You are on page 1of 23

Course Outline

1. PRELIMINARY DATA
1.1 SCOPE
1.2 LOAD DATA
1.3 LOAD ANALYSIS
1.4 TERMINOLOGY
2. ESTIMATION OF LOADS
2.1 PREPARATION OF LOAD DATA
2.2 INDIVIDUAL LOADS
2.3 EMERGENCY LOADS
2.4 AREA LOADS
2.5 ACTIVITY LOADS
3. SELECTION OF ELECTRIC POWER SOURCE
3.1 ELECTRIC POWER SOURCES
3.2 ACCEPTABLE ELECTRIC POWER SOURCES
3.3 PURCHASED ELECTRIC POWER REQUIREMENTS
This course includes a multiple-choice quiz at the end, which is designed to enhance the
understanding of the course materials.

Learning Objective
At the conclusion of this course, the student will:

Learn how to convert motor horsepower to input kilovolt-amperes for preliminary


estimates.
Learn the definition of Demand Factor and how it is used to determine power
requirements.
Learn the definition of Coincidence Factor and how it is used to determine power
requirements.
Learn the definition of Diversity Factor and how it is used to determine power
requirements.
Learn the definition of Load Factor and how it is used to determine power
requirements.
Learn recommended Demand and Load Factors for a wide variety of buildings
and activities.
Learn how to quickly estimate lighting power requirements from illumination
levels for various areas.
Learn guidelines for selection of Demand Factors.
Learn guidelines for selection of Load Factors.
Learn about the three types of emergency power loads.

All electrical systems are susceptible to short circuits and the abnormal current levels
they create. These currents can produce considerable thermal and mechanical
stresses in electrical distribution equipment. Therefore, it's important to protect
personnel and equipment by calculating short-circuit currents during system
upgrade and design. Because these calculations are life-safety related, they're
mandated by 110.9 of the NEC, which states:

Equipment intended to interrupt current at fault levels shall have an interrupting


rating sufficient for the nominal circuit voltage and the current that is available at the
line terminals of the equipment. Equipment intended to interrupt current at other
than fault levels shall have an interrupting rating at nominal circuit voltage sufficient
for the current that must be interrupted.

When you apply these requirements to a circuit breaker, you must calculate the
maximum 3-phase fault current the breaker will be required to interrupt. This
current can be defined as the short-circuit current available at the terminals of the
protective device.

You can assume that 3-phase short circuits are bolted, or have no impedance. In
addition, a 3-phase short circuit can be considered a balanced load, which means you
can use a single-phase circuit to analyze one of the phases and the neutral.

Distribution equipment, such as circuit breakers, fuses, switchgear, and MCCs, have
interrupting or withstand ratings defined as the maximum rms values of symmetrical
current. A circuit breaker can't interrupt a circuit at the instant of inception of a
short. Instead, due to the relay time delay and breaker contact parting time, it will
interrupt the current after a period of five to eight cycles, by which time the DC
component will have decayed to nearly zero and the fault will be virtually
symmetrical.

Closing a breaker against an existing fault makes it possible to intercept the peak of
the asymmetrical short-circuit current, which is greater than the rms value of the
symmetrical current. For this reason, equipment is also tested at a particular test
X/R ratio value typical to a particular electrical apparatus, such as switchgear,
switchboards, or circuit breakers, and is designed and rated to withstand and/or
close and latch the peak asymmetrical current described above.

Fault analysis is required to calculate and compare symmetrical and asymmetrical


current values in order to select a protective device to adequately protect a piece of
electrical distribution equipment.

Methods of calculation

Rather than using a theoretical approach to determine short-circuit currents,


published standards offer methods to compute a symmetrical steady state solution to
which you can apply a multiplier in order to obtain the peak value of an asymmetrical
current. The result is precise enough to fall within an acceptable tolerance to meet
NEC requirements.

The classical approach and the method defined by ANSI/IEEE are two such
industry-accepted methods for calculating short circuits. Both methods assume that
the fault impedance is zero (bolted short circuit) and the pre-fault voltage is constant
during the evolution of the fault. In actuality, the fault has its own impedance, and
the voltage drop, due to the short-circuit current, lowers the driving voltage.

