br />

be an appropriate reach setting. However, the series capacitor<br />

and the system inductance generate subharmonic oscillations<br />

that can cause severe overreach <strong>of</strong> the distance element.<br />

Figure 6 shows the impedance plane plot for the fault location<br />

shown in Figure 5 (where the series capacitor remains in<br />

service).<br />

S<br />

Relay<br />

Figure 5 System With Series Capacitors at One End With a Fault at the End<br />

<strong>of</strong> the Line<br />

R<br />

Ignoring variables such as mutual coupling and fault<br />

resistance, we can see that the calculated voltage equals the<br />

measured voltage. Determining the ratio <strong>of</strong> the measured<br />

voltage to the calculated voltage would result in unity or one.<br />

When the fault moves to the other side <strong>of</strong> the series<br />

capacitor (line-side), the measured voltage increases and the<br />

calculated voltage decreases. The measured voltage increases<br />

because the series capacitor is no longer between the relay and<br />

the fault, and the line appears to be electrically longer. The<br />

calculated voltage decreases because the calculated voltage<br />

always includes the series capacitor. The ratio <strong>of</strong> the measured<br />

voltage to the calculated voltage is greater than one for a fault<br />

at this location.<br />

As the fault nears the relay location, the measured voltage<br />

decreases and the calculated voltage increases. The ratio <strong>of</strong> the<br />

measured voltage to the calculated voltage approaches zero as<br />

the fault location nears the relay location. Figure 7 is a plot <strong>of</strong><br />

the measured voltage, V MEAS , calculated voltage, V CALC , and<br />

the ratio <strong>of</strong> the measured to the calculated voltage. Note that<br />

the scale is arbitrary; the purpose <strong>of</strong> this figure is to illustrate<br />

how the voltage magnitudes and the voltage ratio change with<br />

respect to fault location.<br />

Imaginary (Sec. )<br />

Relay<br />

Ratio<br />

1<br />

V MEAS<br />

V CALC<br />

Figure 6<br />

Apparent Impedance for a Fault at the End <strong>of</strong> the Line<br />

As we can see from the impedance plot, the apparent<br />

impedance magnitude decreases to a value as low as 2 ohms<br />

secondary. This value is close to half <strong>of</strong> the compensated line<br />

impedance! Note that in Figure 6 the capacitor is modeled<br />

with no overvoltage protection. This condition is common for<br />

most external faults, because the overvoltage protection is<br />

typically sized to accommodate external faults (e.g., the<br />

overvoltage protection does not operate for external faults).<br />

B. Scheme to Prevent Distance Element Overreach (Patent<br />

Pending)<br />

From Figure 5, we can calculate the voltage drop for a<br />

bolted A-phase-to-ground fault at the line end as follows:<br />

( ) ( )<br />

VCALC = ⎣⎡ IA + k0 ⋅ IG ⋅ ZL1 ⎤⎦ + ⎡⎣ IA ⋅ − jXC<br />

⎤⎦ (6)<br />

Where:<br />

I A = A-phase current at the relay location<br />

I G = residual or ground current at the relay location<br />

k 0 = zero-sequence compensation factor<br />

Z L1 = positive-sequence line impedance<br />

X C = capacitive reactance that the relay “sees”<br />

Figure 7 Measured and Calculated Voltages and Voltage Ratio for Faults<br />

Along the Line<br />

We can use the ratio we described earlier to supervise<br />

underreaching Zone 1 distance elements to prevent overreach<br />

for external faults. When the ratio <strong>of</strong> the measured to<br />

calculated voltage is less than a pre-defined threshold, the<br />

Zone 1 distance element can operate. Otherwise, the Zone 1<br />

element is blocked. The only additional information that the<br />

relay needs to calculate this ratio is the capacitive reactance<br />

that the relay “sees.” With the ratio supervision, you can set<br />

the Zone 1 reach based on the uncompensated line impedance.<br />

IV. DISTANCE ELEMENT POLARIZATION DURING<br />

POLE-OPEN CONDITIONS<br />

V POL is the polarizing voltage for calculating the distance to<br />

fault, m, as Equation 4 illustrates. The most popular polarizing<br />

quantity for distance protection is positive-sequence voltage<br />

with memory [2]. During pole-open conditions in applications<br />

with line-side potentials, eventual corruption <strong>of</strong> the polarizing<br />

quantity can occur if the input voltage to the memory circuit is<br />

corrupted. Invalid memory polarization may cause distance<br />

44 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


element misoperation. Shunt reactor switching generates<br />

damped oscillations with signals that have frequencies<br />

different from the actual system frequency. Let us look at an<br />

example <strong>of</strong> these signals and the logic that prevents the<br />

memory polarization from using unhealthy voltages.<br />

A. Shunt Reactor Switching<br />

Shunt reactors compensate the line charging currents and<br />

reduce overvoltages in long transmission lines. Figure 8 shows<br />

a 735 kV transmission line with shunt compensation at both<br />

ends <strong>of</strong> the line, 200 MVARs at each line end. The figure also<br />

shows line capacitance that generates 546 MVARs <strong>of</strong> reactive<br />

power. After the circuit breakers open at both line ends, the<br />

remaining circuit is basically an RLC circuit with a natural<br />

frequency <strong>of</strong> about 51.36 Hz; the circuit has stored energy in<br />

the reactor and in the line capacitance. Note that the circuit<br />

natural frequency is close to the nominal system frequency <strong>of</strong><br />

60 Hz. After the three poles <strong>of</strong> each line breaker open, the<br />

shunt reactors interact with the line capacitance and maintain<br />

line voltages for several cycles. The circuit applies these<br />

voltages to the potential transformers (PTs) or capacitive<br />

voltage transformers (CVTs). These voltages corrupt the<br />

distance protection polarization and frequency estimation.<br />

Figure 9 shows the A-phase voltage at the relay location after<br />

deenergization <strong>of</strong> the line (in Figure 8). There is no need to<br />

feed this distorted voltage to the distance protection<br />

polarization and frequency estimation algorithm, as we<br />

explain later.<br />

zeros to the memory filter after the relay detects voltage<br />

ringing or undervoltage conditions.<br />

Figure 10 The Relay Inputs Zeros to the Memory Filter When Input<br />

Voltages are Corrupted<br />

V. EFFICIENT FREQUENCY ESTIMATION DURING<br />

POLE-OPEN CONDITIONS<br />

Figure 11 shows a two-source system with line protection<br />

using line-side potentials in an SPT application. This figure<br />

also shows Breaker 1 (BK1) and Breaker 2 (BK2) A-phase<br />

open, indicating a single-pole open condition for both<br />

breakers. The distance element can misoperate during<br />

frequency excursions and single-pole open conditions if it<br />

does not track the system frequency correctly [2]. To prevent<br />

relay misoperations, the distance element needs a reliable<br />

frequency estimation algorithm for proper frequency tracking<br />

during breaker pole-open conditions.<br />

S<br />

L2<br />

R2<br />

R<br />

Relay<br />

L1<br />

C1<br />

C1<br />

L1<br />

C1 = 1341 nF<br />

L1 = 7.162 H<br />

L2 = 0.1736 H<br />

R2 = 2.33 Ω<br />

QC1 = 273 MVARS<br />

QL1 = 200 MVARS<br />

Figure 8 735 kV Transmission Line With Shunt Compensation at Both Ends<br />

<strong>of</strong> the Line<br />

Figure 9<br />

A-Phase Line Voltage After Line Deenergization<br />

We need to eliminate the positive voltage input (V 1 ) to the<br />

polarizing memory when the voltage is distorted. Figure 9<br />

illustrates a voltage ringing condition after the circuit breakers<br />

open at both line ends. The relay detects this ringing condition<br />

and eliminates the corrupted voltage from the memory filter<br />

input. On a phase-by-phase basis, the relay inputs zeros to the<br />

memory filter during pole-open conditions to prevent distance<br />

element misoperation. Figure 10 shows the logic for inputting<br />

Figure 11 Line Protection Using Line-Side Potential Transformers in a<br />

Single-Pole Tripping Application<br />

Traditionally, relays that calculate frequency must have<br />

circuitry that detects zero-crossings <strong>of</strong> the voltage signals to<br />

determine the signal period. The inverse <strong>of</strong> the signal period is<br />

the frequency. Normally, this circuitry monitors a single-phase<br />

voltage; the relay cannot measure frequency if the monitored<br />

phase is deenergized during the pole-open condition. Some<br />

numerical relays use zero-crossing detection [7] or rate-<strong>of</strong>change<br />

<strong>of</strong> angle algorithms to calculate frequency [8]. Some<br />

<strong>of</strong> these relays use positive-sequence voltage to include<br />

voltage information from the three phases [8]. These relays<br />

calculate frequency reliably as long as the voltages are present<br />

and healthy.<br />

Parallel line mutual coupling [12][13] or line resonance<br />

because <strong>of</strong> shunt reactor compensation can cause voltage<br />

distortion during the pole-open condition; this distortion<br />

introduces frequency estimation errors. This section describes<br />

logic to correctly estimate the system frequency during poleopen<br />

conditions and unhealthy voltage conditions.<br />

Transmission Line Protection System for Increasing <strong>Power</strong> System Requirements | 45


A. Frequency Estimation Sources <strong>of</strong> Error<br />

The following are power system conditions that the<br />

frequency estimation logic evaluates to prevent distance<br />

element misoperation. First, we show a frequency excursion<br />

during a pole-open condition. Second, we show the frequency<br />

<strong>of</strong> the voltage signal after a transmission line deenergization.<br />

1) Frequency Excursion During Pole-Open Condition<br />

During pole-open conditions, the power system is prone to<br />

instability. Particularly in weak source systems, the frequency<br />

changes when one phase is open because <strong>of</strong> the sudden<br />

increase in the impedance between the two line terminals. The<br />

distance relay needs to track this frequency to minimize<br />

distance element overreach. Figure 12 shows a 600 MVA<br />

generator connected to a 275 kV network through two<br />

transmission lines. The frequency changes, as illustrated in<br />

Figure 13, when Line 2 opens A-phase to clear an A phase-toground<br />

fault located 50 km from Station S while Line 1 is<br />

open. The frequency varies from 60.2 to 59.8 Hz during the<br />

pole-open condition.<br />

voltage in Figure 9; the frequency <strong>of</strong> the voltage signal<br />

oscillates between 44.6 and 54.4 Hz after the line<br />

deenergization (Figure 15). These corrupted voltages must be<br />

removed from the frequency estimation logic to prevent errors<br />

in the estimation, as we see later.<br />

S-275<br />

= 250 KM<br />

Line 1<br />

R-275<br />

Figure 15 Frequency Estimation Using A-Phase Voltage After Line<br />

De-energization<br />

600 MVA<br />

600 MVA<br />

22/275<br />

Relay<br />

ØA-G<br />

Line 2<br />

Figure 12 275 kV Network With 600 MVA Generator<br />

System<br />

Equivalent<br />

B. Frequency Estimation Logic (Patent Pending)<br />

Figure 16 shows the logic for frequency estimation that<br />

increases distance element reliability. The figure also shows<br />

alternate methods for determining the power system<br />

frequency, FREQ, for normal system operating conditions,<br />

pole-open conditions, and unhealthy voltage conditions. The<br />

relay uses FREQ to obtain the relay tracking frequency and to<br />

adapt the line protection to power system changes.<br />

Figure 13<br />

Frequency at the Terminal Close to the 600 MVA Generator<br />

If the relay in Figure 12 does not track the system<br />

frequency, the apparent impedance begins to appear as a fault<br />

condition, as shown in Figure 14. The relay must use voltage<br />

information from the unfaulted phases to track the system<br />

frequency and minimize distance element overreach.<br />

mBC Calculation (Pri. Ω)<br />

400<br />

300<br />

200<br />

100<br />

0<br />

Zone 2 Reach<br />

0.5 1 1.5 2 2.5<br />

Time (Seconds)<br />

Figure 14 Distance Element Apparent Impedance During Pole-Open<br />

Conditions<br />

2) Shunt Reactor Compensation<br />

We saw earlier the line voltages resulting from the<br />

deenergization <strong>of</strong> a line with shunt reactor compensation.<br />

Figure 15 shows the frequency estimation for the A-phase<br />

Σ<br />

Figure 16 Frequency Estimation Logic for Transmission Line Protection<br />

Applications<br />

The relay uses the V A , V B , and V C voltages to calculate the<br />

frequency, FREQ. Depending on breaker pole status and<br />

voltage source health, switches SW1, SW2, and SW3 select<br />

the V A , V B , and V C voltages, respectively, or zero. Switches<br />

SW1, SW2, and SW3 have two positions. Position 2 is the<br />

normal state; the switches are in this position if the<br />

corresponding pole is closed. The relay applies the source<br />

voltages when SW1, SW2, and SW3 are in Position 2. The<br />

relay selects Position 1 for SW1, SW2, and SW3 according to<br />

the following:<br />

• SW1 is in Position 1 when there is an A-phase poleopen,<br />

SPOA, condition; a three-pole open, 3PO,<br />

condition; or a loss-<strong>of</strong>-potential (LOP), condition.<br />

When SW1 is in Position 1, V1 is zero. Otherwise,<br />

V1 = V A .<br />

46 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


• SW2 is in Position 1 when there is a B-phase poleopen,<br />

SPOB, condition; a 3PO condition; or an LOP<br />

condition. When SW2 is in Position 1, V2 is zero.<br />

Otherwise, V2 = V B .<br />

• SW3 is in Position 1 when there is a C-phase poleopen,<br />

SPOC, condition; a 3PO condition; or an LOP<br />

condition. When SW3 is in Position 1, V3 is zero.<br />

Otherwise, V3 = V C .<br />

V1, V2, and V3 are the voltages after SW1, SW2, and SW3<br />

switch selection. The relay uses these voltages to calculate the<br />

composite signal, V α [14]. Equation 7 shows the V α<br />

calculation.<br />

⎛ V2 V3⎞<br />

Vα<br />

= ⎜ V1 − − ⎟ ⋅ K<br />

(7)<br />

⎝ 2 2 ⎠<br />

Where the constant K depends on the SW1, SW2, and SW3<br />

switch positions as shown in Table 2.<br />

TABLE 2<br />

K CONSTANT TO CALCULATE THE COMPOSITE VOLTAGE, V α,<br />

DEPENDS ON BREAKER POLE CONDITION<br />

Breaker Pole Condition SW1 SW2 SW3<br />

K<br />

A Pole B Pole C Pole Position Position Position<br />

Closed Closed Closed 2 2 2 2/3<br />

Open Closed Closed 1 2 2 2<br />

Closed Open Closed 2 1 2 2<br />

7<br />

Closed Closed Open 2 2 1 2<br />

Closed Open Open 2 1 1 1<br />

Open Closed Open 1 2 1 2<br />

Open Open Closed 1 1 2 2<br />

Open Open Open 1 1 1 2/3<br />

The digital band-pass filter, DBPF, extracts the<br />

fundamental component signal (60 Hz or 50 Hz) from the<br />

composite voltage, V α , to obtain V α_FUND . The relay uses the<br />

fundamental quantity, V α_FUND , to calculate the frequency,<br />

FREQ. As stated before, there are several methods for<br />

calculating frequency [7][8][9][10]. In this case, the frequency<br />

estimation algorithm uses the zero-crossing detection method.<br />

To prevent erroneous frequency estimation during power<br />

system transients (faults), the relay freezes the output <strong>of</strong> the<br />

frequency estimator if the relay detects a transient or a tripping<br />

condition. The output <strong>of</strong> the logic is the power system<br />

frequency, FREQ. The relay uses this frequency to determine<br />

the tracking frequency.<br />

The frequency estimation logic removes the voltage signal<br />

from the open phase to prevent using corrupted signals during<br />

pole-open conditions or LOP conditions. The composite<br />

signal, V α , allows the relay to combine information from three<br />

phases without additional signal manipulation with the “a”<br />

operator [15], as in the case <strong>of</strong> positive-sequence quantity.<br />

This composite signal does not require additional filtering for<br />

7<br />

implementation and has improved transient response. The<br />

relay uses the composite signal, V α , to track the system<br />

frequency during pole-open conditions. The constant K (see<br />

Table 2) modifies the gain <strong>of</strong> the composite signal, V α , to<br />

provide a constant signal amplitude to the frequency estimator<br />

for different line operating conditions.<br />

VI. SECONDARY ARC EXTINCTION DETECTION AND<br />

ADAPTIVE RECLOSING<br />

One <strong>of</strong> the challenges in SPT applications is to prevent the<br />

breaker pole from reclosing before the fault extinguishes.<br />

Reclosing under fault provides added power system damage<br />

and compromises system stability. There are several<br />

approaches to prevent these damaging conditions. One <strong>of</strong> the<br />

most common approaches is to add shunt and neutral reactors<br />

[17] to suppress the secondary arc present during the poleopen<br />

condition and have a dead time (open phase interval)<br />

long enough to allow for arc suppression and air de-ionization.<br />

Another approach is to increase the dead time and expect the<br />

secondary arc to be self-extinguished. None <strong>of</strong> these<br />

approaches verify expiration <strong>of</strong> the arc before the breaker pole<br />

recloses.<br />

As soon as the arc extinguishes, recovery voltage appears<br />

across the secondary arc path. This recovery voltage may<br />

initiate a restrike and create a reclose-onto-fault condition.<br />

Ideally, we want the breaker to close if and only if there is no<br />

fault on the line. Detection <strong>of</strong> the secondary arc extinction<br />

prevents reclosing under fault conditions and optimizes the<br />

dead time.<br />

A. The Secondary Arc Phenomenon<br />

In a three-phase circuit, there is electromagnetic and<br />

electrostatic coupling between the phase conductors [11]. A<br />

phase-to-ground fault results in formation <strong>of</strong> a primary arc<br />

between the faulted phase and ground. A protective system<br />

isolates the faulted phase from the power system in SPT<br />

applications; the other two healthy phases remain in service,<br />

thereby isolating the primary arc. Because two <strong>of</strong> the three<br />

phases are still at approximately nominal voltage and the<br />

faulted phase is coupled capacitively and inductively to the<br />

healthy phases, a secondary arc is maintained. The air is<br />

already ionized from the primary fault, so the two healthy<br />

phases need little current to maintain this secondary arc.<br />

Figure 17 Transmission Line Π Equivalent Circuit<br />

Figure 17 shows an equivalent Π circuit representing a<br />

transmission line section. If a trans-mission line lacks shunt<br />

reactors, we can simplify the equivalent circuit by considering<br />

only the electrostatic coupling <strong>of</strong> the phase conductors.<br />

Transmission Line Protection System for Increasing <strong>Power</strong> System Requirements | 47


Figure 18 represents a simplified, single, symmetrical, and<br />

fully transposed transmission line. The capacitances between<br />

phases are identical, i.e., C ab = C bc = C ca = C m and the<br />

capacitances to ground for each phase are identical, i.e., C ag =<br />

C bg = C cg = C g .<br />

Figure 18<br />

Equivalent Circuit Considering Only the Line Capacitances<br />

If this system now experiences an A-phase-to-ground fault,<br />

we can represent the system during the pole-open condition as<br />

illustrated in Figure 19. This figure also shows the vector<br />

diagram while the secondary arc is present.<br />

V<br />

REC<br />

= V ⋅<br />

LN<br />

C<br />

m<br />

( 2Cm<br />

+ Cg<br />

)<br />

(10)<br />

If the transmission line has shunt reactors, one must<br />

consider the electromagnetic coupling from the healthy<br />

phases; a portion <strong>of</strong> the secondary arc current becomes<br />

inductive. In this case, the secondary arc current is the vector<br />

sum <strong>of</strong> the electrostatic (I ARC-C ) and electromagnetic ( IARC-M )<br />

currents. Calculation <strong>of</strong> the electromagnetic current is a nontrivial<br />

task; you would need the assistance <strong>of</strong> a transient<br />

analysis s<strong>of</strong>tware package such as EMTP.<br />

B. Secondary Arc Extinction Detectors, SAEDs<br />

A secondary arc extinction detector can be added to<br />

supervise the closing signal to the breaker and prevent<br />

reclosing under fault. This supervision blocks the closing<br />

signal to the breaker when the secondary arc is present and<br />

minimizes the possibility <strong>of</strong> unsuccessful reclosures.<br />

Figure 20 shows the close supervision logic using secondary<br />

arc extinction detection (SAED) for close supervision. The<br />

SAED can minimize the dead time by initiating the reclose<br />

after the secondary arc extinguishes. The SAEDD time delay<br />

provides time for the air formerly occupied by the arc to<br />

regain dielectric capabilities.<br />

Secondary Arc<br />

Extinction Detection<br />

(SAED)<br />

SAEDD<br />

DO<br />

Reclosing<br />

Relay<br />

Supervision<br />

SAED<br />

Close<br />

Enable<br />

Figure 19<br />

Equivalent Circuit While the Secondary Arc is Present<br />

The magnitude <strong>of</strong> the secondary arc current resulting from<br />

the electrostatic coupling is a function <strong>of</strong> the line voltage and<br />

the line length. The line capacitance is a function <strong>of</strong> the<br />

distance between phase conductors and the height <strong>of</strong> these<br />

conductors above ground. If we represent the transmission line<br />

as a series <strong>of</strong> Π circuits, we can see that all these capacitances<br />

are effectively in parallel; the capacitance increases<br />

proportionally with the line length. From the equivalent circuit<br />

in Figure 19, we can calculate the secondary arc current,<br />

disregarding fault resistance, as follows:<br />

IARC− C = VDR−C<br />

⋅ j⋅<br />

ϖ ⋅ 2Cm<br />

(8)<br />

Where the voltage V DR-C for an A-phase-to-ground fault<br />

equals (V B + V C )/2 = V LN /2. We can therefore rewrite<br />

Equation 8 as follows:<br />

IARC− C = VLN<br />

⋅ j⋅<br />

ϖ ⋅ Cm<br />

(9)<br />

For a typical 400 kV line, I ARC-C has an approximate value<br />

<strong>of</strong> 0.1085 A/km, or 10.85 A/ 100 km. Once the secondary arc<br />

extinguishes, a voltage known as the recovery voltage appears<br />

across the ground capacitance, C g . The magnitude and/or rate<br />

<strong>of</strong> rise <strong>of</strong> the recovery voltage can initiate a restrike. We can<br />

calculate the recovery voltage as follows:<br />

Figure 20 Reclosing Relay Close Supervision Using Secondary Arc<br />

Extinction Detection<br />

Figure 21 shows the A-phase voltage after the line relay<br />

clears an A-phase-to-ground fault in an SPT application. After<br />

one pole opens, the voltage to ground, V ARC , exists in the<br />

faulted phase until the secondary arc extinguishes. When the<br />

secondary arc extinguishes, a phase-to-ground voltage, V A ,<br />

results from the mutual coupling between the sound phases<br />

and the faulted phase. This voltage provides information to<br />

detect secondary arc extinction [18].<br />

V C<br />

V A<br />

V B<br />

V Arc<br />

Figure 21 Line Voltages During the Pole-Open Condition<br />

48 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


Complete reclosing relay supervision requires three<br />

secondary arc extinction detectors, one per phase. These<br />

detectors measure the angle, φ, between the phase-to-ground<br />

voltage <strong>of</strong> the faulted phase (V γ ) and the sum <strong>of</strong> the sound<br />

phases phase-to-ground voltages (V Σ ). For -β ≤ φ ≤ β, and |V γ |<br />

> V thre , the detector determines the secondary arc extinction<br />

and asserts the SAED bit. β and V thre determine the region that<br />

detects the secondary arc extinction. Table 3 shows the V γ and<br />

V Σ voltages for A-, B-, and C-phase SAEDs.<br />

TABLE 3<br />

V γ AND V Σ VOLTAGES FOR A-PHASE, B-PHASE, AND C-PHASE SAEDS<br />

Detector V γ V Σ<br />

A-phase V A V B + V C<br />

C. SAED Performance in Transmission Lines With Shunt<br />

Compensation<br />

SAED is suitable for lines with or without shunt reactor<br />

compensation. We will look at the performance <strong>of</strong> the SAED<br />

for an A-phase-to-ground fault at the middle <strong>of</strong> the line in the<br />

power system shown in Figure 8. In this case, we assume that<br />

both breakers at the line ends open A phase poles to create an<br />

A-phase open condition. Figure 24 shows the A-phase voltage<br />

and the fault current through the arc at the fault location. The<br />

fault occurs at cycle eight (133 ms) and the poles open at<br />

cycle 14 (233 ms). The figure also illustrates the presence <strong>of</strong><br />

arc; primary arc exists before the breaker poles open, and<br />

secondary arc exists after the breakers clear the fault. The<br />

secondary arc extinguishes at cycle 23.5 (392 ms).<br />

B-phase V B V C + V A<br />

C-phase V C V A + V B<br />

The angle φ = 180° and V γ is outside <strong>of</strong> the SAED region<br />

when the line is energized under normal conditions as shown<br />

in Figure 22.<br />

During single-pole open conditions, after the secondary arc<br />

extinguishes, the angle φ = 0° and V γ is inside the SAED<br />

region (Figure 23). The A-phase detector asserts the SAED bit<br />

for this condition, allowing the reclosing sequence to<br />

continue.<br />

V C<br />

φ<br />

β<br />

V Σ<br />

= V B<br />

+ V C<br />

V γ<br />

= V A<br />

β<br />

V thre<br />

V B<br />

Figure 22 The Angle φ = 180° and V γ is Outside <strong>of</strong> the SAED Region for<br />