The classical approach is used to calculate the Thevenin equivalent impedance as


seen by the system at the point of the fault. Thevenin impedance is defined as the
impedance seen at any point in a circuit once all the voltage generators have been
short circuited and all the current generators have been opened. Transformer and
utility impedances and rotating machine subtransient reactances describe all
possible contributions to a short circuit. Once we have calculated the symmetrical
and peak duties, we can determine the required rating of the protective device by
direct comparison to manufacturer equipment ratings.

The ANSI/IEEE short-circuit calculation method follows a step-by-step process.

The IEEE standard permits the exclusion of all 3-phase induction motors below 50

The ANSI/IEEE method, which is described in IEEE Std. C37.010-1979 and its revision in
1999, is used for high-voltage (above 100V) equipment. It calls for determining the
momentary network fault impedance, which makes it possible to calculate the close and latch
rating of the breaker. It also calls for identifying the interrupting network fault impedance,
which makes it possible to calculate the interrupting duty of the breaker. The interrupting
network fault impedance value differs from the momentary network fault impedance value in
that the impedance increases from the subtransient to transient level.

hp and all single-phase motors. Hence, no reactance adjustment is needed for these
motors. The Chart at left clarifies the ANSI/IEEE procedure.

Classical calculation
Begin by converting all impedances to per unit values. Per unit base values and formulae
used are as follows:
Sbase =100MVA
Vbase =26.4 kV

Let's run through an example calculation to make this discussion a little more tangible.
Refer to the one-line diagram in the Figure below with the following input data:

Utility: 26.4kV, 1,200MVA, X/R=41

Transformer (T1): 2MVA, 26.4/4.16kV, DY-G, Z=7%, X/R515

Motor 1 (M1): Induction, 4.16kV, 1,000 hp, PF=0.8, efficiency50.8, X"d= 0.16 pu, X/R=28

Motor 2 (M2): Induction, 4.16kV, 49 hp, PF=0.8, efficiency=0.8, X"d=0.17 pu, X/R=10

This over-simplified one-line diagram of a power distribution system included values necessary for working
through the two methods of short-circuit calculation referred to in the text.

Now it's possible to calculate the equivalent Thevenin impedance for a fault at Bus 2 by
combining the per unit X and R values to obtain the relative impedances.
ZFault=(Zutility+ZT1)||ZMotor1||ZMotor2=(0.0021+j0.083+0.005+j0.07)||(0.49+j13.8)||(29.8+j29
8)=0.166+j2.817 pu=2.823ej86.6
We may now calculate the short-circuit current rms at Bus 2:

The peak duty the breaker is required to close and latch may be evaluated using the
following formula, which constitutes a multiplier to the rms current, which was calculated
above:

Use Table 1, page 1 in ANSI C37.06-1997 Preferred Ratings and Related Required
Capabilities to rate new switchgear. It's useful in comparing calculated duty (4,916A and

12,692A) and standard ratings. The Table includes sample values extracted from the ANSI
table.

Compare calculated duty and standard ratings using Table 1 in ANSI C37.06-1997.

These are the short-circuit current ratings required for our switchgear duty corresponding
to a continuous current, for example, 1,200A. No further steps have to be taken, as the table
itself, by comparison, provides the required specifications for the equipment to be installed.

ANSI/IEEE calculation
The ANSI/IEEE calculation method is based on the same per unit quantities as calculated
before. However, it differs from the classical method because it makes it possible to study
two separate circuits derived from the original one: one resistive only and one reactive only.
This will be carried out for both momentary and interrupting network fault impedances.
For each network, Thevenin equivalent resistance and Thevenin equivalent reactance will
then be combined in order to obtain the equivalent Thevenin impedance. This is the
significant difference between the ANSI/IEEE procedure and the classical calculation
method.
As mentioned before, the momentary network fault impedance is based on the subtransient
reactances of the rotating machines, which allows for the calculation of the first-cycle peak
fault duty. The total fault resistance and reactance values will be calculated separately,
following the same formula as the ZFault equation in the classical calculation section, except
the Zs must be replaced with the Rs and Xs.
Then they'll be combined as total fault impedance ZFault, which will yield ISC3-phase and
IPeak according to the formulas.
The interrupting network fault impedance is based on individual equipment transient
reactances. In the previous example, only the reactance of Motor 1 needs to be adjusted. It's
acceptable to neglect Motor 2 at medium voltage levels. The resistances of the network, in
fact, don't vary with respect to time. ANSI C37.010-1999 identifies the adjustment factor as
1.5.

In this case, the total fault resistance and fault reactance (with adjustments) will be
calculated separately as already seen.
ISC3-phase, symmetrical duty is calculated as it was in the classical method. However, it's
typically characterized by a smaller magnitude because the Zfault interrupting current is
larger than the one in the momentary network calculation.
ISC3-phase is essential because a multiplier factor is applied to this quantity for comparison to
the breaker interrupting rating.
This multiplier will account for:

The additive contribution of the DC current component, which might still be alive after the time of
contact parting.

The eventual subtractive contribution of the AC current decay, due to the evolution of the reactances
toward larger values. This effect is possible when the generation of power is local.

The multipliers, in function of time of contact parting and of the ratio X/R at the point of
fault, are described in curves starting from figure A-8, page 60, C37.010-1999
(Figureabove).
Once ISC3-phase has been multiplied by this factor (between 1 and 1.25), you have the
minimum rating of your equipment. As in the classical method, you can also use Table 1,
page 1 in ANSI C37.06-1997 to determine a standard rating.

Which method is better?

Both methods basically provide the same results. There are no theoretical reasons to prefer
one to the other, only practical reasons. The ANSI/IEEE approach is the evolution of a
method conceived in the '70s in the United States, when no computer-assisted calculations
were available. ANSI/IEEE C37.010-1999 can only be used at medium or high voltages and
only at 60 Hz. Calculation programs have been developed to determine fault currents that
apply the multiplier factors called for in this standard. In fact, some clients may ask for the
application of this calculation methodology by contract. Manufacturers may also recall the
ANSI/IEEE standard in their catalogues. The classical method is used mainly in low-voltage
studies and can also be applied at 50 Hz. It's a well-known procedure because it's a common
topic in every power system college course.

Performing short-circuit calculations requires an understanding of various system


components and their interaction.
It's very important to understand the meaning of the term "short-circuit fault." Basically, a
short-circuit fault in a power system is an abnormal condition that involves one or more
phases unintentionally coming in contact with ground or each other. Thus, short-circuit
protection is necessary to protect personnel and apparatus from the destructive effects of
the resulting excessive current flow, which is caused by the relatively low impedance of the
short-circuit fault connection.
To provide the required protection, we must determine the extent of short-circuit current at
various points of our power distribution system. This determination requires a calculation.
In Part one of this three-part series, we'll discuss specific parameters needed to perform
short-circuit calculations of industrial and commercial power systems according to the
ANSI/IEEE 141 standard, Recommended Practice for Electric Power Distribution for
Industrial Plants (Red Book).
Sources of fault current
Where does fault current come from? Basically, it comes from rotating electric machinery,
usually in the form of synchronous generators, synchronous motors and condensers,
induction machines, and electric utility systems. The magnitude of fault current from these
sources is limited by the impedance of the machine itself as well as the impedance between
the machine and the fault itself.
Because a synchronous generator has a prime mover and an externally excited field, its fault
current will continue unless interrupted by some switching means.
Synchronous motors and condensers supply current to a fault in much the same way as
synchronous generators; however, their fault current diminishes as their magnetic fields
decay. Induction motor fault current is generated by inertia that is driving the motor in the
presence of a field flux, which is produced by induction from the motor's stator.
The basics
A balanced 3-phase fault implies that all three phases of the power system are
simultaneously short-circuited to each other through a direct or "bolted" connection.
Although the probability of this happening is small, relative to the probability of other types

of unbalanced faults occurring (e.g., phase-to-ground and phase-to-phase faults), we