Normal Conditions<br />

V C<br />

V γ<br />

= V A<br />

V Σ<br />

= V B<br />

+ V C<br />

Figure 23 V γ is Inside the SAED Region After the Arc Extinguishes<br />

V B<br />

Figure 24 A-Phase Voltage, Arc Current, and Arc Presence Indication for an<br />

A-Phase-to-Ground Fault at the Middle <strong>of</strong> the Line<br />

The A-phase voltage undergoes three significant changes<br />

during fault and pole-open conditions:<br />

• Voltage reduction because <strong>of</strong> the fault presence. The<br />

voltage changes from normal to fault conditions.<br />

• Additional voltage reduction after the breakers clear<br />

the fault. Reduced voltage exists during the pole-open<br />

condition while the secondary arc is present.<br />

• Voltage increase during the pole-open condition. The<br />

voltage increases after the secondary arc extinguishes.<br />

Figure 25 shows the A-phase voltage changes in magnitude<br />

and angle, and the SAED boundaries. During the pole-open<br />

condition before extinction <strong>of</strong> the secondary arc, the voltage<br />

magnitude is less than the voltage threshold, V thre , and the<br />

voltage angle is outside the 2β angle limits. After the<br />

secondary arc extinguishes, the voltage magnitude exceeds the<br />

voltage threshold, V thre , and the voltage angle is within the 2β<br />

angle limits <strong>of</strong> the SAED characteristic. You can see the A-<br />

phase voltage better in the polar plot <strong>of</strong> Figure 26. V A = 1∠0°<br />

(in per unit) before the fault occurs; V A = 0.7∠0° during the<br />

fault condition; V A = 0.1∠254° while the secondary arc is<br />

present; and V A = 0.8∠191° after the secondary arc<br />

extinguishes. The A-phase voltage operating point enters the<br />

A-phase SAED region after the secondary arc extinguishes.<br />

Transmission Line Protection System for Increasing <strong>Power</strong> System Requirements | 49


phases without additional signal manipulation, as in<br />

the case <strong>of</strong> positive-sequence quantity.<br />

5. Secondary arc extinction detection prevents singlepole<br />

reclosing while the fault is present and optimizes<br />

the single pole-open interval, avoiding additional<br />

power system damage and minimizing system<br />

disturbance.<br />

Figure 25 A-Phase Voltage Magnitude and Angle for an A-Phase-to-Ground<br />

Fault With the SAED Boundaries<br />

150<br />

210<br />

120<br />

Secondary Arc<br />

Extinguishes<br />

240<br />

90<br />

180 0<br />

2 70<br />

1<br />

0 .5<br />

1.5<br />

Secondary Arc<br />

Present<br />

Fault<br />

60<br />

30 0<br />

Prefault<br />

Figure 26 A-Phase Voltage Phasor Enters the SAED Characteristic After the<br />

Secondary Arc Extinguishes<br />

VII. CONCLUSIONS<br />

1. The result <strong>of</strong> using the dual-filter scheme technique is<br />

reliable high-speed transmission line protection<br />

compared to conventional one-cycle only filtering<br />

schemes.<br />

2. The ratio <strong>of</strong> the measured voltage, V MEAS , to the<br />

calculated voltage, V CAL , in series compensation<br />

applications provides information to block Zone 1<br />

distance elements and prevent distance element<br />

overreach.<br />

3. The ability to remove the open phase voltage prevents<br />

using corrupted signals for distance element<br />

polarization and frequency tracking during pole-open<br />

conditions or loss-<strong>of</strong>-potential conditions.<br />

4. Using the composite signal, V α , allows relays to track<br />

system frequency during pole-open conditions. The<br />

composite signal combines information from the three<br />

30<br />

330<br />

VIII. REFERENCES<br />

[1] M. A. Adamiak, G. Alexander, and W. Premerlani “Advancements in<br />

Adaptive Algorithms for Secure High-Speed Protection,” 23rd Annual<br />

Western Protective Relay Conference, Spokane, Washington, October<br />

15–17, 1996.<br />

[2] D. Hou, A. Guzman, and J. Roberts, “Inovative Solutions Improve<br />

Transmission Line Protection,” 24th Western Protective Relay<br />

Conference, Spokane, WA, October 21–23, 1997.<br />

[3] E. O. Schweitzer III and J. Roberts, “Distance Relay Element Design,”<br />

19th Annual Western Protective Relay Conference, Spokane,<br />

Washington, October 20–22, 1992.<br />

[4] G. Benmouyal and J. Roberts, “Superimposed Quantities: Their True<br />

Nature and Their Application in Relays,” Proceedings <strong>of</strong> the 26th<br />

Annual Western Protective Relay Conference, Spokane, WA, October<br />

26–28, 1999.<br />

[5] G. Benmouyal and J. Mahzeredjian, “A Combined Directional and<br />

Faulted Type Selector Element Based on Incremental Quantities,” IEEE<br />

PES Summer Meeting 2001, Vancouver, Canada.<br />

[6] A. Guzman, J. Roberts, and D. Hou, “New Ground Directional Elements<br />

Operate Reliably for Changing System Conditions,” 23rd Annual<br />

Western Protective Relay Conference, Spokane, Washington, October<br />

15–17, 1996.<br />

[7] Instruction Manual for Digital Frequency Relay Model BE1-81 O/U,<br />

Basler Electric, Highland, Illinois. Publication: 9 1373 00 990. Revision:<br />

E. December 1992.<br />

[8] A. G. Phadke and J. S. Thorp, “Computer Relaying for <strong>Power</strong> Systems,”<br />

Research Studies Press Ltd. 1988.<br />

[9] A. G. Phadke, J. S. Thorp, and M. G. Adamiak, “A New Measurement<br />

Technique for Tracking Voltage Phasors, Local System Frequency, and<br />

Rate <strong>of</strong> Change <strong>of</strong> Frequency,” IEEE Transactions on <strong>Power</strong> Apparatus<br />

and Systems, Vol. PAS-102, No. 5, May 1983.<br />

[10] P. J. Moore, J. H. Allmeling, and A. T. Johns, “Frequency Relaying<br />

Based on Instantaneous Frequency Measurement,” IEEE/PES Winter<br />

Meeting, January 21–25, 1996, Baltimore, MD.<br />

[11] IEEE PSRC Working Group: J. Esztergalyos et al. “Single-Phase<br />

Tripping and Auto Reclosing <strong>of</strong> Transmission Lines IEEE Committee<br />

Report.”<br />

[12] M. J. Pickett, H. L. Manning, and H. N. V. Geem, “Near Resonant<br />

Coupling on EHV Circuits: I-Field Investigations,” IEEE Transactions<br />

on <strong>Power</strong> Systems, Vol. PAS-87, No. 2, February 1968.<br />

[13] M. H. Hesse and D. D. Wilson, “Near Resonant Coupling on EHV<br />

Circuits: II-Method <strong>of</strong> Analysis,” IEEE Transactions on <strong>Power</strong> Systems,<br />

Vol. PAS-87, No. 2, February 1968.<br />

[14] E. Clarke, “Circuit Analysis <strong>of</strong> A-C <strong>Power</strong> Systems,” General Electric,<br />

Schenectady, N.Y., 1950.<br />

[15] C. F. Wagner and R. D. Evans, “Symmetrical Components as Applied to<br />

the Analysis <strong>of</strong> Unbalanced Electrical Circuits,” McGraw-Hill, New<br />

York, 1933.<br />

[16] D. L. Goldsworthy, “A Linearized Model for MOV-Protected Series<br />

Capacitors,” IEEE Transactions on <strong>Power</strong> Systems, Vol. PWRS-2, No 4,<br />

Nov 1987, pp 953–958.<br />

[17] E. N. Kimbark, “Suppression <strong>of</strong> Ground-Fault Arcs on Singe-Pole<br />

Switched EHV Lines by Shunt Reactors,” IEEE Transactions on <strong>Power</strong><br />

Apparatus and Systems, Vol. 83, March 1964.<br />

[18] CFE-, LAPEM, UIE, ATN, “Secondary Arc Extinction Detection<br />

Project.”<br />

50 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


IX. BIOGRAPHIES<br />

Armando Guzmán, P.E. received his BSEE with honors from Guadalajara<br />

Autonomous University (UAG), Mexico, in 1979. He received a diploma in<br />

Fiber-Optics Engineering from Monterrey Institute <strong>of</strong> Technology and<br />

Advanced Studies (ITESM), Mexico, in 1990. He served as Regional<br />

Supervisor <strong>of</strong> the Protection Department in the Western Transmission Region<br />

<strong>of</strong> the Federal Electricity Commission (the electrical utility company <strong>of</strong><br />

Mexico) for 13 years. He lectured at UAG in power system protection. Since<br />

1993, he has been with Schweitzer Engineering Laboratories, Pullman,<br />

Washington, where he is presently a Research Engineer. He holds several<br />

patents in power system protection. He is a registered pr<strong>of</strong>essional engineer in<br />

Mexico, is a senior member <strong>of</strong> IEEE, and has authored and coauthored several<br />

technical papers.<br />

Joseph B. Mooney, P.E. received his B.Sc. in Electrical Engineering from<br />

Washington State University in 1985. He joined Pacific Gas and Electric<br />

Company upon graduation as a System Protection Engineer. In 1989, he left<br />

Pacific Gas and Electric and was employed by Bonneville <strong>Power</strong><br />

Administration as a System Protection Maintenance District Supervisor. In<br />

1991, he left Bonneville <strong>Power</strong> Administration and joined Schweitzer<br />

Engineering Laboratories as an Application Engineer. Shortly after starting<br />

with <strong>SEL</strong>, he was promoted to Application Engineering Manager. In 1999, he<br />

became Manager <strong>of</strong> the <strong>Power</strong> Engineering Group <strong>of</strong> the Research and<br />

Development department at Schweitzer Engineering Laboratories. He is a<br />

registered pr<strong>of</strong>essional engineer in the states <strong>of</strong> California and Washington.<br />

Gabriel Benmouyal, P.E. received his B.A.Sc. in Electrical Engineering and<br />

his M.A.Sc. in Control Engineering from Ecole Polytechnique, Université de<br />

Montréal, Canada in 1968 and 1970, respectively. In 1969, he joined Hydro-<br />

Québec as an Instrumentation And Control Specialist. He worked on different<br />

projects in the field <strong>of</strong> substation control systems and dispatching centers. In<br />

1978, he joined IREQ, where his main field <strong>of</strong> activity was the application <strong>of</strong><br />

microprocessors and digital techniques to substation and generating-station<br />

control and protection systems. In 1997, he joined Schweitzer Engineering<br />

Laboratories in the position <strong>of</strong> Research Engineer. He is a registered<br />

pr<strong>of</strong>essional engineer in the Province <strong>of</strong> Québec, is an IEEE member, and has<br />

served on the <strong>Power</strong> System Relaying Committee since May 1989.<br />

Normann Fischer joined Eskom as a Protection Technician in 1984. He<br />

received a Higher Diploma in Technology, with honors, from the<br />

Witwatersrand Technikon, Johannesburg, in 1988 and a B.Sc. in Electrical<br />

Engineering, with honors, from the University <strong>of</strong> Cape Town in 1993. He was<br />

a Senior Design Engineer in Eskom’s Protection Design Department for three<br />

years, then joined IST Energy as a Senior Design Engineer in 1996. In 1999,<br />

he joined Schweitzer Engineering Laboratories as a <strong>Power</strong> Engineer in the<br />

Research and Development Division. He was a registered pr<strong>of</strong>essional<br />

engineer in South Africa and a member <strong>of</strong> the South Africa Institute <strong>of</strong><br />

Electrical Engineers.<br />

Copyright © <strong>SEL</strong> 2001<br />

(All rights reserved)<br />

Printed in USA<br />

20020401<br />

Transmission Line Protection System for Increasing <strong>Power</strong> System Requirements | 51


1<br />

Lessons Learned Analyzing Transmission Faults<br />

David Costello, Schweitzer Engineering Laboratories, Inc.<br />

Abstract—Transmission line relays record interesting power<br />

system phenomena and misoperations due to a variety <strong>of</strong><br />

problems. By analyzing event records, some common setting<br />

mistakes and misapplications have become evident. Setting and<br />

testing recommendations can be made to avoid these problems.<br />

In the interest <strong>of</strong> reducing transmission line relay misoperations,<br />

this technical paper shares practical lessons learned through<br />

experience analyzing transmission line relay event reports.<br />

I. FAULT LOCATION ERROR<br />

In December 2006, a single-line-to-ground fault occurred<br />

on a two-terminal 69 kV transmission line. The distance relay<br />

at the local terminal produced a fault location estimate, trip<br />

targets, and an event report. The actual fault location was<br />

found to be 3.6 miles from the local terminal, but the relay<br />

reported 7.31 miles. In addition, the relay targets included<br />

Zone 1 and time indication, which is contradictory because<br />

Zone 1 elements were set with no intentional time delay.<br />

Event data from the local relay are shown in Fig. 1.<br />

Ground current (3I0) reached a maximum magnitude <strong>of</strong> 3600<br />

amperes. The actual fault location was validated using this<br />

information and an ASPEN OneLiner power system model. In<br />

OneLiner, a line-to-ground fault was simulated and slid along<br />

the line until the 3I0 current matched the maximum value<br />

from the event data. The estimated location <strong>of</strong> this simulated<br />

fault matched the actual fault location from the field.<br />

IA IB IC IGMag<br />

VA(kV) VB(kV) VC(kV) VB(kV)Mag<br />

Digitals<br />

2500<br />

0<br />

-2500<br />

50<br />

25<br />

0<br />

-25<br />

-50<br />

Fig. 1.<br />

52A<br />

TRIP<br />

67G1<br />

51G<br />

IA IB IC IGMag<br />

VA(kV) VB(kV) VC(kV) VB(kV)Mag<br />

6.75 cycles<br />

0 1 2 3 4 5 6 7 8 9 10<br />

Cycles<br />

Event Data From 69 kV Transmission Line Relay<br />

The event data capture was triggered at Cycle 2 by an<br />

inverse-time ground overcurrent element (51G). The 3I0<br />

current is small (200 amperes 3I0) at the fault inception<br />

because <strong>of</strong> arcing or fault impedance. In the fifth cycle, the<br />

ground fault evolves into a bolted fault (3600 amperes 3I0).<br />

The instantaneous ground overcurrent element (67G1) asserts<br />

and calls for a trip, and the breaker opens 3 cycles later.<br />

An event record typically contains prefault, fault, and<br />

postfault data. After the event record is stored, the relay<br />

executes the fault location algorithm. This algorithm must first<br />

extract voltage and current phasor information from a point<br />

within the event data. What is the best point in the event data<br />

to choose<br />

A typical fault location algorithm will determine the<br />

contiguous fault data, which are defined by the window <strong>of</strong><br />

time in which the element that triggered the event remains<br />

asserted. In Fig. 1, this window <strong>of</strong> time is 6.75 cycles long.<br />

The algorithm assumes that the contiguous data midpoint is a<br />

reasonably stable point at which to estimate the fault location.<br />

In this case, the fault data midpoint is just before the ground<br />

current magnitude increased to a larger value. The dramatic<br />

evolution <strong>of</strong> the fault and its unlikely timing led to the fault<br />

location error.<br />

An <strong>of</strong>fline analysis <strong>of</strong> this event is required so one can<br />

manipulate where in the event data the fault algorithm extracts<br />

voltages and currents. In Fig. 2, a relay fault location algorithm<br />

is plotted using a MathCad simulation. At Cycle 6.75,<br />

the midpoint <strong>of</strong> the bolted fault data, the fault location<br />

estimate matches the actual fault location.<br />

m<br />

15<br />

10<br />

5<br />

Fig. 2.<br />

0<br />

0 1 2 3 4 5 6 7 8 9 10<br />

Cycles<br />

Fault Location MathCad Simulation<br />

We must now explain why a trip caused by an<br />

instantaneous ground element produced a time target. How<br />

fast is an instantaneous trip versus a time-delayed trip In this<br />

particular relay, the time target asserts when elements selected<br />

by the user have been asserted for longer than 3 cycles before<br />

a trip occurs. If the trip occurs before 3 cycles, we label the<br />

trip as instantaneous [1]. The same element that triggered the<br />

event data capture, 51G, was also set to signify that a fault is<br />

present. The dramatic evolution <strong>of</strong> the fault and its unlikely<br />

timing led to the confusing targets as well.<br />

Single-ended fault location estimates from relays are<br />

accurate in most cases. Offline tools, such as MathCad<br />

worksheets, can improve fault location accuracy for challenging<br />

cases. These tools <strong>of</strong>fer the ability to vary the location at<br />

which the fault location estimate is calculated and can include<br />

data from both ends <strong>of</strong> the transmission line for even better<br />

results [2].<br />

52 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


2<br />

II. COMMON ZONE TIMING<br />

In June 2006, a fault occurred just past the remote terminal<br />

on a two-terminal 138 kV transmission line. The directional<br />

comparison blocking (DCB) scheme was disabled at the time<br />

<strong>of</strong> the fault. Because the fault was within the Zone 2 reach, we<br />

expected a trip within the Zone 2 delay <strong>of</strong> 25 cycles.<br />

Event data from the local relay are shown in Fig. 3. The<br />

event data capture was triggered at Cycle 4 by the Zone 2<br />

phase distance element (M2P). The relay tripped and reported<br />

Zone 2 and time targets 0.5 cycles later.<br />

IA IB IC<br />

IGMag<br />

VA(kV) VB(kV) VC(kV)<br />

Digitals<br />

2000<br />

0<br />

-2000<br />

750<br />

500<br />

250<br />

0<br />

100<br />

0<br />

-100<br />

Fig. 3.<br />

IN105<br />

IN106<br />

TRIP<br />

52A<br />

Z2T<br />

Z2GT<br />

Z2G<br />

M2PT<br />

M2P<br />

IA IB IC IGMag VA(kV) VB(kV) VC(kV)<br />

.5 cycles<br />

0 1 2 3 4 5 6 7 8<br />

Cycles<br />

Event Data From 138 kV Transmission Line Relay<br />

Unlike the ground fault in Section I, which evolved into a<br />

larger magnitude fault <strong>of</strong> the same type, this fault evolved<br />

from a single-line-to-ground fault to a phase-phase-ground<br />

fault.<br />

Distance relays support two philosophies <strong>of</strong> zone timing:<br />

independent or common timing (see Fig. 4). For the independent<br />

timing mode, the phase and ground distance elements for<br />

each zone initiate independent timers. For the common mode,<br />

the phase and ground distance elements for each zone drive a<br />

common timer. The common zone timer is suspended for<br />

1 cycle if the timer input drops out. This feature prevents timer<br />

reset when a fault evolves [3].<br />

Suspend Timing<br />

during the fault transition period, and the common zone timer<br />

was suspended. The M2P element then asserted, the common<br />

zone timer resumed timing, and a trip was issued 0.5 cycles<br />

later. In this example, the benefit <strong>of</strong> common zone timing was<br />

a total relay response time <strong>of</strong> 25.5 cycles instead <strong>of</strong> the<br />

expected 50.5 cycles if we had chosen to use independent<br />

timing.<br />

In this event, the fault location estimate from the relay was<br />

accurate despite the evolving fault data. The contiguous fault<br />

data were roughly from Cycle 4 to 7, which was the window<br />

<strong>of</strong> time the triggering element M2P was asserted. During that<br />

time, the fault had already evolved, and the fault data were<br />

stable.<br />

In Fig. 5, the fault location estimate is plotted for the two<br />

different fault types experienced. Both estimates produce<br />

roughly the same fault location when data are stable. But from<br />

Cycle 2.5 to 3.5 you can see where both estimates veer wildly<br />

from the actual location because <strong>of</strong> the fault type evolution.<br />

The reported location from the relay would have been in error<br />

if it had selected data from that time.<br />

m<br />

m<br />

3.5<br />

2<br />

Single-Line-to-Ground Faults (BG)<br />

0<br />

0 1 2 3 4 5 6 7 8<br />

Cycles<br />

2<br />

Phase-to-Phase Faults (BC)<br />

0<br />

0 1 2 3 4 5 6 7 8<br />

Cycles<br />

Fig. 5. MathCad Fault Location Estimates for BG and BC Fault Types<br />

Event data can be replayed as IEEE COMTRADE files<br />

through test equipment into relays. Unique or challenging<br />

fault cases should be archived as IEEE COMTRADE files and<br />

used to test new relays, challenge standard schemes, and<br />

understand relay responses.<br />

M2P<br />

Z2G<br />

Fig. 4.<br />

Z2D<br />

Z2PD<br />

Z2GD<br />

Independent and Common Zone Timing Schemes<br />

0<br />

0<br />

0<br />

Z2T<br />

M2PT<br />

Z2GT<br />

We can deduce from the data in Fig. 3 that the initial fault<br />

had been present, and the Zone 2 ground distance element<br />

(Z2G) had been timing for 24.5 cycles before the fault<br />

evolved. All distance elements dropped out for 0.5 cycles<br />

III. RECLOSE FAILURE<br />

In May 2006, a 34.5 kV line feeding a hospital failed to<br />

reclose for a line fault. The relay was set for two reclose<br />

attempts.<br />

Event data from the relay are shown in Fig. 6. The event<br />

started as an A-phase-to-ground (AG) fault and evolved to a<br />

B-phase-to-ground (BG) fault. This was most likely caused by<br />

a slapping conductor or tree contact. At fault inception, an instantaneous<br />

ground overcurrent element (67G1) asserts and<br />

produces a trip output. At this same time, the relay enters the<br />

reclosing cycle state (79CY), indicating that reclosing openinterval<br />

timing has been initiated. However, we can see that as<br />

the AG fault progresses to a BG fault, the 67G1 element drops<br />

out for 0.75 cycles before reasserting. When the 67G1 element<br />

Lessons Learned Analyzing Transmission Faults | 53


3<br />

reasserts, the relay advances to lockout (79LO). The root<br />

cause can be found in the relay operation and user settings [1].<br />

A programmable logic equation (79RI) defines the<br />

conditions that initiate reclose. 79RI is set equal to the rising<br />

edge <strong>of</strong> 67G1 plus a number <strong>of</strong> other conditions logically<br />

ORed together. Once 79RI asserts, the relay enters its cycle<br />

state (79CY) and begins to time to reclose. However, the<br />

reclose timer is interrupted when 79RI reasserts. The relay<br />

immediately assumes something has gone awry, such as a<br />

flashover during a breaker opening, and the relay immediately<br />

goes to lockout (79LO) to prevent further trouble.<br />

just after the fault transitioned from AG to BG; therefore, we<br />

are suspicious that the relay used corrupt data for its fault<br />

location estimate.<br />

A fault location algorithm MathCad simulation is shown in<br />

Fig. 7. We can see the fault location error during the fault type<br />

change, but before and after that transition the location estimate<br />

is consistent.<br />

m<br />

12<br />

9<br />

6<br />

Single-Line-to-Ground Faults (BG)<br />

IA IB IC<br />

2500<br />

0<br />

-2500<br />

IA IB IC IGMag VA(kV) VB(kV) VC(kV)<br />

5.25 cycles<br />

3<br />

0<br />

7 7.5 8 8.5 9 9.5 10 10.5 11 11.5 12 12.5 13<br />

Cycles<br />

IGMag<br />

VA(kV) VB(kV) VC(kV)<br />

Digitals<br />

2000<br />

1000<br />

0<br />

25<br />

0<br />

-25<br />

FSB<br />

FSA<br />

52A<br />

79LO<br />

79CY<br />

79RS<br />

67G1T<br />

51G<br />

TRIP<br />

5.0 7.5 10.0 12.5 15.0 17.5<br />

Cycles<br />

Fig. 6. Reclosing Interrupted by Evolving Fault<br />

To prevent this from happening in the future, reclosing<br />

must not be reinitiated before reclose open-interval timing is<br />

complete and the breaker has been reclosed. Use the trip<br />

element to initiate reclose. Relays have a minimum trip<br />

duration time, so trip outputs remain asserted throughout<br />

evolving faults.<br />

It is also common to stall open-interval timing while the<br />

trip condition is still present. Additionally, there are likely<br />

conditions included in the trip logic for which we do not want<br />

to reclose. These conditions can be specified in a programmable<br />

drive-to-lockout equation (79DTL). Such conditions may<br />

include remote or manual OPEN commands, relay trips for<br />

three-phase faults, or time-delayed trips.<br />

An interesting breaker problem is also evident by the data<br />

in Fig. 6. Notice that the breaker status contact (52A) changes<br />

state shortly after the main breaker interrupts current. This is<br />

consistent with breaker operations in the previous two<br />

examples. However, we see that while the main breaker<br />

contacts remain open, the breaker auxiliary status contact<br />

bounces closed after 0.75 cycles and remains closed for 3.5<br />

cycles before opening again. Bouncy auxiliary contacts can<br />

cause reclosing failures. If 52A asserts before the relay calls<br />

for a close, relay logic assumes a manual or remote CLOSE<br />

command asserted. The auxiliary contacts <strong>of</strong> this breaker need<br />

maintenance to prevent future problems.<br />

The relay reported a fault location <strong>of</strong> 3.22 miles and a BG<br />

fault type. The contiguous fault data are roughly from Cycle 8<br />

to 13, which was the window <strong>of</strong> time the triggering element<br />

51G was asserted. The midpoint <strong>of</strong> that data corresponds to<br />

Fig. 7. MathCad Fault Location Estimates for AG and BG Fault Types<br />

The relay fault location algorithm selected data just prior to<br />

Cycle 10.5, where the B-phase current is still increasing to its<br />

maximum value. Therefore, its estimate is slightly further<br />

away from the local terminal because <strong>of</strong> the lower current<br />

magnitude used. The actual fault location is closer to 3.0<br />

miles.<br />

IV. COMMUNICATIONS PROBLEM<br />

In June 2007, a fault occurred beyond the remote terminal<br />

<strong>of</strong> a 115 kV line protected by a DCB scheme. DCB schemes<br />

use a high-speed on/<strong>of</strong>f carrier signal to block high-speed<br />

tripping at the remote terminal for out-<strong>of</strong>-section faults.<br />

Therefore, we expected the remote terminal to send a blocking<br />

signal to the local relay to prevent it from tripping at a high<br />

speed.<br />

The event data recorded from the local relay are shown in<br />

Fig. 8. The received blocking signal was wired to a programmable<br />

input on the relay (IN5) and appears to have been<br />

asserted at the time <strong>of</strong> the trip.<br />

IA IB IC<br />

IRMag<br />

VA VB VC<br />

Digitals<br />

100<br />

0<br />

-100<br />

75<br />

50<br />

25<br />

0<br />

50<br />

0<br />

-50<br />

Fig. 8.<br />

IA IB IC IRMag VA VB VC<br />

OUT 1&2<br />

B<br />

IN 5&6 5 5<br />

67N 2<br />

-1 0 1 2 3 4 5 6 7 8 9 10 11<br />

Cycles<br />

Four Samples-per-Cycle Data From a DCB Scheme Trip<br />

The relay is set so that overreaching phase and ground<br />

elements are allowed to trip after a one-cycle carrier<br />

coordination delay if no block is received. However, the relay<br />

tripped by overreaching ground directional overcurrent ele-<br />

54 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


4<br />

ment 67N2 after a 1-cycle carrier coordination delay.<br />

Apparently, a blocking signal was being received.<br />

Higher resolution data from the same relay for the same<br />

event are shown in Fig. 9. The better resolution illuminates a<br />

0.25-cycle carrier hole that allowed the local relay to trip.<br />

IA IB IC<br />

IRMag<br />

VA VB VC<br />

Digitals<br />

100<br />

0<br />

-100<br />

75<br />

50<br />

25<br />

0<br />

50<br />

0<br />

-50<br />

Fig. 9.<br />

BTX<br />

OUT1<br />

IN5<br />

67N2T<br />

67N2<br />

IA IB IC IRMag VA VB VC<br />

.25 cycles<br />

-1 0 1 2 3 4 5 6 7 8 9 10 11<br />

Cycles<br />

Sixteen Samples-per-Cycle Raw Data From a DCB Scheme Trip<br />

The received blocking input must remain asserted to block<br />

the forward-looking elements after the coordination timer<br />

expires. If the blocking signal drops out momentarily because<br />

<strong>of</strong> noise or other channel-related problems, the local relay will<br />

trip for out-<strong>of</strong>-section faults.<br />

Relays include an extension timer to delay the control<br />

input dropout assigned to receive the blocking signal. The<br />

output <strong>of</strong> this timer (BTX) ensures unwanted tripping does not<br />

occur during momentary lapses <strong>of</strong> the blocking signal (carrier<br />

holes). A typical setting is 0.5 to 1.5 cycles.<br />

One drawback <strong>of</strong> using the block trip (BT) extension timer<br />

is that tripping would be delayed anytime the BT input is<br />

asserted. This can cause unnecessary delays for internal faults<br />

in applications using a nondirectional carrier start. That is, the<br />

BT input asserts momentarily when nondirectional elements<br />

assert then deasserts when the forward elements detect the<br />

fault. BTX would remain asserted and delay tripping until the<br />

extension timer expired.<br />

To avoid this delay, use programmable logic in the relay to<br />

extend the blocking signal (see Fig. 10).<br />

IN10x<br />

1.0 cyc<br />

Set pickup to longer than<br />

remote carrier assertion<br />

Fig. 10.<br />

1 cyc<br />

Set dropout to longer than<br />

expected carrier hole time<br />

BT<br />

0 cyc<br />

BTXD<br />

Set BTXD to zero (0)<br />

BT Extension for Nondirectional Carrier Start Applications<br />

In this logic, we extend the BT only if the carrier has been<br />

initially picked up longer than the remote carrier assertion.<br />

The typical BT extension timer setting can then be set equal to<br />

0 cycles [4].<br />

BTX<br />

V. SENSITIVITY MISCOORDINATION<br />

In January 2001, two parallel 138 kV lines connecting a<br />

utility to a petrochemical plant tripped for an out-<strong>of</strong>-section<br />

fault. The simultaneous outages left the plant’s local<br />

generation and load electrically isolated from the utility. Both<br />

lines use identical permissive overreaching transfer trip<br />

(POTT) schemes with identical relay types at local and remote<br />

terminals.<br />

A simplified system one-line diagram with fault location is<br />

shown in Fig. 11. The tie line that is in parallel with Line 3–4<br />

and other infeeds to the petrochemical plant and utility bus are<br />

not shown. The utility is designated 1, and the petrochemical<br />

plant is designated 2. The actual fault location was behind the<br />

utility terminal (reverse to Terminal 1 and forward to Terminal<br />

2).<br />

2 – Petrochem 1 – Utility<br />

Fig. 11.<br />

1 2 3 4 5 6<br />

Zone 3 Zone 2<br />

Zone 2 Zone 3<br />

Simplified One-line Diagram <strong>of</strong> System With Relay Sensitivities<br />

Event data from both ends <strong>of</strong> one tie line are shown<br />

simultaneously in Fig. 12. The event data match the designations<br />

above, i.e., the utility data are designated 1, and the<br />

petrochemical plant data are designated 2.<br />

1_IA 1_IRMag<br />

2_IA 2_IRMag<br />

2_VA 1_VA<br />

Digitals<br />

1000<br />

500<br />

0<br />

-500<br />

1000<br />

500<br />

0<br />

-500<br />

50<br />

0<br />

-50<br />

Fig. 12.<br />

1_IA 1_IRMag 2_IA 2_IRMag 2_VA 1_VA<br />

.054167 sec<br />

2_OUT 3&4 3 3<br />

B<br />

2_OUT 1&2<br />

2_67N 1<br />

2_32 Q<br />

1_OUT 3&4 3 3<br />

1_OUT 1&2 2 2<br />

1_IN 1&2 1<br />

1_67N<br />

1_32 q<br />

34.725 34.750 34.775 34.800 34.825 34.850 34.875 34.900 34.925<br />

Event Time (Sec) 05:05:34.856396<br />

Utility and Plant Relay Data for Fault<br />

Terminal 2 responds to the forward AG fault by asserting<br />

its forward directional element (2_32Q) and its overreaching<br />

forward directional ground overcurrent element (2_67N2).<br />

The digital trace in Fig. 12 is identified with a 2_67N1, which<br />

is correct. Settings for the level one (67N1) and level two<br />

(67N2) ground elements are set to the same value, and the<br />

relay reports the assertion <strong>of</strong> the nearest zone. Terminal 2 keys<br />

a permission-to-trip (PTT) signal to the remote utility relay<br />

(2_OUT3).<br />

Lessons Learned Analyzing Transmission Faults | 55


5<br />

MTCS<br />

Z3RB<br />

67N3<br />

67Q3<br />

Z3G<br />

M3P<br />

3P0<br />

OR 1<br />

AND 2<br />

0<br />

Z3RBD<br />

(Setting)<br />

AND 5<br />

OR 4<br />

Key<br />

EBLKD = OFF<br />

(Setting)<br />

ETDPU = OFF<br />

(Setting)<br />

PT<br />

(Input)<br />

AND 1<br />

0<br />

EBLKD<br />

(Setting)<br />

AND 3<br />

(Setting)<br />

ETDPU<br />

0<br />

0<br />

EDURD<br />

(Setting)<br />

AND 6<br />

EKey<br />

Fig. 13.<br />

Partial Relay POTT Scheme Logic Diagram<br />

Terminal 1 correctly sees the fault as reverse according to<br />

its directional element (1_32q). The PTT signal sent by<br />

Terminal 2 is received 10 ms later by Terminal 1 (1_OUT2).<br />

OUT2 is programmed to follow the local PT logic bit (PTT<br />

received).<br />

A portion <strong>of</strong> the relay POTT scheme logic diagram is<br />

shown in Fig. 13. The relays include logic that permits rapid<br />

clearing <strong>of</strong> end-<strong>of</strong>-line faults when one line terminal is open or<br />

has a very weak source. This is referred to as “echo” logic [5].<br />

The open breaker or weak source terminal is allowed to echo<br />

the received permission signal as long as two main conditions<br />

are met:<br />

1. A reverse fault must not have been detected by the<br />

reverse-looking elements (67N3, 67Q3, Z3G, M3P).<br />

2. The PTT input must be received for a settable length <strong>of</strong><br />

time.<br />

At Terminal 1, no reverse-looking elements are asserted.<br />

After 2 cycles, Terminal 1 echos the received PTT signal back<br />

to Terminal 2 as a four-cycle pulse. Why did Terminal 1 not<br />

have a reverse-looking element asserted The Zone 3 reverse<br />

ground distance element (Z3G) was desensitized by infeed<br />

from other lines at the utility bus. Reverse ground overcurrent<br />

(67N3) was not enabled at this relay.<br />

Meanwhile, the fault is still visible and on the system, and<br />

3 cycles after Terminal 2 originally sent PTT, it receives its<br />

own echoed signal back. This is received by Terminal 2 as<br />

PTT, and a local high-speed trip is asserted (2_OUT1&2).<br />

In Fig. 11, note the fault between Breakers 5 and 6; the<br />

67N2 ground overcurrent element at Breaker 3 is sensitive<br />

enough to see this fault, and Breaker 3 sends a PTT signal to<br />

the relay at Breaker 4. At Breaker 4, no reverse 67N3 ground<br />

overcurrent element is enabled. The ground distance element,<br />

Z3G, at Breaker 4 is not sensitive enough to see this reverse<br />

fault, so the Breaker 4 echo logic returns the received PTT<br />

signal. The result is that Breaker 3 trips for an out-<strong>of</strong>-section<br />

fault. The parallel line tripped in identical fashion.<br />

The root cause <strong>of</strong> this misoperation was dissimilar<br />

sensitivities in the local and remote relays. This difference was<br />

not because <strong>of</strong> differences in design, make, or model. The<br />

difference was because <strong>of</strong> different settings and elements<br />

enabled at each end <strong>of</strong> the line by different engineers. This<br />

was not discovered in the engineering review or by<br />

commissioning tests.<br />

Engineers and technicians continue to perform detailed<br />

element tests during commissioning. Testing the overall<br />

protection system, including installed settings, is more<br />

important. End-to-end testing in the lab before installation or<br />

end-to-end satellite-synchronized testing during commissioning<br />

may have found this setting problem. In order to do so, the<br />

test voltages and currents for each line terminal would have to<br />

be provided from a model <strong>of</strong> the power system for faults at the<br />

end <strong>of</strong> element reaches.<br />

When installing a POTT scheme, the need for echo logic<br />

should be considered carefully. Do you even have the<br />

possibility <strong>of</strong> a weak infeed condition If you enable echo<br />

logic, it is critical that reverse-looking Zone 3 elements are<br />

enabled so that they can supervise the echo decision. Local<br />

and remote line relay sensitivities should be coordinated, i.e.,<br />

if ground overcurrent elements are enabled at the local end<br />

(forward 67N2, reverse 67N3), they should be enabled at the<br />

remote end as well [6].<br />

In this event, we also notice that once an echo key<br />

condition was established, it continued indefinitely until<br />

stopped by personnel intervention. In Fig. 13, if echo pulse<br />

duration EDURD is set longer than pickup ETDPU, the local<br />

and remote relays will establish a permanent echo key<br />

condition. To prevent this from occurring, set ETDPU to 2<br />

cycles and EDURD to 1 cycle. Limiting the echo duration to<br />

less than the echo time-delay pickup ensures that a permanent<br />

echo key condition will not occur. A one-cycle duration is<br />

more than adequate for recognition by modern relays. The<br />

traditional criteria that the echo signal be longer than the<br />

remote breaker tripping time is not required if relays include<br />

an appropriate minimum trip duration time. The echo signal<br />

only needs to be long enough to initiate the high-speed trip,<br />

after which the trip duration delay determines how long a local<br />

trip signal is maintained to the breaker. A typical trip duration<br />

setting is 9 cycles.<br />

56 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


6<br />

VI. ZONE 1 OVERREACH DUE TO CVT<br />

In July 2002, a 230 kV transmission line experienced a<br />

C-phase-to-ground (CG) fault. The local relay tripped by<br />

Zone 1 ground distance element for a fault that was physically<br />

located beyond the Zone 1 reach setting. A capacitive voltage<br />

transformer (CVT) supplies secondary voltage for the relay.<br />

The unfiltered event data from this misoperation are shown<br />

in Fig. 14. The event shows a severe transient in the C-phase<br />

voltage just before the Zone 1 CG distance element operation.<br />

This transient makes the C-phase voltage magnitude appear<br />

smaller to the relay than the actual voltage value. The apparent<br />

impedance calculated by the relay is smaller (or closer to the<br />

terminal) than actual. The CG distance element asserted in less<br />

than 1 cycle. When the fault clears, the C-phase voltage has<br />

another transient that makes it appear higher in magnitude<br />

than the other nonfault-affected phases.<br />

IA IB IC IRMag<br />

VA VB VC<br />

3V1 3I1<br />

VCMag<br />

Digitals<br />

1000<br />

0<br />

-1000<br />

100<br />

0<br />

-100<br />

1000<br />

0<br />

-1000<br />

IA IB IC IRMag VA<br />

VB VC 3V1 3I1 VCMag<br />

.75 cycles<br />

150<br />

100<br />

50<br />

0<br />

OUT 1&2 1 .<br />

32 Q<br />

ZCG 1<br />

Fig. 14.<br />

2 3 4 5 6 7 8 9 10 11<br />

Cycles<br />

CG Fault Produces CVT Transient<br />

Equation (1) and the data from the event allow us to<br />

calculate the source impedance ratio (SIR) for this application.<br />

CVTs with an active ferroresonance-suppression circuit (FSC)<br />

and applications with a high SIR limit secure Zone 1 reach<br />

settings [7]. The positive-sequence source impedance for this<br />

application is 11 ohms secondary and the line impedance is<br />

0.59 ohms secondary, giving an SIR <strong>of</strong> over 18.<br />

V1FAULT<br />

– V1PREFAULT<br />

SIR = (1)<br />

(I1 FAULT – I1PREFAULT)<br />

• Z1 LINE<br />

For an SIR <strong>of</strong> 18 and an active FSC, the maximum<br />

recommended Zone 1 reach is 0.07 pu <strong>of</strong> the line. With such a<br />

short line (small impedance), this results in a secondary<br />

setting below the minimum value allowed by the relay<br />

(0.05 ohms secondary). In this application, the Zone 1 element<br />

needs to be completely disabled or time-delayed by at least<br />

1.5 cycles. In some applications, another option is to split<br />

Zone 1 into two zones; one zone with reduced reach and set to<br />

trip instantaneously, and the other zone set to the desired or<br />

normal reach with time delay. This is a more complex<br />

solution, but it gives faster tripping for close-in faults and<br />

takes advantage <strong>of</strong> the multiple zones available in relays.<br />

Modern relays now include CVT transient detection logic.<br />

This logic detects SIRs greater than 5 during a fault and<br />

blocks the Zone 1 distance tripping for up to 1.5 cycles or<br />

until the distance calculation “smoothes,” indicating the<br />

transient has subsided. This logic is ideal for preventing<br />

distance element overreach for CVT transients. The raw event<br />

data in Fig. 14 were converted to IEEE COMTRADE files and<br />

replayed into a relay with CVT transient detection logic. With<br />

CVT transient detection logic turned <strong>of</strong>f, the relay<br />

overreached for every simulation. With CVT transient<br />

detection logic turned on, the logic correctly blocked the<br />

Zone 1 tripping for the entire 1.5-cycle duration and<br />

successfully prevented misoperation for all simulations.<br />

VII. CVT CAUSES RECLOSE FAILURE<br />

In June 2005, a 161 kV line experienced a BG fault. Both<br />

line ends use line-side CVTs. Fault data from the remote<br />

terminal are shown in Fig. 15, and fault data from the local<br />

terminal are shown in Fig. 16. Both terminals report balanced<br />

voltages and appropriate load currents flowing from the local<br />

to remote end before the fault occurs.<br />

For a BG fault that is forward to both terminals, we would<br />

expect the phase angle relationships shown in Fig. 15. However,<br />

voltage during the fault at the local end does not match<br />

that at the remote terminal.<br />

Fig. 15.<br />

Fig. 16.<br />

180<br />

135<br />

225<br />

VA<br />

VC<br />

90<br />

270<br />

IC<br />

VB<br />

IA<br />

IB<br />

45<br />

315<br />

Forward BG Fault as Viewed by Remote Line Relay<br />

180<br />

135<br />

225<br />

IC<br />

90<br />

VA<br />

IA<br />

270<br />

IB<br />

VC<br />

45<br />

VB<br />

315<br />

Forward BG Fault as Viewed by Local Line Relay<br />

Multiple CVT secondary grounds were suspected because<br />

the local prefault voltages were balanced and the fault voltages<br />

were corrupt. Technicians found two grounds, one at the<br />

CVT secondary wiring in the yard and another in the control<br />

building at the relay panel.<br />

Despite the CVT grounding problem, both terminals tripped<br />

correctly by a forward ground directional overcurrent element.<br />

0<br />

0<br />

Lessons Learned Analyzing Transmission Faults | 57


7<br />

After the remote terminal tripped, a reclose was expected after<br />

approximately 0.5 seconds. The fault was not permanent, but<br />

the reclose attempt failed.<br />

Oscillography from the local relay is shown in Fig. 17. The<br />

relay’s polarizing voltage has a memory <strong>of</strong> about one second.<br />

The CVT transient response and long-time decay are evident.<br />

This decaying voltage continues to feed the relay positivesequence<br />

(V1) memory and corrupts V1 magnitude and angle.<br />

When the line terminal attempts a reclose, the V1 memory has<br />

not reset, and inrush current from a tapped transformer load is<br />

present (see Fig. 18). The high current combined with the<br />

memory voltage error produces a Zone 1 distance element trip.<br />

IA IB IC<br />

VA VB VC<br />

Digitals<br />

2000<br />

1000<br />

0<br />

-1000<br />

-2000<br />

100<br />

50<br />

0<br />

-50<br />

-100<br />

ZAB<br />

IA IB IC VA VB VC<br />

VA VB VC<br />

V1Mag<br />

V1Ang<br />

100<br />

0<br />

-100<br />

100<br />

75<br />

50<br />

25<br />

0<br />

100<br />

0<br />

-100<br />

Fig. 17.<br />

IA IB IC<br />

2500<br />

0<br />

-2500<br />

VA VB VC V1Mag V1Ang<br />

1 2 3 4 5 6 7 8 9 10 11<br />

Cycles<br />

Forward BG Fault as Viewed by Local Line Relay (Filtered Data)<br />

IA IB IC IAMag IBMag<br />

ICMag VA VB VC<br />

Fig. 19.<br />

0 1 2 3 4 5 6 7 8 9 10 11 12<br />

Cycles<br />

COMTRADE Replay <strong>of</strong> Event Data With V1 Memory Reset<br />

VIII. BLOWN PT FUSE<br />

In June 2000, both primary and backup relays at a 69 kV<br />

line terminal tripped for a potential transformer (PT) fuse<br />

problem. There was no system fault at the time <strong>of</strong> trip. The<br />

relays were different models but both provided distance and<br />

directional overcurrent functions. The primary relay was used<br />

in a DCB scheme and the backup relay provided step-distance<br />

protection and tripped instantaneously for Zone 1 faults.<br />

The false apparent impedance created by abnormal voltages<br />

and load currents caused the apparent impedance to<br />

encroach on the reach <strong>of</strong> the distance relays (see Fig. 20). The<br />

same bus potentials serve both relays.<br />

A B C<br />

Fuses<br />

IAMag IBMag ICMag<br />

2000<br />

1000<br />

0<br />

100<br />

To Relay<br />

VA VB VC<br />

Digitals<br />

0<br />

-100<br />

ZAB 3 1 1 1 1 1 1 1<br />

PT<br />

Loose<br />

Fig. 18.<br />

0 1 2 3 4 5 6 7 8 9 10 11 12<br />

Cycles<br />

Raw Data From Local Relay During Reclose<br />

Event data from the reclose were converted to IEEE<br />

COMTRADE format and replayed into a similar relay in the<br />

lab. With no influence from corrupted V1 memory as in the<br />

real event in the field, the Zone 1 distance element (ZAB_1)<br />

did not trip (see Fig. 19). This was because the V1 memory<br />

was reset at the time the test was started and a fast recharge<br />

circuit instructed the relay to use actual measured V1.<br />

The CVT transient coupled with a fast reclose attempt<br />

caused the reclose failure. Fault detectors that supervise<br />

distance elements should be set above transformer inrush<br />

currents [8]. Reclose-open intervals must take into account<br />

CVT transient response and decay and V1 memory in relays.<br />

Fig. 20.<br />

PT Fuse Problem<br />

Because a blown fuse results in a loss <strong>of</strong> polarizing inputs<br />

to the relays, detection <strong>of</strong> this condition is desirable and was<br />

enabled in both relays (see Fig. 21). The event data show that<br />

the loss-<strong>of</strong>-potential (LOP) detection asserts after the phase<br />

distance element trips. In both relays, there is a three-cycle<br />

delay before the LOP element asserts for unbalanced conditions<br />

to ensure LOP does not block protective elements during<br />

a fault.<br />

58 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


8<br />

1_VA 1_VB 1_VC<br />

2_VA 2_VB 2_VC<br />

3_VA(kV) 3_VB(kV) 3_VC(kV)<br />

25<br />

0<br />

-25<br />

0<br />

-50<br />

25<br />

0<br />

-25<br />

1_VA 1_VB 1_VC 2_VA 2_VB<br />

2_VC 3_VA(kV) 3_VB(kV) 3_VC(kV)<br />

.020833 sec<br />

Digitals<br />

3_LOP<br />

2_OUT TP *<br />

2_LOP * *<br />

2_IN 52A *<br />

L<br />

2_50P<br />

2_21P 1<br />

1_ZBC 4 2<br />

1_OUT 1&2 1<br />

1_LOP * *<br />

1_IN 1&2 B 2<br />

1_50P<br />

L<br />

1_32 Q Q<br />

50.50 50.55 50.60 50.65 50.70 50.75<br />

Event Time (Sec) 23:31:50.588677<br />

Fig. 21. Response <strong>of</strong> Primary (1), Backup (2), and Relay With Improved LOP<br />