nevertheless use a balanced 3-phase fault for a short-circuit study for the following reasons.
* Often, a 3-phase fault produces the largest short-circuit current magnitude; thus, this
worst-case result is then used as the basis to select the short-circuit capabilities of
switchgear from manufacturers' tables.
* Short-circuit calculations are simplest for a balanced 3-phase fault because symmetry of
the fault connection permits us to consider only one of the three phases.
The other types of unbalanced short-circuit faults are important in selecting the timecurrent characteristics and settings of phase-overcurrent and ground-fault protective
devices to provide selective coordination. This coordination assures service continuity and
minimizes damage to switchgear and load equipment. However, unbalanced fault
calculations are more difficult to perform for industrial and commercial power systems and
require a knowledge of the method of symmetrical components.
Symmetrical RMS current versus short-circuit duty
In Fig. 1, we see a 3-phase synchronous generator, previously unloaded, that has been
subjected to a balanced, 3-phase fault across its accessible terminals. In general, an
asymmetrical short-circuit current waveform is produced by a balanced, 3-phase fault.
Fig. 2 shows a typical asymmetrical short-circuit waveform (bottom) for one phase of the 3phase synchronous generator. The diagram also shows that this waveform is a combination
of two components: a unidirectional (DC) component (upper left waveform) and a
symmetrical (AC) component (upper right waveform). The DC component approximately
decays exponentially with a time constant equal to [X.sub.EQ]/[R.sub.EQ], where
[X.sub.EQ] and [R.sub.EQ] are the equivalent inductive reactance and resistance at the fault
location, respectively. The DC component eventually decays to zero, and the amplitude of
the symmetrical AC component eventually decays to a constant amplitude in the steadystate. The sum of these two components at any time instant is equal to the total
asymmetrical short-circuit current at that instant.
The root mean-square (rms) value of the asymmetrical short-circuit current waveform is the
basis for the selection of the short-circuit capabilities of circuit breakers and fuses.
Calculation of the precise rms value of an asymmetrical current at any time after the
inception of a short-circuit may be very involved. Accurate decrement factors to account for

the DC component at any time are required, as well as factors for the rate of change of the
apparent reactance of the generators. This precise method may be used, if desired; however,
simplified methods have evolved whereby the DC component is accounted for by simple
multiplying factors. These multiplying factors convert the rms value of the symmetrical AC
component (symmetrical rms current) into rms current of the asymmetrical waveform,
including the DC component (asymmetrical rms current or short-circuit current duty).
Types of networks to calculate symmetrical rms current
In order to utilize AC circuit theory in calculating symmetrical rms current, three types of
networks are used to represent the power system over three time intervals of the fault-on
time period.
* First-cycle (momentary) network.
* Contact-parting (interrupting) network.
* Approximately 30 cycle network.
These networks only differ from one another by the assignments of constant reactances for
the machines.
First-cycle (momentary) network. This network is used to calculate the first-cycle
(momentary) symmetrical rms current. Here, the rotating machine sources of short-circuit
current are represented, for the most part, by their subtransient reactances, according to the
entries in the first column of Tables 4-1 and 4-2 of the 1993 edition of the IEEE Red Book
(or Tables 24 and 25 of the 1986 edition).
Contact-parting (interrupting) network. This network is used to calculate the contactparting (interrupting) symmetrical rms current for circuit breaker minimum contactparting times of 1.5 to 4 cycles after the inception of the short-circuit fault. Here, the
rotating machine sources of short-circuit current are represented by different constant
reactances than the first-cycle (momentary) network, according to the entries in the second
column of Tables 4-1 and 4-2 of the 1993 edition of the IEEE Red Book (or Tables 24 and 25
of the 1986 edition).
Approximately 30 cycle network. This network is often a minimum-source representation to
investigate whether minimum short-circuit currents are sufficient to operate currentactuated relays. Minimum-source networks might apply at night or when production lines

are down for any reason. Some of the source circuit breakers may be open and all motor
circuits may be off. In-plant generators are represented with transient reactance or a larger
reactance that is related to the magnitude of decaying generator short-circuit current at the
desired calculation time.
In Part 2 (December 1995 issue), we'll use a simple example that shows the necessary steps
needed to determine the appropriate reactances and resistances of the power system
apparatus for the first-cycle (momentary) and contact-parting (interrupting) networks.
In Part 3 (April 1996 issue), we'll explain how to construct the networks, impose the fault
connection, and use basic AC circuit theory to reduce the networks to equivalent reactance
and resistance values. These equivalent reactance and resistance values can then be used to
calculate symmetrical rms short-circuit currents and equivalent short-circuit
[X.sub.EQ]/[R.sub.EQ] ratios.
RELATED ARTICLE: TERMS TO KNOW
Asymmetrical short-circuit current: A fault current whose waveform is asymmetrical to the
zero axis. The peak positive current at any of the waveform loops will be greater than 1.414
times the rms symmetrical current.
Symmetrical short-circuit current: A fault-current whose waveform is symmetrical about
the zero axis. In other words, the positive peak current has the same value as the negative
peak current. These peak (maxium) currents are always equal to 1.414 times the rms
symmetrical current.
rms current: Current which, while flowing through a given ohms resistance, will produce
heat at the same rate as a DC ampere.
Reactance: An opposition to the flow of electrical current in a circuit. It is usually expressed
in ohms and consists of inductive and capacitive parts.
Sub-transient reactance: The value of reactance that determines the amount of short-circuit
current during the first half cycle after the fault occurs.
Performing short-circuit calculations requires an understanding of various system
components and their interaction.