Logic (3) to PT Fuse Problem<br />

In the primary relay, an LOP is detected when negativesequence<br />

voltage (V2) is greater than 14 volts secondary and<br />

negative-sequence current (3I2) is less than 0.5 amperes<br />

secondary [9]. In the backup relay, an LOP is detected when<br />

zero-sequence voltage (V0) is greater than 14 volts secondary<br />

and zero-sequence current (I0) is less than 0.083 amperes<br />

secondary [10]. Once asserted, LOP will block distance and<br />

directional elements that rely on healthy voltage signals.<br />

This event emphasizes that early LOP logic was designed<br />

to protect distance elements from misoperating for system<br />

faults that occurred some time after an initial LOP condition<br />

was detected. Overcurrent fault detectors, set above load, were<br />

used to prevent distance element misoperation when the LOP<br />

condition first occurred. In this event, the fault detector (50L)<br />

was picked up during balanced load flow. Ideally, fault<br />

detectors should be set above expected load currents and<br />

below minimum fault levels to ensure correct distance relay<br />

operation.<br />

Newer relays have LOP logic that operates based on the V1<br />

rate <strong>of</strong> change versus the rate <strong>of</strong> change <strong>of</strong> currents. The new<br />

logic operates in less than 0.5 cycles, so distance element<br />

security is less dependent on the fault detector settings [11].<br />

In Fig. 21, the response <strong>of</strong> the original relays are shown<br />

with that <strong>of</strong> a relay with improved LOP logic. The new relay<br />

LOP logic operates more than 1 cycle before any distance<br />

elements assert, ensuring this misoperation would not happen<br />

again.<br />

IX. DIRECTIONAL ELEMENT CHALLENGE<br />

In June 2007, 80 MW <strong>of</strong> generation was tripped <strong>of</strong>fline<br />

because <strong>of</strong> an unknown cause (at that time). The generator<br />

stepup transformer was fed radially from a 138 kV tie line to a<br />

local utility for several seconds. The line was protected by a<br />

DCB scheme.<br />

A BG fault then occurred on the line. The relay at the<br />

generator end <strong>of</strong> the line saw this fault as reverse and sent a<br />

blocking signal to the utility terminal, delaying fault clearing.<br />

Event data from the relay at the generator end <strong>of</strong> the line are<br />

shown in Fig. 22. The phase currents are all in phase.<br />

Fig. 22.<br />

Fault Data From Relay<br />

A one-line and symmetrical-components diagram for the<br />

fault is shown in Fig. 23. With the generator breaker open,<br />

there is a pure zero-sequence source behind the relay.<br />

S<br />

–<br />

+<br />

Fig. 23.<br />

T<br />

Relay<br />

Z1S Z1T m • Z1L (1–m)Z1L Z1R<br />

Relay<br />

Z2S Z2T m • Z2L (1–m)Z2L Z2R<br />

Relay<br />

Z0S Z0T m • Z0L (1–m)Z0L Z0R<br />

Relay<br />

L<br />

3R F<br />

Symmetrical Components Diagram for BG Line Fault<br />

R<br />

–<br />

+<br />

Lessons Learned Analyzing Transmission Faults | 59


9<br />

The relay’s directional element can use negative-sequence<br />

voltage polarization (Q), zero-sequence voltage polarization<br />

(V), or current polarization (I) [10]. Only one element can be<br />

used at any one time. The user selected the negative-sequence<br />

element; its torque equation is shown as (2).<br />

32Q = V I cos ∠−V<br />

– ∠I<br />

+ MTA (2)<br />

( ( )<br />

T ∠<br />

2 2<br />

2 2<br />

With the local generation connected, negative-sequence<br />

polarization would have been dependable. However, with the<br />

generation isolated, the negative-sequence polarized element<br />

does not correctly determine the fault direction. This is<br />

because <strong>of</strong> the lack <strong>of</strong> negative-sequence current from the pure<br />

zero-sequence source (wye-grounded transformer) behind the<br />

relay. By replaying the event data into a relay, we prove that<br />

zero-sequence voltage polarization correctly determines this to<br />

be a forward fault. The event data indicate that zero-sequence<br />

current polarization would have also correctly declared a<br />

forward fault for this event; I POL and 3I0 are in phase.<br />

The user can choose to change the polarization quantity.<br />

This decision, however, sacrifices a more reliable directional<br />

element in order to trip faster during a rare operating condition.<br />

A practical improvement would be to wire a 52b contact<br />

from the generator breaker in parallel with the carrier squelch<br />

contact from the relay. This would allow high-speed clearing<br />

from the utility source for faults that occur when the<br />

generation is <strong>of</strong>fline.<br />

The directional elements in newer relays <strong>of</strong>fer the ability to<br />

choose the best choice for the current system configuration<br />

and fault without requiring settings changes [3]. Multiple<br />

elements with different polarizing quantities are processed<br />

simultaneously. One setting allows the user to determine<br />

which polarizing source to use first, second, and third. If the<br />

first-choice element does not have adequate quantities, then<br />

the second choice will be evaluated. If the second choice does<br />

not have adequate signal, then the third and final choice is<br />

evaluated.<br />

Event data were replayed into a newer relay with best<br />

choice ground directional element logic (see Fig. 24). 32VE<br />

is the zero-sequence polarizing enable element, and F32V and<br />

32GF are the forward ground fault directional declaration bits.<br />

32IE is the current polarizing enable element, and F32I and<br />

32GF are the forward ground fault directional declaration bits.<br />

32QGE and 32QE are the negative-sequence element enables,<br />

and 32QF and 32GF are the forward ground fault directional<br />

declaration bits.<br />

In the simulation, we chose an order that always gives<br />

preference to negative-sequence polarization (Q, V, I). In this<br />

case, there is little negative-sequence, so the relay checks<br />

zero-sequence and makes the proper forward directional<br />

declaration. This secure operation comes at the expense <strong>of</strong> a<br />

slight processing delay. Overall, this directional logic results<br />

in faster operating times for all system states. When the<br />

generation is online, the negative-sequence directional element<br />

operates most reliably, and when generation is <strong>of</strong>fline,<br />

the zero-sequence element operates correctly.<br />

IA IB IC<br />

VA(kV) VB(kV) VC(kV)<br />

Digitals<br />

1000<br />

0<br />

-1000<br />

50<br />

0<br />

-50<br />

Fig. 24.<br />

32QF<br />

F32I<br />

32GF<br />

F32V<br />

32GR<br />

R32V<br />

32IE<br />

32VE<br />

32QE<br />

32QGE<br />

IA IB IC VA(kV) VB(kV) VC(kV)<br />

.125 cycles<br />

0.0 2.5 5.0 7.5 10.0 12.5 15.0<br />

Cycles<br />

Best Choice Ground Directional Element Logic for BG Fault<br />

If none <strong>of</strong> the directional element enables (32QGE, 32VE,<br />

and 32IE) assert, then the relay defaults to the negativesequence<br />

element (as long as Q is listed in the order setting).<br />

When the relay defaults in this manner, it also bypasses a ratio<br />

check <strong>of</strong> negative-sequence to zero-sequence. It was this ratio<br />

check, |I2| > k2 • |I0|, that prevented the negative-sequence<br />

element from operating in the simulation.<br />

X. WHEN STANDARD SETTINGS DO NOT WORK<br />

In August 2006, a 230 kV line terminal tripped incorrectly<br />

for a reverse CG fault. The line is a three-terminal DCB<br />

scheme. There is no positive-sequence source behind the<br />

relay, but there are two ground sources—an autotransformer<br />

with a delta tertiary and a wye-grounded high-side, delta lowside<br />

power transformer.<br />

The fault data from the event are shown in Fig. 25. During<br />

the fault, V2 leads I2 by 83 degrees and has a magnitude <strong>of</strong><br />

15 volts secondary. The fault current magnitude is<br />

3700 amperes primary. The forward directional element<br />

(32QF) asserts incorrectly and allows a trip by the directional<br />

ground overcurrent element (67N2dcb).<br />

IA IB IC IRMag<br />

VA VB VC<br />

V2Ang I2Ang<br />

Digitals<br />

2500<br />

0<br />

-2500<br />

100<br />

0<br />

-100<br />

100<br />

0<br />

-100<br />

Fig. 25.<br />

32QF<br />

3PT<br />

67N2pu<br />

67N2dcb<br />

IA IB IC IRMag VA<br />

VB VC V2Ang I2Ang<br />

2 3 4 5 6 7 8 9 10 11<br />

Cycles<br />

Forward Fault Declaration for Reverse CG Fault<br />

60 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


10<br />

The phasors during the fault are shown in Fig. 26. C-phase<br />

fault current leads a reduced-magnitude C-phase voltage as<br />

expected for a reverse fault. The expected operation would<br />

have been for the relay to see this fault as reverse and send a<br />

blocking signal to the remote terminals.<br />

Fig. 26.<br />

180<br />

135<br />

225<br />

IC<br />

VB<br />

CG Reverse Fault Phasors<br />

VA<br />

90<br />

270<br />

IA<br />

IB<br />

VC<br />

45<br />

315<br />

Fig. 27 shows a MathCad simulation <strong>of</strong> the directional<br />

element. The directional element measures the negativesequence<br />

impedance (Z2) during the fault and compares it to<br />

settings to determine the fault direction.<br />

0.9<br />

m 0.4<br />

–0.1<br />

Fig. 27.<br />

3 4 5 6 7 8 9 10 11<br />

Cycles<br />

Directional Element Response With As-Set Settings<br />

For forward faults, Z2 is equal to the source impedance<br />

behind the relay (a negative value). For reverse faults, Z2 is<br />

equal to the remote source impedance plus the line impedance<br />

(a positive value). In Fig. 27, Z2 plots below the forward<br />

threshold, so the relay declares this a forward fault.<br />

The initial root cause theory blamed the forward and<br />

reverse threshold settings for the misoperation. The original<br />

forward threshold (Z2F) was set at one-half Z1MAG<br />

(positive-sequence line impedance in secondary ohms). The<br />

original reverse threshold (Z2R) was set at Z2F plus 0.1 ohms<br />

secondary. This is typical and is used by some relays that<br />

calculate thresholds automatically.<br />

There are four reasons this theory seemed credible. First,<br />

the directional decision made by the relay was incorrect.<br />

Second, by modifying the Z2F and Z2R settings the<br />

directional decision made by the relay would have been<br />

correct (reverse). Fig. 28 shows a MathCad simulation <strong>of</strong> the<br />

directional element with Z2F set to –0.1 ohms and Z2R set to<br />

+0.1 ohms secondary. Third, the engineers experienced a few<br />

cases where the standard settings did not work. Lastly, a<br />

subsequent reverse fault operated correctly with the modified<br />

settings.<br />

0<br />

m<br />

1<br />

0.5<br />

0<br />

–0.5<br />

–1<br />

3 4 5 6 7 8 9 10 11<br />

Cycles<br />

Fig. 28. Directional Element Response With Z2F = –0.1 Ohms and<br />

Z2R = +0.1 Ohms<br />

Standard settings <strong>of</strong>fer convenience but should not be<br />

applied blindly. Directional element performance is critical<br />

and should be checked, especially for extreme conditions. Run<br />

fault studies for very strong or very weak sources, and check<br />

Z2 against thresholds Z2F and Z2R for proper operation.<br />

Consider a three-terminal line where one terminal has a<br />

very weak source (i.e., a large transformer). If the line is<br />

protected by a DCB scheme, any terminal will trip if an<br />

internal fault is seen without receiving a block. If the line is<br />

energized from two ends and the weak source breaker is<br />

closed to energize the large transformer, the relay at that<br />

terminal may be subjected to nearly balanced voltages (low<br />

V2) but unbalanced currents. Standard directional settings<br />

based on line parameters default to a forward decision for very<br />

low V2 values. Therefore, the weak terminal could trip<br />

because <strong>of</strong> unbalanced current (3I0) and an incorrect<br />

directional decision (forward for reverse infeed). For<br />

transformer applications, it is more secure to set Z2F to –0.1<br />

to –0.5 ohms and Z2R to +0.1 to +0.5 ohms, respectively [12].<br />

Upon closer inspection <strong>of</strong> these event data, however, it<br />

appears that the Z2F and Z2R thresholds may not be the<br />

culprits at all. Several data anomalies are present. Standard<br />

settings tend to give us problems when V2 is small, such as in<br />

the transformer inrush example above. In this event, however,<br />

we have plenty <strong>of</strong> V2 (15 volts secondary). The line impedance<br />

and line length settings in the relay give us an estimate<br />

<strong>of</strong> 1.6 ohms per mile for a 230 kV line, which appears too<br />

large (half that value would be more reasonable). Z2 measured<br />

during a reverse fault should equal at least the line impedance;<br />

the data show a Z2 measurement <strong>of</strong> 0.5 ohms, less than half <strong>of</strong><br />

the line impedance setting. This led to an alternative theory.<br />

Perhaps the standard thresholds were just fine, but the<br />

measured impedance was wrong. An incorrect CT ratio (CTR)<br />

setting or an incorrect CT tap could do just that. The CTR<br />

setting in the relay is 40 (200:5) for a 230 kV line. Assume<br />

that the CT is tapped at 100:5, while the relay settings expect<br />

200:5. If we leave directional thresholds at their original<br />

standard settings, we get a correct reverse decision for the<br />

original fault (see Fig. 29).<br />

Lessons Learned Analyzing Transmission Faults | 61


11<br />

m<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

3 4 5 6 7 8 9 10 11<br />

Cycles<br />

Fig. 29. Directional Element Response With CT Tap Lower Than CTR<br />

Setting<br />

It is suspected that the CTR setting in the relay does not<br />

match the actual CT tap in the field. However, this is part <strong>of</strong><br />

an ongoing investigation and, at the time <strong>of</strong> publication, root<br />

cause is yet to be determined.<br />

XI. DIRECTIONAL ELEMENT AFFECTS CARRIER EXTENSION<br />

In June 2006, a 161 kV line tripped incorrectly for a<br />

reverse AG fault. The line was protected by a DCB scheme.<br />

The ground directional element used the best choice logic<br />

described in Section IX. The element used ground current<br />

polarization first (I), followed by zero-sequence voltage<br />

polarization (V), and then by negative-sequence voltage<br />

polarization (Q).<br />

The event data from the misoperation are shown in Fig. 30.<br />

For a reverse fault, we expect the polarizing current (IP) and<br />

the terminal ground current (IG) to be 180 degrees out <strong>of</strong><br />

phase. In the event data, however, IP and IG were in phase. A<br />

wiring error was found; the polarizing current CT was<br />

connected reverse polarity. This allowed the local breaker to<br />

trip.<br />

IA IB IC<br />

VA(kV) VB(kV) VC(kV)<br />

IP IG<br />

Digitals<br />

1000<br />

0<br />

-1000<br />

100<br />

0<br />

-100<br />

5000<br />

0<br />

-5000<br />

DSTRT<br />

Z3XT<br />

50G3<br />

32QR<br />

F32I<br />

67G1T<br />

IA IB IC VA(kV) VB(kV) VC(kV) IP IG<br />

2 3 4 5 6 7 8<br />

Cycles<br />

Fig. 30. Directional Element Misoperation due to IP Wiring Mistake<br />

The remote breakers were allowed to trip because <strong>of</strong> the<br />

carrier-blocking signal dropout as the breaker was interrupting<br />

the fault. In Fig. 31, directional carrier blocking (DSTRT) was<br />

asserted due to a negative-sequence element (50Q3) and a<br />

negative-sequence directional element (R32Q). At the end <strong>of</strong><br />

the event, the DSTRT element chattered then dropped out.<br />

This was because <strong>of</strong> the phase angle error introduced by<br />

arcing during interruption. Carrier blocking extension is<br />

enabled in this relay but requires Zone 3 elements to be<br />

asserted for 2 cycles; they were only asserted for 1.5 cycles.<br />

During the investigation, it was noted that the directional<br />

element processing order <strong>of</strong> IVQ was not typically used by<br />

this utility; an order setting <strong>of</strong> QV was more common. Would<br />

carrier extension have asserted if we had used the typical<br />

order<br />

In Fig. 31, the ground-polarized directional elements (F32I<br />

and 32GF) assert because <strong>of</strong> the directional logic order and<br />

wiring error. The phase angle <strong>of</strong> the faulted phase current<br />

varies wildly during the beginning <strong>of</strong> the fault because <strong>of</strong> the<br />

fault arcing and evolution. This delays the voltage-based<br />

directional elements but has little effect on the current-based<br />

directional element. Also in Fig. 31, negative-sequence<br />

directional elements (32QR and R32Q) asserted later in the<br />

fault. 32QR and R32Q supervise phase distance and negativesequence<br />

overcurrent elements. Their ground logic equivalent<br />

(R32QG) is controlled by best choice logic. R32QG did not<br />

assert because the IP-based element has priority, but the<br />

directional decision it would make would match that <strong>of</strong> 32QR<br />

and R32Q when enabled. From this, we realize that the order<br />

and best choice logic processing did not delay the R32QG<br />

element; the evolving fault did.<br />

IA IB IC<br />

VA(kV) VB(kV) VC(kV)<br />

Digitals<br />

1000<br />

0<br />

-1000<br />

100<br />

0<br />

-100<br />

50Q3<br />

50G3<br />

NSTRT<br />

DSTRT<br />

Z3XT<br />

R32V<br />

F32V<br />

R32QG<br />

F32QG<br />

R32Q<br />

F32Q<br />

R32I<br />

F32I<br />

32QE<br />

32IE<br />

32QGE<br />

32VE<br />

32GR<br />

32GF<br />

32QR<br />

IA IB IC VA(kV) VB(kV) VC(kV)<br />

0.0 2.5 5.0 7.5 10.0 12.5 15.0<br />

Cycles<br />

Fig. 31. Detailed Relay Response to Fault<br />

The evolving fault should delay a directional decision for<br />

all the relays. Remote relays additionally include a carrier<br />

coordination delay for tripping. How fast the directional<br />

element sent blocking is not the real issue in this fault; it is<br />

how fast we enable carrier extension. Setting the carrier<br />

extension pickup delay (Z3XPU) to a lower value, even zero,<br />

is recommended [13].<br />

XII. COMPARING PRIMARY AND BACKUP METERING<br />

In November 2005, a 138 kV transmission line experienced<br />

an AG fault. The backup relay correctly saw the fault as<br />

forward and tripped by the directional ground overcurrent<br />

element (67G1). The primary relay, however, declared a<br />

reverse fault and did not operate.<br />

62 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


12<br />

The prefault data from the backup relay are shown in<br />

Fig. 32, and the prefault data from the primary relay are<br />

shown in Fig. 33. The phase voltages seen by the primary<br />

relay in the prefault state did not look normal or balanced, and<br />

a significant standing zero-sequence voltage was present. The<br />

backup relay, on the contrary, reported normal prefault<br />

voltages.<br />

180<br />

135<br />

225<br />

135<br />

90<br />

VC (kV)<br />

IB<br />

IA<br />

VB (kV)<br />

270<br />

90<br />

IC<br />

45<br />

VA (kV)<br />

0<br />

315<br />

45<br />

The neutral bus <strong>of</strong> the primary relay three-phase voltage<br />

connections should have been connected at a terminal block to<br />

station ground; this wire was missing. The result was that the<br />

primary relay voltages were floating, and this distorted phaseto-neutral<br />

magnitudes, angles, and sequence components both<br />

before the fault and during the fault. The backup relay had<br />

properly terminated voltages and, therefore, its zero-sequence<br />

voltage polarized directional element performed correctly.<br />

The fault data collected from the backup relay are shown<br />

in Fig. 34. The zero-sequence current leads the zero-sequence<br />

voltage by about 120 degrees, as expected for a forward AG<br />

fault. Negative-sequence relationships are similar. The relay is<br />

set by the user to enable only zero-sequence quantities for<br />

directional decisions. However, the data show that the zerosequence<br />

directional element used (F32V) and the disabled<br />

negative-sequence directional element (F32Q) both made the<br />

correct directional decision.<br />

IA IB IC<br />

VA(kV) VB(kV) VC(kV)<br />

V0Ang I0Ang<br />

Digitals<br />

10000<br />

0<br />

-10000<br />

50<br />

0<br />

-50<br />

100<br />

0<br />

-100<br />

F32V<br />

F32Q<br />

67G1<br />

IA IB IC VA(kV) VB(kV) VC(kV) V0Ang I0Ang<br />

0.0 2.5 5.0 7.5 10.0 12.5 15.0<br />

Cycles<br />

Fig. 32.<br />

180<br />

225<br />

Prefault Data From Backup Relay<br />

135<br />

180<br />

225<br />

135<br />

180<br />

270<br />

90<br />

VC (kV)<br />

IB (A)<br />

IC (A)<br />

IA (A)<br />

VB (kV)<br />

270<br />

90<br />

V0<br />

V1<br />

0<br />

315<br />

45<br />

VA (kV)<br />

0<br />

315<br />

45<br />

V1<br />

0<br />

Fig. 34.<br />

Fault Data From Backup Relay<br />

The fault data collected from the primary relay are shown<br />

in Fig. 35. The zero-sequence current leads the zero-sequence<br />

voltage by about 210 degrees, which causes the zero-sequence<br />

directional element misoperation. Negative-sequence relationships<br />

match those reported by the backup relay and are<br />

correct. The relay is set by the user to enable only zerosequence<br />

quantities for directional decisions. However, the<br />

data show that the disabled negative-sequence directional<br />

element (F32Q) made the correct directional decision.<br />

IA(A) IB(A) IC(A)<br />

VA(kV) VB(kV) VC(kV)<br />

V0Ang I0Ang<br />

10000<br />

0<br />

-10000<br />

100<br />

0<br />

-100<br />

200<br />

0<br />

-200<br />

IA(A) IB(A) IC(A) VA(kV) VB(kV) VC(kV) V0Ang I0Ang<br />

225<br />

270<br />

315<br />

Digitals<br />

67G1<br />

F32V<br />

F32Q<br />

TRIP<br />

IN105<br />

Fig. 33.<br />

Prefault Data From Primary Relay<br />

Fig. 35.<br />

0.0 2.5 5.0 7.5 10.0 12.5 15.0<br />

Cycles<br />

Fault Data From Primary Relay<br />

Synchronized phasor measurement during commissioning<br />

is a useful tool for finding mistakes like these before they<br />

cause misoperations. When relays are connected to a common<br />

voltage or current source, automation systems can be easily<br />

designed to periodically retrieve metering data, compare them,<br />

and alarm when differences are discovered. Troubleshooting<br />

and repair can then be performed to fix problems before they<br />

are discovered by misoperations.<br />

Lessons Learned Analyzing Transmission Faults | 63


13<br />

XIII. CONCLUSIONS<br />

• Evolving faults can challenge fault location and<br />

targeting algorithms; <strong>of</strong>fline tools can improve fault<br />

location accuracy.<br />

• Evolving faults can delay distance element tripping;<br />

common zone timing allows faster fault clearing.<br />

• Evolving faults can create reclosing problems; use trip<br />

elements to initiate reclosing and nonreclosing elements<br />

in drive-to-lockout logic.<br />

• Carrier holes cause DCB scheme problems; BT<br />

extension logic adds security.<br />

• Sensitivities <strong>of</strong> local and remote line relays should be<br />

matched.<br />

• Commissioning tests should concentrate on proving the<br />

overall protection system.<br />

• Unique event reports should be archived as<br />

COMTRADE files and used to test new schemes.<br />

• CVT transients can cause Zone 1 overreach; CVT<br />

transient detection logic adds security.<br />

• Extend reclosing open intervals beyond CVT transient<br />

response and set fault detectors above inrush on lines<br />

with tapped loads.<br />

• Older LOP logic relied on fault detectors set above<br />

normal load currents; improved LOP logic operates<br />

faster and is less dependent on fault detector settings<br />

for security.<br />

• Relays that include multiple polarization schemes and<br />

choose the appropriate element for the system and fault<br />

<strong>of</strong>fer the best sensitivity and security.<br />

• Do not apply standard settings blindly; directional<br />

element performance is critical and should be checked,<br />

especially for extreme conditions.<br />

• Set carrier extension pickup delay to zero in DCB<br />

schemes.<br />

• Missing or multiple ground wires in instrument<br />

transformer circuits cause directional element<br />

misoperations; thorough commissioning tests and<br />

automated meter data comparisons should find these<br />

errors before misoperations occur.<br />

[4] G. Alexander, K. Zimmerman, and J. Mooney, “Setting the <strong>SEL</strong>-421<br />

Relay With Sub-Cycle Elements in a Directional Comparison Blocking<br />

Scheme,” Schweitzer Engineering Laboratories, Inc. Application Guide<br />

2007-10. [Online] Available: http://www.selinc.com/aglist.htm.<br />

[5] A. Guzman, J. Roberts, and K. Zimmerman, “Applying the <strong>SEL</strong>-321<br />

Relay to Permissive Overreaching Transfer Trip (POTT) Schemes,”<br />

Schweitzer Engineering Laboratories, Inc. Application Guide 95-29.<br />

[Online] Available: http://www.selinc.com/aglist.htm.<br />

[6] J. Roberts, E. O. Schweitzer III, R. Arora, and E. Poggi, “Limits to the<br />

Sensitivity <strong>of</strong> Ground Directional and Distance Protection,” presented at<br />

the 1997 Spring Meeting <strong>of</strong> the Pennsylvania Electric Association Relay<br />

Committee. [Online] Available: http://www.selinc.com/techpprs.htm.<br />

[7] D. Hou and J. Roberts, “Capacitive Voltage Transformers: Transient<br />

Overreach Concerns and Solutions for Distance Relaying,” presented at<br />

the 22nd Annual Western Protective Relay Conference in Spokane, WA,<br />

October 1995. [Online] Available: http://www.selinc.com/techpprs.htm.<br />

[8] J. Mooney and S. Samineni, “Distance Relay Response to Transformer<br />

Energization: Problems and Solutions,” presented at the Texas A&M<br />

Conference for Protective Relay Engineers, College Station, TX,<br />

March 2007. [Online] Available: http://www.selinc.com/techpprs.htm.<br />

[9] <strong>SEL</strong>-321 Instruction Manual, Schweitzer Engineering Laboratories, Inc.,<br />

Pullman, WA.<br />

[Online] Available: http://www.selinc.com/instruction_manual.htm.<br />

[10] <strong>SEL</strong>-221G5 Instruction Manual, Schweitzer Engineering Laboratories,<br />

Inc., Pullman, WA.<br />

[Online] Available: http://www.selinc.com/instruction_manual.htm.<br />

[11] J. Roberts and R. Folkers, “Improvements to the Loss-<strong>of</strong>-Potential<br />

(LOP) Function in the <strong>SEL</strong>-321,” Schweitzer Engineering Laboratories,<br />

Inc. Application Guide 2000-05.<br />

[Online] Available: http://www.selinc.com/aglist.htm.<br />

[12] K. Zimmerman, “Negative-Sequence Directional Element Revisited,”<br />

presented at the 21st Annual <strong>SEL</strong> Technical Seminar at the Western<br />

Protective Relay Conference in Spokane, WA, October 2005.<br />

[13] G. Alexander, K. Zimmerman, and J. Mooney, Schweitzer Engineering<br />

Laboratories, Inc. Application Guide 2007-10.<br />

[Online] Available: http://www.selinc.com/aglist.htm.<br />

XVI. BIOGRAPHY<br />

David Costello graduated from Texas A&M University in 1991 with a BSEE.<br />

He worked as a system protection engineer at Central <strong>Power</strong> and Light and<br />

Central and Southwest Services in Texas and Oklahoma. He has served on the<br />

System Protection Task Force for ERCOT. In 1996, David joined Schweitzer<br />

Engineering Laboratories, Inc., where he has served as a field application<br />

engineer and regional service manager. He presently holds the title <strong>of</strong> senior<br />

application engineer and works in Boerne, Texas. He is a senior member <strong>of</strong><br />

IEEE, and a member <strong>of</strong> the planning committee for the Conference for<br />

Protective Relay Engineers at Texas A&M University.<br />

XIV. ACKNOWLEDGMENT<br />

The author gratefully acknowledges Normann Fischer,<br />

Karl Zimmerman, Bill Fleming, Joe Mooney, and Matt Leoni<br />

for their contributions to the original event analyses.<br />

XV. REFERENCES<br />

[1] <strong>SEL</strong>-311B Instruction Manual, Schweitzer Engineering Laboratories,<br />

Inc., Pullman, WA.<br />

[Online] Available: http://www.selinc.com/instruction_manual.htm.<br />

[2] K. Zimmerman and D. Costello, “Impedance-Based Fault Location<br />

Experience,” presented at the 31st Annual Western Protective Relay<br />

Conference in Spokane, WA, October 2004.<br />

[Online] Available: http://www.selinc.com/techpprs.htm.<br />

[3] <strong>SEL</strong>-311C Instruction Manual, Schweitzer Engineering Laboratories,<br />

Inc., Pullman, WA.<br />

[Online] Available: http://www.selinc.com/instruction_manual.htm.<br />

© 2007 Schweitzer Engineering Laboratories, Inc.<br />

All rights reserved.<br />

20070913 • TP6280-01<br />

64 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


1<br />

Adaptive Phase and Ground<br />

Quadrilateral Distance Elements<br />

Fernando Calero, Armando Guzmán, and Gabriel Benmouyal, Schweitzer Engineering Laboratories, Inc.<br />

Abstract—Quadrilateral distance elements can provide<br />

significantly more fault resistance coverage than mho distance<br />

elements for short line applications. Quadrilateral phase and<br />

ground distance element characteristics result from the<br />

combination <strong>of</strong> several distance elements. Directional elements<br />

discriminate between forward and reverse faults, while reactance<br />

and resistance elements are fundamental to the proper<br />

performance <strong>of</strong> the quadrilateral characteristic. Load flow<br />

considerations determine the choice <strong>of</strong> the polarizing quantity for<br />

these elements. Reactance elements must accommodate load flow<br />

and adapt to it. Resistive blinders should detect as much fault<br />

resistance as possible without causing excessive overreach or<br />

underreach <strong>of</strong> the quadrilateral distance element. In this paper,<br />

we discuss an adaptive quadrilateral distance scheme that can<br />

detect greater fault resistance than a previous implementation.<br />

We also discuss application considerations for quadrilateral<br />

distance elements.<br />

Phase B are shown), the tower structure, the insulator chains,<br />

the ground wire, and the different impedances to the flow <strong>of</strong><br />

fault current. These impedances are simplified to be resistive<br />

values only [4][5].<br />

I. OVERVIEW<br />

Whereas the literature debates the differences and benefits<br />

<strong>of</strong> mho and quadrilateral ground distance elements [1], this<br />

paper describes the theory, application, and characteristics <strong>of</strong> a<br />

particular implementation <strong>of</strong> phase and ground quadrilateral<br />

distance elements.<br />

It is well accepted that a quadrilateral characteristic is<br />

beneficial when protecting short transmission lines [1][2]. It is<br />

also accepted that sensitive pilot protection schemes do not<br />

rely on distance elements only; these schemes also rely on<br />

ground directional overcurrent (67G), a unit that provides<br />

higher fault resistance (Rf) detecting capabilities than ground<br />

distance elements <strong>of</strong> any shape [3].<br />

Generally, high Rf faults have been associated with singleline-to-ground<br />

faults (AG, BG, CG). For these faults, the<br />

associated Rf is considerable. On the other hand, phase faults<br />

are less susceptible to high Rf values. However, because short<br />

transmission lines are much more affected by high Rf values,<br />

the element with the most fault detecting capabilities should<br />

be used [1][2][3].<br />

A. Fault Resistance<br />

Short circuits along the transmission line will have some<br />

degree <strong>of</strong> additional impedance. If this additional impedance<br />

is negligible, the line impedance is prevalent, and the apparent<br />

impedance measured will reflect it by reporting an impedance<br />

with the same angle as the line impedance. On the other hand,<br />

if this additional impedance is not negligible, the measured<br />

apparent impedance no longer appears at the line angle.<br />

Fig. 1 shows the different components <strong>of</strong> fault resistance<br />

for transmission line faults. Although extremely simplified,<br />

the figure shows the phase conductors (only Phase A and<br />

Fig. 1. Visualizing the Rf component<br />

The Rtower resistance is generally called the “footing<br />

resistance.” It is a critical parameter regarding the design and<br />

construction <strong>of</strong> transmission lines [6]. For an insulation<br />

flashover fault, the return path is through the tower itself.<br />

When a foreign object touches the conductors, the current<br />

distributes between adjacent towers but returns through the<br />

footing resistance. Ideally, the smaller the footing resistance,<br />

the better the transmission line ground fault detection<br />

performance will be. However, even though smaller values<br />

exist, practical values range from 5 to 20 ohms; and in rocky<br />

terrain, the resistance could be 100 ohms or more [1].<br />

In Fig. 1, Raφg represents the arc resistance for an insulator<br />

flashover for a phase-to-ground fault. This is in the path <strong>of</strong> the<br />

ground fault flashover current Igffo. Raφφ is the arc resistance<br />

for a phase-to-phase fault.<br />

The arc resistance value is dependent on the arc length and<br />

the current flowing through the arc. A well-accepted formula<br />

is the one empirically derived by A. Van Warrington,<br />

expressed in (1). Other equations yield similar results [7]. In<br />

(1), the arc length is expressed in meters.<br />

Rarc length<br />

= 28688.5 Ω (1)<br />

I 1.4<br />

The arc initially presents a few ohms <strong>of</strong> impedance. Over<br />

time, it could develop into 50 or more ohms [1]. Importantly,<br />

its value is dependent on the arc length and the current<br />

Adaptive Phase and Ground Quadrilateral Distance Elements | 65


2<br />

flowing through the arc. In Fig. 1, the arc length is denoted as<br />

dφg for ground faults and dφφ for phase faults.<br />

Rstruc is the tower structure resistance. Although<br />

insignificant for a metallic structure, this resistance may carry<br />

a significant value if built from a nonconductive material like<br />

wood.<br />

Rtreeg and Rtreep are the resistances <strong>of</strong> foreign objects that<br />

could be causing a power system fault. A tree is chosen as an<br />

example. These resistance values could be a few hundred<br />

ohms.<br />

1) Phase-to-Ground Faults<br />

Phase-to-ground faults are the most common type <strong>of</strong> faults<br />

in the power system. They involve a single phase that<br />

conducts fault current to ground. There are two possible<br />

single-phase-to-ground fault scenarios: insulator flashover and<br />

an object creating a path to ground.<br />

a) Insulator Flashover<br />

An insulator flashover (arc resistance Raφg), which may be<br />

due to a lightning strike or any other event that would stress<br />

the insulator, conducts fault current from the phase conductor<br />

to the tower structure (Igffo) and then to ground through the<br />

“tower footing resistance” (Rtower). The arc forms on the dφg<br />

length. This length is the “creepage distance” <strong>of</strong> the insulator<br />

string, which is the shortest electrical distance between the<br />

conductor and the tower measured along the insulator string<br />

structure.<br />

b) Ground Fault Through an Object<br />

Another possible phase-to-ground fault may occur when<br />

the phase conductor contacts an object, such as a tree (Rtree),<br />

which is in contact with ground (see Fig. 1). The contact is<br />

most likely not at the tower location. It could occur any place<br />

along the span from one tower to another tower. The fault<br />

current distributes to ground through the tower resistances,<br />

with a larger percentage <strong>of</strong> current flowing to the footing<br />

resistance <strong>of</strong> the closer tower. Conservatively, we can assume<br />

that current is only flowing through a single tower footing<br />

resistance. This simple assumption contrasts with other<br />

advanced and accurate analysis techniques [8].<br />

Regardless <strong>of</strong> the two possible scenarios, the path to<br />

ground involves the equivalent Rtower, which is the resistance<br />

<strong>of</strong> the composite path from earth to system ground. For an<br />

insulator flashover, Rf is the sum <strong>of</strong> Raφg and Rtower,<br />

ignoring the tower resistance (Rstruc). For a ground fault<br />

occurring because <strong>of</strong> contact with an object to ground, Rf is<br />

the sum <strong>of</strong> Rtreeg and Rtower. The Rf component for this type<br />

<strong>of</strong> fault can be significant.<br />

The presence <strong>of</strong> ground wires in the tower distributes the<br />

fault current differently. A portion <strong>of</strong> the fault current will<br />

return to ground (Igw) through these wires. The ground wires<br />

are part <strong>of</strong> the zero-sequence impedance and therefore not<br />

associated to Rf.<br />

2) Phase-to-Phase Faults<br />

Phase-to-phase faults, as Fig. 1 illustrates, do not involve<br />

the ground return path. As with phase-to-ground faults, an<br />

insulator flashover or phase-to-phase connection through an<br />

object could be the cause <strong>of</strong> the fault.<br />

If the fault is due to insulation flashover, Rf is expressed by<br />

(1), and the arc length could be a straight line or a path around<br />

the tower (dφφ). The important factor is that Rf is fully due to<br />

the arc resistance.<br />

Because <strong>of</strong> the spacing between phases in high-voltage<br />

(HV) and extra-high-voltage (EHV) transmission networks<br />

and even in subtransmission levels, it is highly improbable<br />

that an object could produce a phase-to-phase fault because <strong>of</strong><br />

contact. However, in distribution networks, phase-to-phase<br />

faults have a higher probability <strong>of</strong> occurrence because the<br />

conductor can have contact with different objects, like tree<br />

branches, flying debris, etc.<br />

B. The Need for a Quadrilateral Element in Transmission<br />

Networks<br />

The following three conclusions can be made based on<br />

Fig. 1:<br />

• The arc component <strong>of</strong> the fault, Raφg or Raφφ, has a<br />

value that can be estimated. Equation (1) indicates that<br />

the value may not be significant for transmission<br />

levels.<br />

• Ground faults may have significant values <strong>of</strong> Rf. The<br />

tower footing resistances or foreign object resistances<br />

can have large values.<br />

• Phase faults in transmission networks will most likely<br />

have a small arc resistance.<br />

When discussing protective distance relaying for<br />

transmission lines, it is <strong>of</strong> interest to understand the relay<br />

impedance characteristics and schemes used. Per the<br />

discussion above, ground distance relaying for short lines,<br />

which can be complicated, benefits from the use <strong>of</strong> a<br />

quadrilateral characteristic because ground faults involve more<br />

than the arc resistance. Phase distance relaying, on the other<br />

hand, detects faults where only the arc resistance is involved,<br />

and therefore the complications <strong>of</strong> a quadrilateral element are<br />

not generally required. For these reasons, protective relaying<br />

distance schemes that implement mho phase distance<br />

algorithms to detect phase faults and a combination <strong>of</strong> mho<br />

and quadrilateral ground distance elements to detect ground<br />

faults are justified.<br />

For the majority <strong>of</strong> transmission line applications, from<br />

subtransmission to EHV voltage levels, the mho phase<br />

element and mho quadrilateral ground distance scheme have<br />

proven to be adequate. Extremely short lines may be a<br />

challenge to this scheme. Zero-sequence and negativesequence<br />

directional overcurrent elements have proven to be<br />

the solution for distance element limitations for short lines.<br />

66 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


3<br />

C. Short Line Applications<br />

A short transmission line will generally have lowimpedance<br />

and short length values. On an R-X diagram, like<br />

the one shown in Fig. 2, the line impedance is electrically very<br />

far from the expected maximum load. For some applications,<br />

the line impedance reach (Zset) values challenge the<br />

measurement accuracies <strong>of</strong> the relay itself.<br />

Even for a ground fault with no arc resistance (Raφg equals<br />

zero), the Rf component will have the value <strong>of</strong> the tower<br />

footing resistance, as discussed previously. Mho ground<br />

elements have an intrinsic ability to expand and accommodate<br />

more Rf. This expansion is proportional to the source<br />

impedance (Zs), as shown in Fig. 2 [9]. However, if the tower<br />

footing resistances are in the range <strong>of</strong> the line impedances,<br />

which add to Rf, the mho element will have difficulty<br />

detecting faults even with no arc resistance. The situation is<br />

negatively amplified if the source behind the relay is very<br />

strong—implying a very small Zs.<br />

D. Directional Overcurrent<br />

Directional overcurrent protection is a more sensitive fault<br />

detecting technique than any type <strong>of</strong> distance element [1][10].<br />

The reach <strong>of</strong> these elements varies with the source impedance<br />

<strong>of</strong> a transmission network. Ground directional elements are<br />

polarized with zero-sequence or negative-sequence voltage.<br />

Negative-sequence polarization is also used for phase<br />

directional overcurrent protection. Other phase directional<br />

schemes are also possible.<br />

In line protection schemes, directional overcurrent is used<br />

as a backup scheme for pilot channel loss.<br />

Directional comparison pilot relaying schemes compare the<br />

direction to the fault between two or more terminals. It is<br />

recommended to include directional overcurrent elements (67)<br />

to complement the traditional distance elements (21), as<br />

illustrated in Fig. 3.<br />

jX<br />

Load<br />

Zset<br />

Rf<br />

Fig. 3.<br />

Directional comparison with directional overcurrent elements<br />

Zs<br />

R<br />

Pilot schemes for ground directional overcurrent, as shown<br />

in Fig. 3, will make up for any lack <strong>of</strong> sensitivity <strong>of</strong> mho<br />

elements for short lines. In fact, greater sensitivity is achieved<br />

by using directional overcurrent elements in the scheme,<br />

regardless <strong>of</strong> the types <strong>of</strong> line and distance elements.<br />

Fig. 2.<br />

Short line apparent impedance<br />

Quadrilateral ground distance elements can provide a larger<br />

margin to accommodate Rf. These elements are better suited to<br />

protect short lines. There are some limitations in the amount<br />

<strong>of</strong> Rf that they can accommodate (see Section IV).<br />

Nevertheless, their performance is better than that <strong>of</strong> a mho<br />

circle.<br />

The situation for phase fault detection is similar to that <strong>of</strong><br />

ground fault detection in short line applications. If the<br />

expected arc resistance is approximately the same magnitude<br />

as the transmission line impedance, the mho phase circle will<br />

experience problems detecting the fault. In significantly short<br />

line applications, quadrilateral phase distance elements<br />

provide notably better coverage than a mho phase element.<br />

Nevertheless, it is accepted that directional overcurrent<br />

elements are the most sensitive fault detecting elements and<br />

should be included in pilot relaying schemes [1][3].<br />

E. Fault Resistance on the Apparent Impedance Plane<br />

Relay engineers use the apparent impedance plane to<br />

analyze distance element performance during load, fault, and<br />

power oscillation conditions, either with mho or quadrilateral<br />

elements. In this plane, we can represent the apparent<br />

impedance for line faults with different values <strong>of</strong> Rf and line<br />

loading conditions. Fig. 4 shows the system that we used to<br />

calculate the apparent impedance for phase-to-ground faults at<br />

85 percent from the sending end.<br />

Fig. 4. <strong>Power</strong> system parameters and operating conditions to analyze the<br />

performance <strong>of</strong> distance elements<br />

Adaptive Phase and Ground Quadrilateral Distance Elements | 67


4<br />

Fig. 5 shows the apparent impedance locus for different<br />

loading conditions (δ equal to –20, –10, 0, 10, and 20 degrees)<br />

and all possible values <strong>of</strong> Rf.<br />

forward direction, ZLOAD is on the right side (positive values<br />

<strong>of</strong> resistance) <strong>of</strong> the plane. As Rf starts decreasing, the<br />

apparent impedance describes the locus that Fig. 7 shows.<br />

Notice that with Rf equal to 0, the apparent impedance is<br />

exactly equal to 85 percent <strong>of</strong> the line impedance.<br />

200<br />

4<br />

Reactance (ohms)<br />

100<br />

0<br />

-10<br />

-20<br />

20<br />

-100 0 100<br />

Resistance (ohms)<br />

Fig. 5. Apparent impedance for δ equal to –20, –10, 0, 10, and 20 degrees<br />

while Rf varies from 0 to ∞<br />

Fig. 6 shows that the apparent impedance can cause<br />

distance elements with fixed characteristics over- and<br />

underreach and have limited Rf coverage if the distance<br />

element does not have an adaptive characteristic [11].<br />

20<br />

10<br />

0<br />

Reactance (ohms)<br />

2<br />

0<br />

-2<br />

-4<br />

Rf = 0<br />

Rf = ∞<br />

ZLOAD<br />

0 10 20 30 40<br />

Resistance (ohms)<br />

Fig. 7. Apparent impedance locus for load in the forward direction (δ equal<br />

to 10 degrees)<br />

Fig. 8 shows the impedance locus for incoming load flow<br />

(δ equal to –10 degrees). This apparent impedance makes it a<br />

challenge for the distance elements to detect large values <strong>of</strong> Rf<br />

and avoid element underreach.<br />

150<br />

10<br />

Reactance (ohms)<br />

0<br />

Reactance (ohms)<br />

100<br />

50<br />

Rf = ∞<br />

-10<br />

-10 0 10 20<br />

Resistance (ohms)}<br />

Fig. 6. Apparent impedance can cause distance element over- and<br />

underreach and have limited Rf coverage<br />

Fig. 7 shows the impedance locus for load flow in the<br />

forward direction (δ equal to 10 degrees). In this case, the<br />

remote-end voltage VR equals 0.98 pu. Regardless <strong>of</strong> the<br />

impedance loop measurement (ground fault loop or phase fault<br />

loop), the apparent impedance starts at a load value, ZLOAD,<br />

that corresponds to Rf equals ∞. For active power flow in the<br />

0<br />

ZLOAD<br />

Rf = 0<br />

-50<br />

-100 -80 -60 -40 -20 0 20 40 60 80 100<br />

Resistance (ohms)<br />

Fig. 8. Apparent impedance locus for incoming load flow (δ equal to<br />

–10 degrees)<br />

68 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


5<br />

II. QUADRILATERAL DISTANCE ELEMENTS<br />

Mho distance elements describe a natural and smooth<br />

curvature on the impedance plane. The shape is the result <strong>of</strong> a<br />

phase comparison <strong>of</strong> two quantities that yield the familiar<br />

circle on the apparent impedance plane [9]. Quadrilateral<br />

distance elements are not as straightforward. Combining<br />

distance elements has allowed designers to create all types <strong>of</strong><br />

shapes and polygonal characteristics.<br />

An impedance function with a quadrilateral characteristic<br />

requires the implementation <strong>of</strong> the following:<br />

• A directional element<br />

• A reactance element<br />

• Two left and right blinder resistance calculations<br />

Fig. 9 illustrates a typical quadrilateral element composed<br />

<strong>of</strong> three distance elements. The element that determines the<br />

impedance reach is the reactance element X. The element that<br />

determines the resistive coverage for faults is the right<br />

resistance element Rright. The element that limits the<br />

coverage for reverse flowing load is the left resistance element<br />

Rleft. A directional check keeps the unit detecting faults in the<br />

forward direction only.<br />

Fig. 9.<br />

Components <strong>of</strong> a quadrilateral distance element<br />

The setting <strong>of</strong> the reach on the line impedance angle locus<br />

is denoted by Zset in Fig. 9. It is not a setting on the X axis but<br />

is the reach on the line impedance. We will show that this<br />

setting is the pivot point <strong>of</strong> the line impedance reach. The Rset<br />

setting is the resistive <strong>of</strong>fset from the origin. A line parallel to<br />

the line impedance is shown in Fig. 9.<br />

The impedance lines in Fig. 9 are straight lines for practical<br />

purposes. The theory, however, shows that these lines are<br />

infinite radius circles [9]. The polarizing quantity for creating<br />

these large circles is the measured current at the relay location.<br />

A. Adaptive Reactance Element<br />

Several protective relaying publications report that serious<br />

overreach problems are experienced by nonadaptive reactance<br />

elements because <strong>of</strong> forward load flow and Rf [1][2][11]. If<br />

the reactance element in a quadrilateral characteristic is not<br />

designed to accommodate the situation shown in Fig. 10, an<br />

external fault with Rf may enter the operating area. The<br />

intrinsic curvature and beneficial shift <strong>of</strong> the mho circle are<br />

sufficient to overcome this problem. However, reactance lines<br />

need to be designed to accommodate this issue.<br />

Fig. 10.<br />

jX<br />

The reactance and mho elements adapt to load conditions<br />

Fig. 10 shows the desired behavior <strong>of</strong> the reactance line for<br />

forward load flow. A tilt in the shown direction is required.<br />

Several techniques have been proposed for this purpose,<br />

including a fixed characteristic tilt and the use <strong>of</strong> prefault load.<br />

Interestingly, an infinite diameter mho circle provides the<br />

same tilt as a regular mho circle, and the reactive line becomes<br />

adaptive [9]. The proper polarizing current is the negativesequence<br />

component [12]. The homogeneity <strong>of</strong> the negativesequence<br />

network and the closer proximity <strong>of</strong> the I2 angle to<br />

the fault current (If) angle makes the I2 current an ideal<br />

polarizing quantity.<br />

To obtain the desired reactance characteristic for the AG<br />

loop, the following two quantities can be compared with a<br />

90-degree phase comparator:<br />

S1 = VRA − Zset(IRA + k0 3I0)<br />

(2)<br />

( ) jT<br />

S2 = j IR2 e<br />

(3)<br />

Equations (4) and (5) define the resulting a and b vectors<br />

used to plot the reactance element characteristic [9].<br />

a = Zset<br />

(4)<br />

⎡<br />

o ⎛ IA1 IA0 Zset0 ⎞⎤<br />

− j ⎢90 − T+ ang ⎜1+ +<br />

⎟⎥<br />

⎣ ⎝ IA2 IA2 Zset ⎠⎦<br />

b = ∞ e<br />

(5)<br />

Zset0 − Zset1<br />

k0 = (6)<br />

3• Zset1<br />

where:<br />

k0 is the zero-sequence compensating factor.<br />

Zset0 is the zero-sequence impedance reach derived from<br />

k0 and Zset.<br />

R<br />

Adaptive Phase and Ground Quadrilateral Distance Elements | 69


6<br />

Fig. 11 shows the adaptive behavior <strong>of</strong> the reactance line<br />

derived from (4) and (5). Calculating a proper homogeneity<br />

angle tilt (denoted by T in the equation), the unit ensures<br />

correct reach regardless <strong>of</strong> the direction <strong>of</strong> the load flow.<br />