In Part 1 of this article, which was featured in the June 1995 issue, we discussed the types of
networks to calculate short-circuit current (i.e., symmetrical rms current). In Parts 2 and 3
(April 1996 issue), we'll describe the per-unit method of performing short-circuit
calculations in accordance with ANSI/IEEE 141-1993, IEEE Recommended Practice for
Electric Power Distribution for Industrial Plants (the Red Book). We'll use a simple example
of an industrial power system to show the data preparation steps necessary in determining
the appropriate per-unit reactances and resistances of power system equipment for firstcycle (momentary) and contact-parting (interrupting) networks.
Overview of per-unit analysis
Per-unit analysis is based on "normalized" representations of the electrical quantities (i.e.,
voltage, current, impedance, etc.). The per-unit equivalent of any electrical quantity is
dimensionless and is defined as the ratio of the actual quantity in units (i.e., volts, amperes,
ohms, etc.) to an appropriate base value of the electrical quantity. This is expressed by the
following equation.
per-unit value = actual value [divided by] base value (eq. 1)
The actual value can be a phasor or complex number (i.e., magnitude and phase) in units,
whereas the base value is simply a real number in units.
The key factor of a per-unit normalization procedure is the selection of the base values of
the electrical quantities. In practice, the base values of 3-phase apparent power in MVA (i.e.
base MVA) and line-to-line voltage in kV (i.e. base kV) are assigned on each side of every 3phase power transformer in accordance with the following simple rules.
First, a convenient base MVA is chosen (e.g., base MVA = 1, 10, or 100) and is common
throughout the entire power system. Second, base kV on any side of a 3-phase power
transformer is designated to be equal to the nominal, line-to-line nameplate kV rating of the
transformer.
The base values of per-phase base impedance in ohms (i.e., base Z) and line current in
kiloamperes (i.e., base kA) are then derived from base MVA and base kV, according to the
following equations. base Z = [(base kV).sup.2] [divided by] base MVA (eq. 2) base kA =
base MVA [divided by] ([square root of 3] x base kV) (eq. 3)
You gain an important advantage by assigning the base values per the above selection rules
and using them to normalize (i.e., "per-unitize") the electrical quantities. This advantage can

be seen in the per-phase equivalent circuit model of any 3-phase transformer connection: It
is simply a per-unit series impedance that accounts for the conductor losses and leakage
fluxes.
The following equation is often used in [TABULAR DATA FOR TABLE 1 OMITTED] the
data preparation stage to adjust the per-unit impedance of power system apparatus
whenever the 3-phase nameplate ratings are different than the power system's 3-phase base
quantities.
adjusted [Z.sub.pu] = unadjusted [Z.sub.pu] x (base MVA [divided by] MVA rating) x [(kV
rating [divided by] base kV).sup.2](eq. 4)
Data preparation for a simplified example In the absence of nameplate information or data
from equipment manufacturers, typical data can be referenced from tables and figures
included in the IEEE Red Book. The data preparation steps to determine the appropriate
per-unit reactances and resistances for first-cycle (momentary) and con-tacting-parting
(interrupting) networks for the simplified industrial power system shown in Fig. 1 are as
follows.
Base values of voltage. The base values of line-to-line voltage are simply the nominal
nameplate, line-to-line voltage ratings of the 3-phase transformers. Base MVA is chosen to
be 10 MVA and is constant throughout the system.
Utility. The per-phase equivalent circuit model of the utility tie at the plant is a voltage
source in series with an impedance. This source is the nominal per-unit line-to-neutral
voltage at the service entrance point. The per-unit impedance is the same for both first-cycle
and interrupting networks; thus, no superscript "f" or "I" is necessary for its symbol. (In the
equations that follow, superscripts "f" and "I" refer to the per-unit reactance and resistance,
respectively, for the first-cycle and interrupting networks.)
The magnitude of the impedance ([Z.sub.u] = 0.01 per unit) is calculated by using the
following equation.
[Z.sub.u] = base MVA [divided by] (SCA MVA)(eq. 5)
Here, SCA MVA is equal to 1000 and represents the available short-circuit apparent power
delivered by the utility from all sources outside the plant.