Similarly, Equations (13) and (14) define the adaptive<br />

phase resistance element for phase faults.<br />

S1 = (VRB − VRC) − Rset(IRB − IRC) (13)<br />

j L1<br />

S2 = (IRB2 − IRC2) e θ<br />

(14)<br />

jX<br />

ZL<br />

Fig. 11.<br />

Adaptive ground reactance element characteristic<br />

For ground distance elements, I0 is another choice for<br />

polarizing the reactance element. This option is acceptable if<br />

the homogeneity factor, T, for the zero-sequence impedances<br />

is known.<br />

For phase distance elements, using the negative-sequence<br />

current is also an option.<br />

S1 = (VRB − VRC) − Zset(IRB − IRC) (7)<br />

jT<br />

S2 = j(IRB2 − IRC2)e<br />

(8)<br />

The resulting a and b vectors are shown in (9) and (10).<br />

a = Zset<br />

(9)<br />

⎛ ⎛ IB1− IC1 ⎞⎞<br />

− j ⎜90− T+ ang ⎜1+<br />

⎟⎟<br />

⎝ ⎝ IB2 − IC2 ⎠⎠<br />

b = ∞ e<br />

(10)<br />

As described in [9], vector b defines the infinite diameter<br />

and the tilt angle, both <strong>of</strong> which are expressed in (5) for the<br />

ground reactance line and (10) for the phase reactance line.<br />

The resulting line is adaptive to the load flow direction, as<br />

shown in Fig. 11. The reactance line adapts properly to load<br />

flow and Rf.<br />

B. Adaptive Resistance Element<br />

Fig. 9 shows that the right resistance element is responsible<br />

for the resistive coverage in a quadrilateral distance element.<br />

This component <strong>of</strong> the quadrilateral distance element should<br />

accommodate and detect as much Rf as possible.<br />

In proposing an adaptive resistance line, it is possible to<br />

make the line static or adaptive as the reactance line. An<br />

adaptive resistive blinder is obtained by defining Rset in (2)<br />

and shifting (3) by (θL1 – 90°), where θL1 is the angle <strong>of</strong> the<br />

positive-sequence line impedance. The benefit shown in<br />

Fig. 12 is a shift <strong>of</strong> the resistance element to the right, which<br />

accommodates faults with forward load flow. Equations (11)<br />

and (12) implement the adaptive ground resistance element.<br />

S1 = VRA − Rset(IRA + k0 3I0) (11)<br />

j L1<br />

S2 = IR2 e θ<br />

(12)<br />

Rset<br />

Fig. 12. Adaptive ground resistance element characteristic<br />

While the use <strong>of</strong> negative-sequence current yields a<br />

beneficial tilt <strong>of</strong> the resistance element for load in the forward<br />

direction, as shown in Fig. 12, the tilt is in the opposite<br />

direction for load in the reverse direction. Therefore, the tilt is<br />

not beneficial under this condition.<br />

Additional polarizing options, like that the alpha<br />

component (I1 + I2) for ground and I1 for phase distance<br />

elements, yield satisfactory tilt behavior for reverse load flow.<br />

The reverse load flow behavior is the same.<br />

C. Left Resistance Element<br />

The left resistive line in Fig. 9 is responsible for limiting<br />

the operation <strong>of</strong> the quadrilateral element for reverse load<br />

flow. It does not need to be adaptive. Care has been taken not<br />

to include the origin to ensure satisfactory operation for very<br />

reactive lines.<br />

D. High-Speed Implementation<br />

In many transmission line protection applications, subcycle<br />

operation is required for distance elements. In many relays,<br />

distance elements with mho or quadrilateral characteristics are<br />

available. When the distance elements selected have<br />

quadrilateral characteristics only, the same high-speed<br />

requirement is applicable for faults with low-resistance value.<br />

In order to obtain subcycle operation with quadrilateral<br />

elements, the same dual-filter concept presented in [14] for<br />

mho elements is used here. The basic principle is to process<br />

the same distance function twice, using two types <strong>of</strong> voltage<br />

and current phasors: the function is processed first using halfcycle<br />

(high-speed) filter phasors and a second time with fullcycle<br />

(conventional) filter phasors. The final function state is<br />

obtained by the logical OR operation from the two processes.<br />

For single-pole tripping applications, these three ground<br />

distance elements (AG, BG, and CG) need to be supervised<br />

with a faulted phase selection function.<br />

R<br />

70 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


7<br />

For the purpose <strong>of</strong> implementing the directional element<br />

and the faulted phase selection for the high-speed part <strong>of</strong> the<br />

quadrilateral function, the algorithm described in [14] and [15]<br />

uses a function known as high-speed directional and fault type<br />

selection (HSD-FTS). It processes signals using half-cycle<br />

filters and superimposed quantities to provide the<br />

14 directional signals listed in Table I.<br />

TABLE I<br />

HIGH-SPEED DIRECTIONAL SIGNALS<br />

Signal<br />

HSD-AGF, HSD-AGR<br />

HSD-BGF, HSD-BGR<br />

HSD-CGF, HSD-CGR<br />

HSD-ABF, HSD-ABR<br />

HSD-BCF, HSD-BCR<br />

HSD-CAF, HSD-CAR<br />

HSD-ABCF, HSD-ABCR<br />

Fault Description<br />

Forward, reverse AG<br />

Forward, reverse BG<br />

Forward, reverse CG<br />

Forward, reverse AB<br />

Forward, reverse BC<br />

Forward, reverse CA<br />

Forward, reverse ABC<br />

Because the HSD-FTS signals are derived from<br />

incremental currents and voltages, they will be available only<br />

for 2 cycles following the inception <strong>of</strong> a fault. Consequently,<br />

the high-speed quadrilateral signals are available for the same<br />

interval <strong>of</strong> time following the detection <strong>of</strong> a fault.<br />

For the reactance element, the high-speed part <strong>of</strong> the<br />

quadrilateral characteristic implementation uses the same<br />

equations for the ground elements as the conventional<br />

counterpart uses with polarization based on negative- or zerosequence<br />

current. During a pole open, the polarization by the<br />

sequence current (negative or zero) is replaced by the<br />

incremental impedance loop current so that the ground<br />

elements remain operational for single-pole tripping<br />

applications.<br />

For the phase elements, polarization is based on the loopimpedance<br />

incremental current so that phase faults and singlepole<br />

tripping applications are automatically covered.<br />

For the two resistance blinder calculations, the equations<br />

are identical to their conventional counterpart so that the<br />

steady-state resistance reach will be identical.<br />

With the high-speed quadrilateral elements, reactance and<br />

resistance blinder calculations use a half-cycle filtering system<br />

to obtain fast operation.<br />

The logic for an A-phase-to-ground fault is presented in<br />

Fig. 13. Similar logic is used for the two other ground fault<br />

elements and the phase elements.<br />

Fig. 13.<br />

faults<br />

High-speed quadrilateral characteristic logic for A-phase-to-ground<br />

To illustrate the parallel operation <strong>of</strong> the high-speed and<br />

conventional quadrilateral elements, an A-phase-to-ground<br />

fault is staged at 33 percent <strong>of</strong> the line length <strong>of</strong> the highvoltage<br />

transmission line in the power network <strong>of</strong> Fig. 4. The<br />

impedance reach is set to 85 percent <strong>of</strong> ZL1. The fault is<br />

staged at 100 milliseconds <strong>of</strong> the EMTP (Electromagnetic<br />

Transients Program) simulation.<br />

Fig. 14 shows the distance to the fault calculations <strong>of</strong> the<br />

two reactance elements (high-speed and conventional) for Rf<br />

equal to 0 ohms. The high-speed element operates in<br />

12.5 milliseconds, and the conventional element operates in<br />

21 milliseconds.<br />

Distance to Fault (pu)<br />

Fig. 14. High-speed and conventional distance element calculations for a<br />

0-ohm, A-phase-to-ground fault at 33 percent <strong>of</strong> the line length<br />

Adaptive Phase and Ground Quadrilateral Distance Elements | 71


8<br />

Fig. 15 depicts the same experiment but with a primary Rf<br />

equal to 50 ohms. The high-speed element has an operating<br />

time <strong>of</strong> 14.5 milliseconds, whereas the conventional element<br />

has an operating time <strong>of</strong> 25 milliseconds.<br />

1<br />

0.8<br />

High-speed distance calculation<br />

Fig. 16 illustrates the negative-sequence network <strong>of</strong> a<br />

simple transmission line and the respective source impedances<br />

at both terminals. If possible, this two-source network should<br />

be evaluated. If the system is slightly more complex (e.g.,<br />

parallel lines), a short-circuit program can provide the IF2 and<br />

IR2 currents. The calculation should be done for a fault at the<br />

reach <strong>of</strong> the Zone 1, where m is approximately 80 percent.<br />

Distance to Fault (pu)<br />

0.6<br />

0.4<br />

Conventional distance calculation<br />

0.2<br />

Conventional trip<br />

High-speed trip<br />

0<br />

0.1 0.11 0.12 0.13 0.14 0.15 0.16<br />

Time (s)<br />

Fig. 15. High-speed and conventional distance element calculations for a<br />

50-ohm, A-phase-to-ground fault at 33 percent <strong>of</strong> the line length<br />

As a general rule, the quadrilateral high-speed logic will<br />

send an output signal a half cycle before the conventional<br />

logic. This corresponds most <strong>of</strong> the time to an overall subcycle<br />

operation for low Rf values. As illustrated in the two examples<br />

above, as Rf increases, both the fault current and the voltage<br />

dip will be reduced. Under these circumstances, the operation<br />

times <strong>of</strong> the high-speed and conventional quadrilateral<br />

elements will increase so that overall operation times close to<br />

or above 1 cycle will be more typical for high-resistance<br />

faults.<br />

III. QUADRILATERAL DISTANCE ELEMENT APPLICATION<br />

A. Homogeneity Calculation<br />

The reactive line in a quadrilateral distance element can be<br />

polarized with either negative-sequence (IR2) or zerosequence<br />

(IR0) current to properly adapt to load flow, as<br />

shown in Fig. 11. Polarizing with these currents makes the<br />

line adaptive and less susceptible to overreach. A check is<br />

needed, however, to ensure effective Zone 1 quadrilateral<br />

ground and phase distance element behavior [13]. This check<br />

is for the homogeneity <strong>of</strong> the negative-sequence impedances<br />

(or zero-sequence impedances, if zero-sequence polarization<br />

has been used).<br />

In a ground fault or asymmetrical phase fault, the total fault<br />

current always lags the source voltages. This fault current, IF,<br />

is the perfect polarizing current. It is in the same direction<br />

regardless <strong>of</strong> the type <strong>of</strong> fault (same angle but with different<br />

magnitude). Because the IF current is not measurable, the<br />

measured currents at the relay location are the only ones<br />

available. The negative-sequence current is an option for<br />

polarizing the reactance line <strong>of</strong> the quadrilateral element. The<br />

protective relay is measuring the local IR2 (negative-sequence<br />

current). The IF2 current is the proper current to use.<br />

Fig. 16.<br />

Two-source negative-sequence network<br />

The variable T is the homogeneity factor, and it is the angle<br />

difference between the fault current and the current measured<br />

at the relay location. Reference [12] illustrates the evaluation<br />

<strong>of</strong> this factor, which is the following current divider<br />

expression:<br />

⎛ IF2 ⎞ ⎡ ZS1 + ZL1 + ZR1 ⎤<br />

T = arg ⎜ ⎟ = arg ⎢ ⎥ (15)<br />

⎝ IR2 ⎠ ⎣ ZR1 + (1 − m) ZL1 ⎦<br />

The angle T in (15) adjusts the measured IR2 current to the<br />

angle <strong>of</strong> the fault current IF2. It is used in (3) and (8) to<br />

properly polarize the reactance line <strong>of</strong> the quadrilateral<br />

element.<br />

When the ground quadrilateral element is polarized with<br />

zero-sequence current (IR0), use a similar expression to<br />

calculate T (15), except that the currents and impedances are<br />

zero sequence.<br />

Equation (15) also provides some extra information<br />

regarding the homogeneity <strong>of</strong> the sequence network. For most<br />

transmission networks, the impedance angles in the negativesequence<br />

network are very similar. Evaluating (15) yields a<br />

small angle, usually in the range <strong>of</strong> ±5 degrees. On the other<br />

hand, in the zero-sequence network, the homogeneity angle<br />

varies considerably more.<br />

In (3) and (8), the reactance line is effectively tilted by the<br />

T angle.<br />

B. Load Encroachment<br />

The quadrilateral distance elements discussed in this paper<br />

are inherently immune to load encroachment. The reactive line<br />

that defines the reach is polarized with negative-sequence<br />

currents, as shown in (3) and (8). The phase and ground<br />

reactive lines start their computation when there is a fault<br />

condition that implies an unbalance <strong>of</strong> (I2/I1) or (I0/I1) greater<br />

than the natural unbalance <strong>of</strong> the system, which is less than<br />

10 percent.<br />

In a full protection scheme, however, there should be<br />

provisions to detect three-phase faults. Although rare, this<br />

type <strong>of</strong> fault is possible. It usually is a fault with almost no Rf.<br />

72 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


9<br />

The three-phase fault detection element is obtained by<br />

using current self-polarization. For example, the BC loop<br />

would be polarized with:<br />

predict the impedance trajectory, and stability studies may be<br />

required.<br />

S2 = j(IRB − IRC)e − jT<br />

(16)<br />

To avoid overreach because <strong>of</strong> forward flowing load, the<br />

setting T in degrees is a tilt, most likely downward, for the<br />

reactive line. The resistive reach is polarized with positivesequence<br />

current.<br />

The three-phase quadrilateral element just described is set<br />

with the same reach as the phase-to-phase distance elements.<br />

It does require certain load considerations to avoid load<br />

encroachment.<br />

If the transmission line is long and the resistive setting<br />

chosen conflicts with load, a load encroachment element is<br />

required. This element should clearly define the load area in<br />

the forward load flow direction. Fig. 17 illustrates a traditional<br />

and already widely used load-encroachment logic<br />

characteristic. The operating point <strong>of</strong> the load impedance in<br />

this region will clearly identify load conditions and prevent<br />

the three-phase fault detection algorithm from operating.<br />

Fig. 17.<br />

jX<br />

Load<br />

Load encroachment for quadrilateral three-phase distance elements<br />

C. Out <strong>of</strong> Step<br />

Much <strong>of</strong> the theory and discussion in literature on out-<strong>of</strong>step<br />

detection can be applied to quadrilateral distance<br />

elements [16][17]. When power flows are oscillating in a<br />

power system, the apparent impedances measured by the<br />

distance elements describe a trajectory on the R-X plane.<br />

These oscillations can be caused by angular instability or<br />

simply switching lines in or out [17]. If the oscillations are<br />

contained within a maximum oscillation envelope and are<br />

damped over time, the power swings are considered stable. On<br />

the other hand, if the power swings are not damped over time,<br />

the power swings are said to be unstable.<br />

On the R-X diagram shown in Fig. 18, a stable power<br />

swing impedance trajectory is contained on the right side (or<br />

the left side for reverse power flow) and eventually rests on a<br />

new load-impedance operating point. An unstable power<br />

swing, in contrast, will show a trajectory that crosses the plane<br />

from left to right (or right to left). Theoretically, and assuming<br />

the simple two-source network shown in Fig. 18, the unstable<br />

power swing will cross the electrical center <strong>of</strong> the system<br />

when the angle’s difference between the two source voltages<br />

is close to 180 degrees apart. Unless the power system can be<br />

reduced to a two-source model, it is not a simple matter to<br />

R<br />

Fig. 18.<br />

Traditional dual-zone out-<strong>of</strong>-step characteristic<br />

During power system oscillations, stability requirements<br />

demand that transmission lines remain in the power system.<br />

Tripping transmission lines unnecessarily jeopardizes the<br />

stability <strong>of</strong> the power system. It is therefore necessary to<br />

ensure that unstable trajectories on the R-X diagram entering<br />

distance element characteristics (shown in Fig. 18) do not<br />

unnecessarily trip the transmission line. However, some<br />

applications require tripping transmission lines in a controlled<br />

manner.<br />

Out-<strong>of</strong>-step detection techniques traditionally take<br />

advantage <strong>of</strong> the slower speed <strong>of</strong> the apparent impedance<br />

trajectory on the R-X diagram for power swing conditions.<br />

The trajectory <strong>of</strong> the operating point changes from load to<br />

fault almost instantaneously for fault conditions.<br />

Fig. 18 illustrates a traditional scheme comprised <strong>of</strong> two<br />

zones. If the inner zone operates after a set time delay (2 to<br />

5 cycles), an out-<strong>of</strong>-step condition is detected. If the trajectory<br />

is due to a power system fault, both zones will operate within<br />

a short time window.<br />

There are several philosophies to follow when setting the<br />

parameters <strong>of</strong> this scheme [17]. Some <strong>of</strong> the most important<br />

considerations are:<br />

• The inner zone should not operate for stable swings.<br />

As shown in Fig. 18, a stable swing eventually returns<br />

to the load impedance.<br />

• The outer zone should not include any possible load<br />

impedance. If load is included by the outer zone, there<br />

is a risk <strong>of</strong> incorrectly declaring a power swing<br />

condition.<br />

• The distance from the inner to the outer zone on the<br />

impedance plane should be made as wide as possible<br />

to allow the detection <strong>of</strong> the power swing condition.<br />

• The inner zone should not include any distance<br />

element zone that is to be blocked. For long line<br />

applications, achieving this goal for all distance zones<br />

may not be possible. We can place the inner zone<br />

across part <strong>of</strong> the distance element characteristic. This<br />

will effectively cut part <strong>of</strong> the characteristic.<br />

Adaptive Phase and Ground Quadrilateral Distance Elements | 73


10<br />

Fig. 18 illustrates some <strong>of</strong> the these considerations. Short<br />

lines present sufficient margin to accommodate the inner and<br />

outer zones together with any type <strong>of</strong> distance element, such<br />

as a quadrilateral distance unit, following the above<br />

guidelines. Long transmission lines, however, may not allow<br />

sufficient margin. Engineering judgment should be used to set<br />

the inner and outer zones, as well as the resistive reach <strong>of</strong> the<br />

quadrilateral element.<br />

When determining the setting parameters, it may be very<br />

difficult to cover all possible scenarios <strong>of</strong> instability with a<br />

simple two-source model. Therefore, transient studies will be<br />

needed to understand the effectiveness <strong>of</strong> the scheme in<br />

Fig. 18.<br />

Recently, a power swing detection algorithm was proposed<br />

that requires little information from the user [18]. This<br />

algorithm will detect and declare a power swing based on the<br />

estimation <strong>of</strong> the swing center voltage (SCV), which is the<br />

voltage at the electrical center <strong>of</strong> a two-source model. This<br />

voltage can be estimated with local measurements and its<br />

behavior used to detect an out-<strong>of</strong>-step condition. The<br />

advantage <strong>of</strong> this methodology is that no network information<br />

is required.<br />

D. Series Capacitor Applications<br />

It is common to apply directional comparison relaying<br />

systems in the protection <strong>of</strong> series-compensated transmission<br />

lines. Protective relays intended to protect these lines should<br />

be designed to accommodate the changing measured<br />

impedance (because <strong>of</strong> the MOVs [metal oxide varistors] and<br />

spark gaps in parallel with the capacitor bank) and<br />

subsynchronous voltages and currents that are characteristic <strong>of</strong><br />

series capacitor-compensated systems [19]. Moreover,<br />

protective relaying systems located in adjacent lines should<br />

reliably determine the direction to a fault.<br />

For distance elements that are polarized with voltage, like<br />

mho distance elements, the voltage inversion because <strong>of</strong> the<br />

series capacitor is properly handled with memory voltage<br />

[19][20]. Moreover, directional elements determine the correct<br />

direction to the fault [21].<br />

Identifying the fault direction is important to keep the<br />

reactance and resistance lines <strong>of</strong> the quadrilateral distance<br />

element from operating improperly. An impedance-based<br />

negative-sequence polarized directional element (or an<br />

alternate zero-sequence polarized element for ground faults)<br />

will properly determine the direction to the fault, unless a<br />

current inversion is present. Depending on the location <strong>of</strong> the<br />

capacitor bank and the location <strong>of</strong> the voltage transformers<br />

(VTs), suggested settings for the directional thresholds (Z2F<br />

and Z2R) can be found in [20] and [21]. For the impedances<br />

<strong>of</strong> the compensated system in Fig. 19, the directional element<br />

threshold Z2F should be set to:<br />

(ZL1 − XC)<br />

Z2F ≤ (17)<br />

2<br />

Setting this threshold as close to the origin as possible will<br />

ensure proper directional determination, unless a current<br />

reversal is possible in the power system.<br />

jX<br />

–jXC<br />

ZL1<br />

Fig. 19. Series capacitor applications<br />

Uncompensated<br />

Compensated<br />

In Fig. 19, the perspective <strong>of</strong> a long line is shown. Seriescompensated<br />

lines are long lines that require compensation to<br />

transfer more power. There are no short lines compensated<br />

with series capacitors. Also, in the vicinity <strong>of</strong> a series<br />

capacitor installation, subsynchronous oscillations <strong>of</strong> the<br />

voltages and currents are possible [19][20][21]. While the<br />

filtering in protective relays is very good at eliminating highfrequency<br />

components, the filtering is not efficient at<br />

eliminating lower frequencies. These subsynchronous<br />

transients, shown as impedance oscillations on the apparent<br />

impedance plane, eventually converge on the true apparent<br />

impedance, as illustrated in Fig. 20. This figure also shows<br />

that distance element overreach is a possibility.<br />

jX<br />

Z1L<br />

–jXc<br />

Fig. 20. Subharmonic frequency transients can cause distance elements to<br />

overreach<br />

R<br />

R<br />

74 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


11<br />

Zone 1 distance elements should account for the above<br />

phenomena by reducing the reach [20][21]. A good suggestion<br />

is to set the reach <strong>of</strong> the reactive line to half <strong>of</strong> the<br />

compensated line impedance [20]. On the other hand,<br />

protective relays can have an automatic reach adjustment<br />

based on a measured apparent impedance compared to a<br />

theoretically calculated value [14][21]. This way, the reach is<br />

automatically reduced to half <strong>of</strong> the compensated line<br />

impedance when transients are detected. The resistive reach<br />

should follow the recommendations for a long line (e.g., Rset<br />

equal to one-half Zset).<br />

The presence <strong>of</strong> the series capacitor in the power system<br />

modifies the homogeneity <strong>of</strong> the negative- and zero-sequence<br />

impedances. Therefore, when adjusting the homogeneity<br />

factor T, described in (18) and (19), the capacitor impedance<br />

should be considered. When using a negative-sequence current<br />

polarized reactance element:<br />

⎛ IF2 ⎞ ⎡ ZS1 + ZL1 − XC + ZR1 ⎤<br />

T = arg⎜ ⎟ = arg<br />

IR2<br />

⎢<br />

ZR1 + (1 − m)(ZL1 − XC)<br />

⎥ (18)<br />

⎝ ⎠ ⎣ ⎦<br />

And when using a zero-sequence polarized reactance line:<br />

⎛ IF0 ⎞ ⎡ ZS0 + ZL0 − XC + ZR0 ⎤<br />

T = arg ⎜ ⎟ = arg<br />

IR0<br />

⎢<br />

ZR0 + (1 − m)(ZL0 − XC)<br />

⎥ (19)<br />

⎝ ⎠ ⎣ ⎦<br />

Notice that the zero- and negative-sequence impedance <strong>of</strong> a<br />

series capacitor are the same as the positive-sequence<br />

impedance.<br />

Equation (18) for the uncompensated line should also be<br />

evaluated. The minimum calculated T value (most negative)<br />

should be used.<br />

When applying any protective relaying scheme to seriescompensated<br />

lines, transient simulation and testing are<br />

recommended [19][21]. This step ensures dependability and<br />

confirms proposed settings.<br />

E. Single-Pole Trip Applications<br />

In transmission line protection, it is common to use singlepole<br />

trip schemes. The scheme trips the faulted phase only for<br />

a single-line-to-ground fault. Once the pole is open, the other<br />

two phases are still conducting power, and the system is<br />

capable <strong>of</strong> remaining synchronized. During the open-pole<br />

interval, it is expected that the arc deionizes. After the openpole<br />

interval, a reclosing command is sent to the breaker.<br />

Current polarization with negative-sequence current (I2) or<br />

zero-sequence current (I0) is not reliable during the open-pole<br />

interval. The open pole makes the power system unbalanced,<br />

causing negative- and zero-sequence currents to flow. The<br />

consequence to distance elements polarized with sequence<br />

component currents, as in (3) and (8), is that the polarization<br />

becomes unreliable. Depending on the load flow direction, I2<br />

and I0 will have different directions. Fortunately, there are<br />

other distance elements that will reliably operate during an<br />

open-pole condition [14]. The positive-sequence voltagepolarized<br />

mho element is stable during open-pole intervals and<br />

will reliably detect power system faults during this condition.<br />

In a practical scheme, the phase and ground quadrilateral<br />

elements should be disabled when an open-pole condition is<br />

detected. The high-speed quadrilateral distance element is<br />

implemented with incremental quantities and does not need to<br />

be disabled during the open-pole interval.<br />

IV. SETTING THE QUADRILATERAL DISTANCE ELEMENT<br />

Consider the A-phase-to-ground fault circuit <strong>of</strong> Fig. 4.<br />

Equation (20) determines the apparent impedance (Zapp) that<br />

the relay installed at the left side <strong>of</strong> the line measures as a<br />

function <strong>of</strong> fault voltages and currents. Equation (21)<br />

determines Zapp as a function <strong>of</strong> Rf and fault location m.<br />

VA<br />

Zapp =<br />

IA + k0 • IR<br />

(20)<br />

Zapp = m • ZL1 + KR • Rf<br />

(21)<br />

In (21), KR is a factor that depends upon the positive- and<br />

zero-sequence current distribution factors (C1 and C0) and is<br />

equal to:<br />

KR 3<br />

=<br />

(22)<br />

2 • C1 + C0(1 + 3• k0)<br />

C1 and C0 are equal to:<br />

(1 − m) • ZL1 + ZR1<br />

C1 =<br />

(23)<br />

ZS1 + ZL1 + ZR1<br />

(1 − m) • ZL0 + ZR0<br />

C0 =<br />

(24)<br />

ZS0 + ZL0 + ZR0<br />

k0 is the zero-sequence compensation factor equal to:<br />

ZL0 − ZL1<br />

k0 = (25)<br />

3• ZL1<br />

For no-load conditions (δ equal to 0) and homogeneous<br />

systems, the resistive blinder <strong>of</strong> the adaptive quadrilateral<br />

element will assert for an Rf that satisfies this condition:<br />

Rapp < Rset<br />

(26)<br />

Rapp = Real(KR) • Rf<br />

(27)<br />

where Rset is the resistive reach setting. Alternatively, we can<br />

calculate Rapp using relay voltage and currents for a fault at m<br />

according to (28).<br />

Rapp = Real( Zapp) − m • Real( ZL1)<br />

(28)<br />

Adaptive Phase and Ground Quadrilateral Distance Elements | 75


12<br />

For the system in Fig. 4, Fig. 21 represents the values <strong>of</strong><br />

Real(KR) as a function <strong>of</strong> m with a constant value <strong>of</strong> Rset. The<br />

increasing values <strong>of</strong> Real(KR) indicate that the maximum<br />

detectable Rf at no load decreases as the distance to the fault<br />

increases.<br />

10<br />

9<br />

8<br />

7<br />

Real (KR)<br />

6<br />

5<br />

4<br />

Fig. 22. CT and VT error evaluation for Zone 1<br />

Fig. 23 shows the Rmax pu as a function <strong>of</strong> Zset for θL1<br />

equal to 40, 55, 70, and 75 degrees and θε equal to 2 degrees.<br />

30<br />

3<br />

2<br />

1<br />

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

Fault Location (pu)<br />

Fig. 21. Factor Real(KR) for the system <strong>of</strong> Fig. 4<br />

Another consideration in determining the setting <strong>of</strong> the<br />

resistive coverage involves VT and CT (current transformer)<br />

errors. Reference [22] indicates that a composite angle error in<br />

the measurement θε can be assumed.<br />

A. Zone 1<br />

For a Zone 1 application, the requirement is that Zone 1<br />

never overreaches for any fault at the end <strong>of</strong> the line.<br />

Assuming that for resistive faults at the end <strong>of</strong> the line there is<br />

an angle error θε, the effective path for increasing Rf will tilt<br />

down an extra θε degrees, as shown in Fig. 22. For increasing<br />

Rf, the intersection with the Zone 1 reactive line is the<br />

indication <strong>of</strong> the maximum Rset or Rmax. Using the law <strong>of</strong><br />

sines and trigonometry, Rmax can be expressed as:<br />

sin ( θε + θ L1)<br />

Rmax = • ( 1− Zset _ pu ) • ZL1 (29)<br />

sin θε<br />

( )<br />

Equation (29) defines Rmax, the maximum secure resistive<br />

reach setting for Zone 1, taking into account CT, VT, relay<br />

measurement errors, and θε. Rmax is a function <strong>of</strong> the<br />

impedance reach setting Zset, the positive-sequence line<br />

impedance magnitude |ZL1| and angle θL1, and the total<br />

angular error in radians θε [22].<br />

Maximum Resistive Reach Setting Rmax (pu)<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

Impedance Reach Setting Zset (pu)<br />

Fig. 23. Maximum resistive reach setting as a function <strong>of</strong> the impedance<br />

reach due to measurement errors<br />

A typical Zone 1 impedance reach setting for short lines is<br />

70 percent. For the system in Fig. 4, Zset_zone1 is equal to<br />

1.4 ohms secondary.<br />

With Zset, |ZL1|, θL1, and θε, we can calculate Rmax using<br />

(29) or obtain the pu value Rmax_pu with respect to the total<br />

positive-sequence line impedance from Fig. 23. In this case,<br />

Rmax equals 17.17 ohms secondary, or Rmax_pu equals<br />

8.58 pu.<br />

Additionally, we need to verify that the fault current is<br />

above the maximum relay sensitivity. In this case, the residual<br />

current is 3.0 A secondary, and the relay sensitivity is 0.25 A.<br />

Therefore, the relay can see the fault at 70 percent <strong>of</strong> the line<br />

with Rf equal to 25 ohms primary.<br />

40<br />

55<br />

70<br />

85<br />

76 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


13<br />

Fig. 24 shows the apparent resistance for different Rf<br />

values for a fault at 70 percent <strong>of</strong> the line. Note that with<br />

Rset_zone1 equal to 11.52 ohms, the quadrilateral element can<br />

see 3 ohms secondary or 25 ohms primary.<br />

8<br />

7<br />

—— Rapp (ohms)<br />

Fault Resistance (ohms)<br />

6<br />

5<br />

4<br />

3<br />

2<br />

Rset = 11.52 ohms<br />

δ = 0 degrees<br />

1<br />

Fig. 24.<br />

Apparent resistance for a fault at 70 percent <strong>of</strong> the line<br />

Fig. 25 shows the margin <strong>of</strong> the selected Rset with respect<br />

to Rmax for the selected Zset.<br />

Fig. 25.<br />

Margin <strong>of</strong> Rset at Zset_zone1 equal to 0.7 pu<br />

Fig. 26 shows that the quadrilateral distance element can<br />

see up to 3 ohms for faults at 70 percent <strong>of</strong> the line for Rset<br />

equal to 11.52 ohms.<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

Fault Location (pu)<br />

Fig. 26. Maximum Rf coverage with Rset equal to 11.5 ohms for faults<br />

along the line<br />

The analysis carried out for a phase-to-ground fault can<br />

also be applied for phase-to-phase faults. In these cases, the<br />

factor KR is equal to:<br />

1<br />

KR = (30)<br />

2 • C1<br />

For no-load conditions (δ equal to 0) and homogeneous<br />

systems, the resistive blinder <strong>of</strong> the phase quadrilateral<br />

element will assert for an Rf that satisfies (31) or (32).<br />

Rf<br />

( )<br />

Rset<br />

2 • Real C1 < (31)<br />

⎛ Vϕϕ<br />

⎞<br />

Rapp = Real⎜<br />

⎟ − m • Real ZL1<br />

⎝ Iϕϕ<br />

⎠<br />

( )<br />

(32)<br />

B. Zone 2<br />

When considering overreaching zones, it is important to<br />

determine the maximum underreach and verify that the zone<br />

covers at least the expected Rf. For example, in a Zone 2<br />

application, it is expected that all faults on the line and those<br />

at the remote terminal will be detected. It is a common<br />

practice to set the Zone 2 reach to 120 percent <strong>of</strong> the line<br />

length. However, in certain circumstances, this impedance<br />

reach would not guarantee coverage for faults with Rf, and a<br />

longer reach would be required.<br />

Adaptive Phase and Ground Quadrilateral Distance Elements | 77


14<br />

Following is a conservative approach to set Zone 2 that<br />

guarantees that the overreaching element sees all faults with<br />

specific Rf coverage. Fig. 27 shows the apparent resistance for<br />

a homogeneous system and no-load conditions.<br />

Fig. 28 shows Rset_zone2_pu for θL1 equal to 40, 55, 70,<br />

and 85 degrees and for θε equal to –2 degrees.<br />

Resistive Reach Setting Rset Zone 2 (pu)<br />

15<br />

10<br />

5<br />

40<br />

55<br />

70<br />

85<br />

Fig. 27. Apparent impedance for an end-<strong>of</strong>-line fault considering<br />

measurement errors<br />

From Fig. 27, we can estimate the required resistive reach<br />

Rset_zone2 and impedance reach Zset_zone2 settings<br />

according to (33) and (34) for a desired Rf coverage.<br />

( θL1<br />

+ θε )<br />

( θL1)<br />

sin ( θε )<br />

sin( θL1<br />

+ θε )<br />

sin<br />

− j⋅θε<br />

Rset _ zone2 = • Real(KR) • Rf • e (33)<br />

sin<br />

Zset _ zone2 = ZL1 −<br />

• R set _ zone2<br />

(34)<br />

We can represent Rset_zone2_pu as a function <strong>of</strong> Ppu<br />

according to (35). These values are the normalized values <strong>of</strong><br />

Rset_zone2 and P (see Fig. 27) with respect to |ZL1|.<br />

( θL1<br />

+ θε )<br />

sin ( θε )<br />

sin<br />

Rset _ zone2 _ pu = − • Ppu (35)<br />

0<br />

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5<br />

P (pu)<br />

Fig. 28. The resistive reach setting as a function <strong>of</strong> the impedance reach<br />

setting for several values <strong>of</strong> θL1<br />

For the relay in Fig. 4, we calculate Rset_zone2 for a<br />

desired Rf equal to 3 ohms secondary. Using (33) with θε<br />

equal to –2 degrees and a fault at the end <strong>of</strong> the line, we obtain<br />

Rset_zone2 equal to 28.69 ohms secondary and<br />

Rset_zone2_pu equal to 14.35 pu. From Fig. 28, we obtain the<br />

required value <strong>of</strong> Ppu and the Zone 2 impedance reach<br />

Zset_zone2_pu equal to 1.5 or 150 percent <strong>of</strong> |ZL1|.<br />

V. DISTANCE ELEMENT PERFORMANCE<br />

A. Traditional Distance Element Characteristics<br />

Adaptive quadrilateral phase and ground distance elements<br />

were designed to improve Rf coverage in short line<br />

applications. A previous distance relay included a ground<br />

quadrilateral distance element characteristic with an adaptive<br />

reactance element and two resistance elements that calculate<br />

Rf according to (36) and a ground mho distance characteristic<br />

with an adaptive mho element that calculates the distance to<br />

the fault according to (37) [12]<br />

j⋅θ<br />

L1<br />

( )<br />

*<br />

Im<br />

⎡<br />

V • I • e<br />

⎤<br />

⎢<br />

⎥<br />

R relay1 =<br />

⎣<br />

⎦<br />

⎡ 3<br />

Im<br />

⎢<br />

• I2 I0 • I e<br />

⎣ 2<br />

j⋅θ<br />

L1<br />

( + ) ( ⋅ )<br />

( )<br />

*<br />

Re ⎡V • V1_ mem ⎤<br />

m relay1 =<br />

⎣<br />

⎦<br />

j L1<br />

Re ⎡ ⋅θ<br />

I • e • V1_ mem<br />

⎣<br />

( )<br />

*<br />

*<br />

⎤<br />

⎥<br />

⎦<br />

⎤<br />

⎦<br />

(36)<br />

(37)<br />

78 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


15<br />

B. Adaptive Resistance Element<br />

Fig. 29 shows the quadrilateral distance element<br />

characteristic that uses the adaptive reactance element <strong>of</strong> the<br />

previous design with an adaptive resistive element.<br />

X<br />

Adaptive Reactance Element<br />

setting <strong>of</strong> 11.52 ohms and an impedance reach (Zset) setting <strong>of</strong><br />

120 percent <strong>of</strong> ZL1. The mho distance element has a reach <strong>of</strong><br />

120 percent <strong>of</strong> ZL1. The sensitivity <strong>of</strong> all the distance<br />

elements is 0.05 • Inom.<br />

Fig. 30, Fig. 31, and Fig. 32 show the Rf coverage <strong>of</strong> mho<br />

and quadrilateral distance elements. We observe that the Rf<br />

coverage is severely reduced as the fault approaches the end <strong>of</strong><br />

the line. As expected, the mho element is the one with less Rf<br />

coverage, and the adaptive resistance element has the greatest<br />

Rf coverage, especially for power flow in the forward<br />

direction, δ equal to 10 degrees.<br />

20<br />

18<br />

Mho Distance Element<br />

Quadrilateral Distance Element<br />

Adaptive Resistive Element<br />

Z LINE<br />

Adaptive<br />

Resistive<br />

Element<br />

16<br />

14<br />

Rset = 11.52 ohms<br />

δ = -10 degrees<br />

R<br />

Fault Resistance (ohms)<br />

12<br />

10<br />

8<br />

6<br />

Fig. 29. Quadrilateral distance element characteristic with adaptive<br />

resistance element<br />

The adaptive resistive element calculates Rf according to<br />

(38) and (39), and compares the minimum <strong>of</strong> the two<br />

calculation results against the resistive setting.<br />

j⋅θ<br />

L1<br />

( )<br />

j⋅θ<br />

L1<br />

*<br />

( )<br />

Im<br />

⎡<br />

V • I2 • e<br />

⎢<br />

R2 =<br />

⎣<br />

Im<br />

⎡<br />

I • I2 • e<br />

⎢⎣<br />

Im<br />

⎡<br />

V • I<br />

⎢<br />

Rα<br />

=<br />

⎣<br />

Im<br />

⎡<br />

I • I<br />

⎢⎣<br />

⎤<br />

⎥⎦<br />

⎤<br />

⎥⎦<br />

*<br />

j⋅θ<br />

L1<br />

( α • e )<br />

j⋅θ<br />

L1<br />

*<br />

( α • e )<br />

⎤<br />

⎥⎦<br />

⎤<br />

⎥⎦<br />

*<br />

(38)<br />

(39)<br />

Equation (38) is equivalent to (11) and (12). It uses a<br />

different form <strong>of</strong> phase comparator equation. Equation (39)<br />

uses the alpha component (Iα = I1 + I2).<br />

C. Resistive Coverage<br />

To compare the resistive coverage <strong>of</strong> the traditional<br />

distance elements with the adaptive resistive element, we use<br />

the system in Fig. 4 and perform the tests discussed in the<br />

following sections.<br />

1) Faults at Multiple Locations<br />

We calculate the maximum Rf that the distance elements<br />

can detect for an A-phase-to-ground fault for m values from 0<br />

to 1 and load angles <strong>of</strong> δ equal to –10, 0, and 10 degrees. The<br />

quadrilateral distance elements have a resistive reach (Rset)<br />

4<br />

2<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

Fault Location (pu)<br />

Fig. 30. Rf coverage <strong>of</strong> mho and quadrilateral distance elements for<br />

δ equal to –10 degrees<br />

Fault Resistance (ohms)<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

Mho Distance Element<br />

Quadrilateral Distance Element<br />

Adaptive Resistive Element<br />

Rset = 11.52 ohms<br />

δ = 0 degrees<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

Fault Location (pu)<br />

Fig. 31. Rf coverage <strong>of</strong> mho and quadrilateral distance elements for δ equal<br />

to 0 degrees<br />

Adaptive Phase and Ground Quadrilateral Distance Elements | 79


16<br />

Fault Resistance (ohms)<br />

12<br />

10<br />

8<br />

6<br />

4<br />

Mho Distance Element<br />

Quadrilateral Distance Element<br />

Adaptive Resistive Element<br />

Rset = 11.52 ohms<br />

δ = 10 degrees<br />

D. Adaptive Behavior<br />

A carefully designed quadrilateral characteristic should<br />

have an adaptive reactance line to avoid overreach because <strong>of</strong><br />

load in the forward direction and Rf. Moreover, this paper has<br />

presented the concept <strong>of</strong> an adaptive resistive line that<br />

beneficially tilts to detect more Rf.<br />

Two figures will be used to illustrate the adaptive behavior<br />

<strong>of</strong> the reactive line. Fig. 34 illustrates a ground fault detected<br />

from the terminal with forward load flow. Fig. 35 shows the<br />

same fault, with the same Rf, detected from the other terminal<br />

(i.e., the terminal with reverse direction load flow).<br />

2<br />

0<br />

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1<br />

Fault Location (pu)<br />

Fig. 32. Rf coverage <strong>of</strong> mho and quadrilateral distance elements for δ equal<br />

to 10 degrees<br />

2) Faults at m Equal to 0.7 for Multiple Load Angles<br />

We calculate the Rf coverage for a fault at 70 percent <strong>of</strong> the<br />

line and different load angles (see Fig. 33). The adaptive<br />

resistance element has the highest Rf coverage, while the mho<br />

element has the lowest Rf coverage.<br />

12<br />

10<br />

Adaptive Resistive Element<br />

Mho Distance Element<br />

Quadrilateral Distance Element<br />

Rset = 11.52 ohms<br />

m = 0.7<br />

Fault Resistance (ohms)<br />

8<br />

6<br />

4<br />

Fig. 34. Example <strong>of</strong> a ground fault detected from the forward load flow<br />

direction terminal<br />

2<br />

0<br />

-10 -5 0 5 10<br />

Load Angle (degrees)<br />

Fig. 33.<br />

angles<br />

Rf coverage for faults at 70 percent <strong>of</strong> the line with different load<br />

Fig. 35. Example <strong>of</strong> a ground fault detected from the reverse load flow<br />

direction terminal<br />

80 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


17<br />

Fig. 34 and Fig. 35 illustrate the adaptive behavior <strong>of</strong> the<br />

quadrilateral distance element. Because these figures are for<br />

illustrative purposes, the power system impedances, fault<br />

location, and/or Rf value are not relevant to the discussion.<br />

These two figures simply illustrate the adaptive behavior <strong>of</strong><br />

the reactance and resistive lines.<br />

The fault type is a ground fault; and the ground<br />

quadrilateral element is formed with a reactance line polarized<br />

with negative-sequence current (the preferred polarization).<br />

The two resistance elements are polarized with negative<br />

sequence (I2) and with the alpha component (I1 + I2). The<br />

polarization is what makes the lines adaptive, as explained in<br />

the previous sections.<br />

Fig. 34 and Fig. 35 provide a wealth <strong>of</strong> information about<br />

the behavior <strong>of</strong> these impedance lines, including the<br />

following:<br />

• The reactance element pivots on the line impedance<br />

reach, and that point is fixed. The resistance elements<br />

pivot on the resistive reach, which is a setting.<br />

• The degree at which these lines tilt is determined by<br />

the power system parameters and operating<br />

conditions. These include line impedances, load flow,<br />

and Rf.<br />

• The mho circle and reactance line tilt at the same time<br />

and in the same direction.<br />

• The resistance element trip decision is the OR<br />

combination <strong>of</strong> either resistive line. Their behavior is<br />

dependent on the direction <strong>of</strong> the load flow, and their<br />

operation is complementary to each other.<br />

For the forward load flow terminal in Fig. 34, we conclude:<br />

• The reactance element tilts beneficially in a clockwise<br />

direction. This behavior prevents overreaching<br />

because <strong>of</strong> high-resistance load flow.<br />

• The resistance element polarized with negativesequence<br />

current will adapt to provide better resistive<br />

reach coverage. This resistive line will make the<br />

decision for faults with load in the forward direction.<br />

• The resistance element polarized with the alpha<br />

component current tilts in the opposite direction. The<br />

resistive coverage <strong>of</strong> this characteristic is less effective<br />

compared to the other resistive line.<br />

For the reverse load flow terminal in Fig. 35, we conclude:<br />

• The reactance element moves in the direction that the<br />

apparent impedance locus moves, as shown in Fig. 10.<br />

• The resistance element polarized with negativesequence<br />

current moves in the opposite direction and<br />

with less resistive coverage.<br />

• The resistance element polarized with the alpha<br />

component provides more effective coverage and will<br />

detect the fault.<br />

VI. CONCLUSIONS<br />

<strong>Power</strong> system faults present different values <strong>of</strong> Rf. Ground<br />

faults present a larger value <strong>of</strong> Rf because <strong>of</strong> the arc resistance<br />

and tower footing resistance.<br />

Short line protection applications with distance elements<br />

favor the use <strong>of</strong> quadrilateral distance elements for phase and<br />

ground fault protection. The expected Rf for short lines can be<br />

in the same order <strong>of</strong> magnitude as the impedance <strong>of</strong> the<br />

transmission line.<br />

Rf and power flow have the effect <strong>of</strong> modifying the<br />

apparent impedance measured by the distance element. The<br />

description <strong>of</strong> the Rf influence was plotted for different Rf<br />

values and load flows.<br />

Especially for short lines, quadrilateral distance elements<br />

can detect faults with higher Rf than mho distance elements.<br />

Instrument transformers and relay measurement errors limit<br />

the Rf coverage in short line applications.<br />

An adaptive characteristic for ground and phase<br />

quadrilateral distance elements was presented. The reactance<br />

element, polarized with negative-sequence current, adapts<br />

based on the direction <strong>of</strong> the load flow and prevents<br />

overreaching issues associated with load flow in the forward<br />

direction. Resistance elements are polarized with two<br />

quantities simultaneously. The negative-sequence polarization<br />

has a better Rf coverage for forward load flow. Using ground<br />

distance (alpha component) and phase distance (positivesequence<br />

component) provides better coverage for faults with<br />

reverse load flow. For the resistance elements, running two<br />

polarizations at the same time helps to detect as much Rf as<br />

possible.<br />

A high-speed version <strong>of</strong> the quadrilateral elements<br />

typically improves the speed <strong>of</strong> operation by half a cycle.<br />

These elements are required in applications where high-speed<br />

tripping times are required. These elements provide subcycle<br />

operating speeds and operate reliably during open-pole<br />

conditions.<br />

The paper presented an improved distance element with an<br />

adaptive quadrilateral characteristic that can be part <strong>of</strong> a line<br />

protection relay.<br />

The performance <strong>of</strong> the adaptive quadrilateral distance<br />

element was compared to a previous quadrilateral<br />

implementation, showing the benefits. A graphical illustration<br />

<strong>of</strong> the performance expected from the reactance and resistive<br />

lines was presented with an example in Fig. 34 and Fig. 35.<br />

The Rf coverage <strong>of</strong> the adaptive quadrilateral element<br />

increases for terminals with forward direction load flow.<br />

The phase quadrilateral distance elements presented in this<br />

paper are suitable for any transmission line application, but<br />

because <strong>of</strong> their nature, they fit better in short line<br />

applications. No distance element, however, can provide better<br />

sensitivity and Rf coverage than directional overcurrent<br />

elements in a pilot scheme.<br />

Adaptive Phase and Ground Quadrilateral Distance Elements | 81


18<br />

VII. REFERENCES<br />

[1] S. Ward, “Comparison <strong>of</strong> Quadrilateral and Mho Distance<br />

Characteristic,” proceedings <strong>of</strong> the 26th Annual Western Protective<br />

Relay Conference, Spokane, WA, October 1999.<br />

[2] J. Holbach, V. Vadlamani, and Y. Lu, “Issues and Solutions in Setting a<br />

Quadrilateral Distance Characteristic,” proceedings <strong>of</strong> the 61st Annual<br />

Conference for Protective Relay Engineers, College Station, TX, April<br />

2008.<br />

[3] J. Mooney and J. Peer, “Application Guidelines for Ground Fault<br />

Protection,” proceedings <strong>of</strong> the 24th Annual Western Protective Relay<br />

Conference, Spokane, WA, October 1997.<br />

[4] S. Sebo, “Zero-Sequence Current Distribution Along Transmission<br />

Lines,” IEEE Transactions on Apparatus and Systems, PAS-88, Issue 6,<br />

June 1969.<br />

[5] G. Swift, D. Fedirchuk, and T. Ernst, “Arcing Fault ‘Resistance’ (It<br />

Isn’t),” proceedings <strong>of</strong> the Georgia Tech Fault and Disturbance Analysis<br />

Conference, Atlanta, GA, May 2003.<br />

[6] H. Khodr, A. Menacho e Moura, and V. Miranda, “Optimal Design <strong>of</strong><br />

Grounding System in Transmission Line,” proceedings <strong>of</strong> the<br />

International Conference on Intelligent Systems Applications to <strong>Power</strong><br />

Systems, November 2007.<br />

[7] V. Terzija and H. Koglin, “New Approach to Arc Resistance<br />

Calculation,” IEEE PES Winter Meeting, Vol. 2, 2001.<br />

[8] L. Popovic, “A Digital Fault-Location Algorithm Taking Into Account<br />

the Imaginary Part <strong>of</strong> the Grounding Impedance at the Fault Place,”<br />

IEEE Transactions on <strong>Power</strong> Delivery, Vol. 18, Issue 4, October 2003.<br />

[9] F. Calero, “Distance Elements: Linking Theory With Testing,”<br />

proceedings <strong>of</strong> the 35th Annual Western Protective Relay Conference,<br />

Spokane, WA, October 2008.<br />

[10] J. Roberts, E. O. Schweitzer, III, R. Arora, and E. Poggi, “Limits to the<br />

Sensitivity <strong>of</strong> Ground Directional and Distance Protection,” proceedings<br />

<strong>of</strong> the 1997 Spring Meeting <strong>of</strong> the Pennsylvania Electric Association<br />

Relay Committee, Allentown, PA, May 1997.<br />

[11] J. Roberts, A. Guzmán, and E. O. Schweitzer, III, “Z = V/I Does Not<br />

Make a Distance Relay,” proceedings <strong>of</strong> the 20th Annual Western<br />

Protective Relay Conference, Spokane, WA, October 1993.<br />

[12] E. O Schweitzer, III and J. Roberts, “Distance Relay Element Design,”<br />

proceedings <strong>of</strong> the 46th Annual Conference for Protective Relay<br />

Engineers, College Station, TX, April 1993.<br />

[13] <strong>SEL</strong>-421 Instruction Manual. Available: http://www.selinc.com.<br />

[14] A. Guzmán, J. Mooney, G. Benmouyal, and N. Fischer, “Transmission<br />

Line Protection System for Increasing <strong>Power</strong> System Requirements,”<br />

proceedings <strong>of</strong> the 55th Annual Conference for Protective Relay<br />

Engineers, College Station, TX, April 2002.<br />

[15] G. Benmouyal and J. Roberts, “Superimposed Quantities: Their True<br />

Nature and Application in Relays,” proceedings <strong>of</strong> the 26th Annual<br />

Western Protective Relay Conference, Spokane, WA, October 1999.<br />

[16] D. Tziouvaras and D. Hou, “Out-<strong>of</strong>-Step Protection Fundamentals and<br />

Advancements,” proceedings <strong>of</strong> the 30th Annual Western Protective<br />

Relay Conference, Spokane, WA, October 2003.<br />

[17] J. Mooney and N. Fischer, “Application Guidelines for <strong>Power</strong> Swing<br />

Detection on Transmission Systems,” proceedings <strong>of</strong> the 32nd Annual<br />

Western Protective Relay Conference, Spokane, WA, October 2005.<br />

[18] G. Benmouyal, D. Tziouvaras, and D. Hou, “Zero-Setting <strong>Power</strong>-Swing<br />

Blocking Protection,” proceedings <strong>of</strong> the 31st Annual Western<br />

Protective Relay Conference, Spokane, WA, October 2004.<br />

[19] H. Altuve, J. Mooney, and G. Alexander, “Advances in Series-<br />

Compensated Line Protection,” proceedings <strong>of</strong> the 35th Annual Western<br />

Protective Relay Conference, Spokane, WA, October 2008.<br />

[20] J. Mooney and G. Alexander, “Applying the <strong>SEL</strong>-321 Relay on Series-<br />

Compensated Systems,” <strong>SEL</strong> Application Guide AG2000-11. Available:<br />

http://www.selinc.com.<br />

[21] F. Plumptre, M. Nagpal, X. Chen, and M. Thompson, “Protection <strong>of</strong><br />

EHV Transmission Lines With Series Compensation: BC Hydro’s<br />

Lessons Learned,” proceedings <strong>of</strong> the 62nd Annual Conference for<br />

Protective Relay Engineers, College Station, TX, March 2009.<br />

[22] E. O. Schweitzer, III, K. Behrendt, and T. Lee, “Digital<br />

Communications for <strong>Power</strong> System Protection: Security, Availability<br />

and Speed,” proceedings <strong>of</strong> the 25th Annual Western Protective Relay<br />

Conference, Spokane, WA, October 1998.<br />

VIII. BIOGRAPHIES<br />

Fernando Calero received his BSEE in 1986 from the University <strong>of</strong> Kansas,<br />

his MSEE in 1987 from the University <strong>of</strong> Illinois (Urbana-Champaign), and<br />

his MSEPE in 1989 from the Rensselaer Polytechnic Institute. From 1990 to<br />

1996, he worked in Coral Springs, Florida, for the ABB relay division in the<br />

support, training, testing, and design <strong>of</strong> protective relays. Between 1997 and<br />

2000, he worked for Itec Engineering, Florida <strong>Power</strong> and Light, and Siemens.<br />

Since 2000, Fernando has been an application engineer in international sales<br />

and marketing for Schweitzer Engineering Laboratories, Inc., providing<br />

training and technical assistance.<br />

Armando Guzmán received his BSEE with honors from Guadalajara<br />

Autonomous University (UAG), Mexico. He received a diploma in fiberoptics<br />

engineering from Monterrey Institute <strong>of</strong> Technology and Advanced<br />

Studies (ITESM), Mexico, and his MSEE from the University <strong>of</strong> Idaho, USA.<br />

He lectured at UAG and University <strong>of</strong> Idaho in power system protection and<br />

power system stability. Since 1993, Armando has been with Schweitzer<br />

Engineering Laboratories, Inc. in Pullman, Washington, where he is research<br />

engineering manager. He holds several patents in power system protection<br />

and metering. He is a senior member <strong>of</strong> IEEE.<br />

Gabriel Benmouyal, P.E. received his BASc in Electrical Engineering and<br />

his MASc in Control Engineering from Ecole Polytechnique, Université de<br />

Montréal, Canada, in 1968 and 1970. In 1969, he joined Hydro-Québec as an<br />

instrumentation and control specialist. He worked on different projects in the<br />

fields <strong>of</strong> substation control systems and dispatching centers. In 1978, he<br />

joined IREQ, where his main field <strong>of</strong> activity was the application <strong>of</strong><br />

microprocessors and digital techniques for substation and generating station<br />

control and protection systems. In 1997, he joined Schweitzer Engineering<br />

Laboratories, Inc. in the position <strong>of</strong> principal research engineer. Gabriel is an<br />

IEEE Senior Member, a registered pr<strong>of</strong>essional engineer in the Province <strong>of</strong><br />

Québec, and has served on the <strong>Power</strong> System Relaying Committee since May<br />

1989. He holds over six patents and is the author or coauthor <strong>of</strong> several papers<br />

in the fields <strong>of</strong> signal processing and power networks protection and control.<br />

© 2009 by Schweitzer Engineering Laboratories, Inc.<br />

All rights reserved.<br />

20091214 • TP6378-01<br />

82 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


Line Protection Bibliography<br />

Issue I <strong>of</strong> the <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong><br />

Costello, David, and Karl Zimmerman. “Determining the<br />

Faulted Phase.” March 2010.<br />

Stevens, Ian, Normann Fischer, and Bogdan Kasztenny.<br />

“Performance Issues With Directional Comparison<br />

Blocking Schemes.” March 2010.<br />

Zimmerman, Karl, and David Costello. “Fundamentals<br />

and Improvements for Directional Relays.” January 2010.<br />

Miller, Hank, John Burger, Normann Fischer, and<br />

Bogdan Kasztenny. “Modern Line Current Differential<br />

Protection Solutions.” January 2010.<br />

Calero, Fernando, Armando Guzmán, and Gabriel<br />

Benmouyal. “Adaptive Phase and Ground Quadrilateral<br />

Distance Elements.” December 2009.<br />

Calero, Fernando. “Distance Elements: Linking Theory<br />

With Testing.” August 2009.<br />

Plumptre, Frank, Mukesh Nagpal, Xing Chen, and<br />

Michael Thompson. “Protection <strong>of</strong> EHV Transmission<br />

Lines With Series Compensation: BC Hydro’s Lessons<br />

Learned.” January 2009.<br />

Calero, Fernando. “Rebirth <strong>of</strong> Negative-Sequence<br />

Quantities in Protective Relaying With Microprocessor-<br />

Based Relays.” November 2008.<br />

Calero, Fernando. “Mutual Impedance in Parallel<br />

Lines — Protective Relaying and Fault Location<br />

Considerations.” November 2008.<br />

Maezono, Paulo Koiti, Enrique Altman, Kennio Brito,<br />

Vanessa Alves dos Santos Mello Maria, and Fabiano<br />

Magrin. “Very High-Resistance Fault on a 525 kV<br />

Transmission Line — Case Study.” October 2008.<br />

Altuve, Héctor, Joseph Mooney, and George Alexander.<br />

“Advances in Series-Compensated Line Protection.”<br />

October 2008.<br />

Costello, David, and Karl Zimmerman. “Distance<br />

Element Improvements — A Case Study.” January 2008.<br />

Schweitzer III, Edmund O., and Stanley E. Zocholl.<br />

“Introduction to Symmetrical Components.”<br />

December 2007.<br />

Guzmán, Armando, Venkat Mynam, and Greg Zweigle.<br />

“Backup Transmission Line Protection for Ground Faults<br />

and <strong>Power</strong> Swing Detection Using Synchrophasors.”<br />

September 2007.<br />

Mooney, Joe. “Distance Element Performance Under<br />

Conditions <strong>of</strong> CT Saturation.” September 2007.<br />

Costello, David. “Lessons Learned Analyzing<br />

Transmission Faults.” September 2007.<br />

Hubertus, James, Joe Mooney, and George Alexander.<br />

“Application Considerations for Distance Relays on<br />

Impedance-Grounded Systems.” September 2007.<br />

Sánchez, Servando, Alfredo Dionicio, Martín Monjarás,<br />

Manuel Guel, Guillermo González, Octavio Vázquez,<br />

José L. Estrada, Héctor J. Altuve, Ignacio Muñoz,<br />

Iván Yánez, and Pedro Loza. “Directional Comparison<br />

Protection Over Radio Channels for Subtransmission<br />

Lines: Field Experience in Mexico.” September 2007.<br />

Hou, Daqing. “Relay Element Performance During<br />

<strong>Power</strong> System Frequency Excursions.” August 2007.<br />

Mooney, Joe, and Satish Samineni. “Distance Relay<br />

Response to Transformer Energization: Problems and<br />

Solutions.” January 2007.<br />

Stokes-Waller, Edmund. “Distance Protection: Pushing<br />

the Envelope.” October 2006.<br />

Plumptre, Frank, Stephan Brettschneider, Allen Hiebert,<br />

Michael Thompson, and Mangapathirao “Venkat”<br />

Mynam. “Validation <strong>of</strong> Out-<strong>of</strong>-Step Protection With a<br />

Real Time Digital Simulator.” September 2006.<br />

Araujo, Chris, Fred Horvath, and Jim Mack. “A<br />

Comparison <strong>of</strong> Line Relay System Testing Methods.”<br />

September 2006.<br />

Line Protection Bibliography | 83


Benmouyal, Gabriel, and Joe Mooney. “Advanced<br />

Sequence Elements for Line Current Differential<br />

Protection.” September 2006.<br />

Tziouvaras, Demetrios. “Relay Performance During<br />

Major System Disturbances.” September 2006.<br />

Roberts, Jeff, and Armando Guzmán. “Directional<br />

Element Design and Evaluation.” August 2006.<br />

Tziouvaras, Demetrios. “Protection <strong>of</strong> High-Voltage AC<br />

Cables.” January 2006.<br />

Abboud, Ricardo, Walmer Ferreira Soares, and Fernando<br />

Goldman. “Challenges and Solutions in the Protection <strong>of</strong><br />

a Long Line in the Furnas System.” November 2005.<br />

Zimmerman, Karl, and Dan Roth. “Evaluation <strong>of</strong><br />

Distance and Directional Relay Elements on Lines With<br />

<strong>Power</strong> Transformers or Open-Delta VTs.” September<br />

2005.<br />

Benmouyal, Gabriel. “The Trajectories <strong>of</strong> Line Current<br />

Differential Faults in the Alpha Plane.” September 2005.<br />

Zhou, Zexin, Xia<strong>of</strong>an Shen, Daqing Hou, and Shaojun<br />

Chen. “Analog Simulator Tests Qualify Distance Relay<br />

Designs to Today’s Stringent Protection Requirements.”<br />

September 2005.<br />

Mooney, Joe, and Normann Fischer. “Application<br />

Guidelines for <strong>Power</strong> Swing Detection on Transmission<br />

Systems.” September 2005.<br />

Zimmerman, Karl, and David Costello. “Impedance-<br />

Based Fault Location Experience.” August 2004.<br />

Benmouyal, Gabriel, Daqing Hou, and Demetrios<br />

Tziouvaras. “Zero-Setting <strong>Power</strong>-Swing Blocking<br />

Protection.” March 2005.<br />

Tziouvaras, Demetrios, Jeff Roberts, and Gabriel<br />

Benmouyal. “New Multi-Ended Fault Location Design<br />

for Two- or Three-Terminal Lines.” November 2004.<br />

Topham, Graeme, and Edmund Stokes-Waller. “Steady-<br />

State Protection Study for the Application <strong>of</strong> Series<br />

Capacitors in the Empangeni 400 kV Network.”<br />

October 2004.<br />

Calero, Fernando, and Daqing Hou. “Practical<br />

Considerations for Single-Pole-Trip Line-Protection<br />

Schemes.” September 2004.<br />

Benmouyal, Gabriel, and Tony Lee. “Securing Sequence-<br />

Current Differential Elements.” September 2004.<br />

Fodero, Ken, and Girolamo Rosselli. “Applying Digital<br />

Current Differential Systems Over Leased Digital<br />

Service.” September 2004.<br />

Moxley, Roy. “Analyze Relay Fault Data to Improve<br />

Service Reliability.” March 2004.<br />

Schweitzer III, Edmund O., Ken Behrendt and Tony Lee.<br />

“Digital Communications for <strong>Power</strong> System Protection:<br />

Security, Availability, and Speed.” January 2004.<br />

Tziouvaras, Demetrios, and Daqing Hou. “Out-<strong>of</strong>-Step<br />

Protection Fundamentals and Advancements.”<br />

December 2003.<br />

Henville, Charlie, Ralph Folkers, Allen Hiebert, and Rudi<br />

Wierckx. “Dynamic Simulations Challenge Protection<br />

Performance.” October 2003.<br />

Benmouyal, Gabriel, Michael Bryson, and Marc Palmer.<br />

“Implementing a Line Thermal Protection Element Using<br />

Programmable Logic.” September 2003.<br />

Carroll, Debra, John Dorfner, Tony Lee, Ken Fodero,<br />

and Chris Huntley. “Resolving Digital Line Current<br />

Differential Relay Security and Dependability Problems:<br />

A Case History.” September 2002.<br />

Tziouvaras, Demetrios, Héctor Altuve, Gabriel<br />

Benmouyal, and Jeff Roberts. “Line Differential<br />

Protection With an Enhanced Characteristic.”<br />

July 2002.<br />

Guzmán, Armando, Joe Mooney, Gabriel Benmouyal,<br />

and Normann Fischer. “Transmission Line Protection<br />

System for Increasing <strong>Power</strong> System Requirements.”<br />

April 2001.<br />

Alexander, George, Joe Mooney, and William Tyska.<br />

“Advanced Application Guidelines for Ground Fault<br />

Protection.” October 2001.<br />

Roberts, Jeff, Demetrios Tziouvaras, Gabriel Benmouyal,<br />

and Hector J. Altuve. “The Effect <strong>of</strong> Multiprinciple Line<br />

Protection on Dependability and Security.” February<br />

2001.<br />

Costello, David. “Understanding and Analyzing Event<br />

Report Information.” October 2000.<br />

84 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


Hou, Daqing, and Jeff Roberts. “Capacitive Voltage<br />

Transformers: Transient Overreach Concerns and<br />

Solutions for Distance Relaying.” July 2000.<br />

Benmouyal, Gabriel, and Jeff Roberts. “Superimposed<br />

Quantities: Their True Nature and Application in Relays.”<br />

October 1999.<br />

Zocholl, Stanley E. “Symmetrical Components: Line<br />

Transposition.” March 1999.<br />

Behrendt, Kenneth. “Relay-to-Relay Digital Logic<br />

Communication for Line Protection, Monitoring, and<br />

Control.” November 1998.<br />

Fleming, Bill. “Negative-Sequence Impedance<br />

Directional Element.” February 1998.<br />

Schweitzer III, Edmund O., Bill Fleming, Tony Lee, and<br />

Paul Anderson. “Reliability Analysis <strong>of</strong> Transmission<br />

Protection Using Fault Tree Methods.” October 1997.<br />

Hou, Daqing, Armando Guzmán, and Jeff Roberts.<br />

“Innovative Solutions Improve Transmission Line<br />

Protection.” October 1997.<br />

Mooney, Joe, and Jackie Peer. “Application Guidelines<br />

for Ground Fault Protection.” October 1997.<br />

Schweitzer III, Edmund O., and John Kumm. “Statistical<br />

Comparison and Evaluation <strong>of</strong> Pilot Protection<br />

Schemes.” October 1996.<br />

Guzmán, Armando, Jeff Roberts, and Daqing Hou.<br />

“New Ground Directional Elements Operate Reliably for<br />

Changing System Conditions.” September 1996.<br />

Mooney, Joe. “Microprocessor-Based Transmission Line<br />

Relay Applications.” March 1996.<br />

Roberts, Jeff, Edmund O. Schweitzer III, Renu Arora,<br />

and Ernie Poggi. “Limits to the Sensitivity <strong>of</strong> Ground<br />

Directional and Distance Protection.” October 1995.<br />

Zocholl, Stanley E. “Three-Phase Circuit Analysis and<br />

the Mysterious k 0<br />

Factor.” September 1995.<br />

Roberts, Jeff, Armando Guzmán, and Edmund O.<br />

Schweitzer III. “Z = V/I Does Not Make a Distance<br />

Relay.” April 1994.<br />

Schweitzer III, Edmund O. “A Review <strong>of</strong> Impedance-<br />

Based Fault Locating Experience.” June 1993.<br />

Schweitzer III, Edmund O., and Jeff Roberts. “Distance<br />

Relay Element Design.” April 1993.<br />

Zimmerman, Karl, and Joe Mooney. “Comparing Ground<br />

Directional Element Performance Using Field Data.”<br />

April 1993.<br />

Schweitzer III, Edmund O., and Daqing Hou. “Filtering<br />

for Protective Relays.” April 1993.<br />

Roberts, Jeff, and Edmund O. Schweitzer III. “Analysis<br />

<strong>of</strong> Event Reports.” April 1991.<br />

Schweitzer III, Edmund O. “New Developments in<br />

Distance Relay Polarization and Fault Type Selection.”<br />

October 1989.<br />

Line Protection Bibliography | 85


<strong>SEL</strong> University 2010 Course Schedule<br />

Course No. Course Name July August September October November December<br />

Fundamentals<br />

PWRS 400<br />

<strong>Power</strong> System Fundamentals for Engineers<br />

Pullman, WA<br />

3—6<br />

PROT 401<br />

Protecting <strong>Power</strong> Systems for Engineers<br />

Chicago, IL<br />

19—23<br />

Pullman, WA<br />

9—13<br />

Los Angeles, CA<br />

6—10<br />

PROT 403<br />

Distribution System Protection<br />

Dothan, AL<br />

19—21<br />

PROT 405<br />

Industrial <strong>Power</strong> System Protection<br />

Pullman, WA<br />

21—24<br />

PROT 407<br />

Transmission Line Protection<br />

Pullman, WA<br />

23—25<br />

PROT 409<br />

Generation System Protection<br />

Pullman, WA<br />

13—15<br />

PROT 411<br />

Substation Equipment Protection<br />

Folsom, CA<br />

12—14<br />

Applications<br />

APP 87<br />

<strong>SEL</strong>-387 and <strong>SEL</strong>-587 Percentage-Restrained Differential Relays<br />

Folsom, CA<br />

15—16<br />

APP 300G<br />

<strong>SEL</strong>-300G Generator Relay<br />

Pullman, WA<br />

16—17<br />

APP 311L<br />

<strong>SEL</strong>-311L Line Current Differential Relay<br />

Pullman, WA<br />

26—27<br />

Folsom, CA<br />

16—17<br />

APP 351<br />

<strong>SEL</strong>-351 Directional Overcurrent and Reclosing Relay<br />

Chicago, IL<br />

3—4<br />

Folsom, CA<br />

28—29<br />

Orlando, FL<br />

14—15<br />

APP 351R<br />

<strong>SEL</strong>-351R Recloser Control<br />

Charlotte, NC<br />

9—10<br />

APP 403<br />

Enhancing Distribution Protection Using the <strong>SEL</strong>-351S<br />

Dothan, AL<br />

22—23<br />

APP 405<br />

Protecting Induction Motors Using the <strong>SEL</strong>-710, <strong>SEL</strong>-701<br />

Charlotte, NC<br />

13—15<br />

Pullman, WA<br />

28—30<br />

Dallas, TX<br />

7—9<br />

APP 421<br />

<strong>SEL</strong>-421 Protection, Automation, and Control System<br />

Orlando, FL<br />

16—18<br />

APP 487B<br />

<strong>SEL</strong>-487B Bus Protection Relay<br />

Raleigh, NC<br />

5—6<br />

APP 651R<br />

<strong>SEL</strong>-651R Advanced Recloser Control<br />

Atlanta, GA<br />

26—27<br />

APP 751A Feeder Protection Using the <strong>SEL</strong>-751A Relay NEW in 2010<br />

APP 3530 <strong>SEL</strong>-3530 Real-Time Automation Control NEW in 2010<br />

Pullman, WA<br />

13—14<br />

Philadelphia, PA<br />

14—15<br />

Testing<br />

TST 101<br />

<strong>SEL</strong> Relay Testing Basics<br />

Chicago, IL<br />

5—6<br />

Nashville, TN<br />

16—17<br />

TST 107<br />

<strong>SEL</strong> Transmission Substation Relay Testing<br />

Denver, CO<br />

14—16<br />

Web-Based Training<br />

WBT 351<br />

<strong>SEL</strong>-351 Relay Coordination Techniques<br />

Online<br />

2 & 4<br />

WBT 421<br />

Applying Dual-Breaker Reclosing With the <strong>SEL</strong>-421 Relay<br />

Online<br />

6 & 8<br />

WBT 2411S RTU Replacement Using IEC 61850<br />

Online<br />

31<br />

Online<br />

2<br />

WBT 3332<br />

Configuring Logical Expressions Using the <strong>SEL</strong>-3332 Intelligent<br />

Server and <strong>SEL</strong>-3351 System Computing Platform<br />

Online<br />

14 & 16<br />

Since its inception, <strong>SEL</strong> University has had one clear mission—to provide the education and training needed to make electric power safer, more reliable,<br />

and more economical. We’ve developed programs that help you meet the technical challenges and complexities <strong>of</strong> integrating digitally based technologies<br />

into your expanding power system infrastructure.<br />

For the most up-to-date course descriptions, dates, and locations, visit www.selinc.com/selu • Phone: +1.509.338.4026 • selu@selinc.com<br />

86 | <strong>Journal</strong> <strong>of</strong> <strong>Reliable</strong> <strong>Power</strong>


©2010 Schweitzer Engineering Laboratories, Inc.<br />

20100616

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!