To resolve [Z.sub.u] into reactive ([X.sub.u]) and resistive ([R.sub.u]) components, the
following equations are used.
X = Z sin ([tan.sup.-1] X/R)(eq. 6a)
R = Z cos ([tan.sup.-1] X/R)(eq. 6b)
In this event, a conservative assumption is to let [X.sub.u] = [Z.sub.u] = 0.01 per unit and
[R.sub.u] = 0.
In our example, as shown in Fig. 1, the short-circuit X/R ratio at the utility tie is unavailable.
Thus, we assume [X.sub.u] = 0.01 per unit and [R.sub.u] = 0.
Transformers. The per-phase equivalent circuit model of any 3-phase transformer
[TABULAR DATA FOR TABLE 2 OMITTED] connection is simply a per-unit series
impedance. This impedance is the same for both first-cycle and interrupting networks; thus,
no superscript "f" or "I" is necessary for its symbol.
The unadjusted impedance of the transformer is provided on its nameplate and is expressed
as a percentage of rated impedance. In other words, you simply divide it by 100% to arrive
at its unadjusted per-unit value. This per-unit impedance is adjusted with respect to the
system base quantities per Equation 4, and the adjusted per-unit impedance ([Z.sub.t]) is
resolved into reactive ([X.sub.t]) and resistive ([R.sub.t]) components per Equations 6a and
6b. (The typical short-circuit X/R ratios of the transformers in Fig. 1 are taken from Fig. 4A1 of the 1993 IEEE Red Book.) The results for the transformers in Fig. 1 are shown in Table
1.
Cable. The per-phase equivalent circuit model of the cable in Fig. 1 is simply a per-unit
series impedance. This impedance of the cable is the same for both first-cycle and
interrupting networks; thus, no superscript "f" or "I" is necessary for its symbol.
To find the cable's per-unit reactance ([X.sub.c]) and resistance ([R.sub.c]), the following
equations are used.
[X.sub.c] = [(X ohms per 1000ft) x (length of run in ft)] [divided by] [number of parallel
conductors per phase x base Z in ohms] (eq. 7a)
[R.sub.c] = [(R ohms per 1000ft) x (length of run in ft)] [divided by] [no. of parallel
conductors per phase x base Z in ohms] (eq. 7b)

Using the above equations, [X.sub.c] is 0.0030 per unit and [R.sub.c] = 0.0043 per unit.
(The approximate reactance and resistance data (in ohms per 1000 ft) noted alongside the
cable in Fig. 1 are taken from Table 4A-7 in the 1993 IEEE Red Book.)
Turbine-generator. In general, the per-phase equivalent circuit model of a rotating machine
is a voltage source in series with an impedance that varies with time during the fault. Based
on Table 4-1, Chapter 4 of the 1993 IEEE Red Book, the unadjusted per-unit reactance of
the turbine-generator for both the first-cycle and interrupting networks is 1.0 [X.sub.d]",
where [X.sub.d]" is the saturated direct-axis subtransient reactance of the generator in perunit. You adjust this reactance with respect to the system base quantities by using Equation
9, with the corresponding adjusted resistance calculated by using Equation 8.
adjusted R = adjusted X [divided by] short-circuit X/R (eq. 8) [X.sup.f]=(1.0 [X.sub.d]") x
(base MVA [divided by] MVA rating) x [(kV rating [divided by] base kV).sup.2] (eq. 9)
The calculation results are shown in Table 2. (The typical machine reactance and shortcircuit X/R data are from Table 4A-1 and Figs. 4A-2 and 4A-3 of the 1993 IEEE Red Book.
Incidentally, there is a typographical error in Table 4A-1 of the first printing of this book:
the left-most column of Table 4A-1 should be labeled [X.sub.d]" and the right-most column
[X.sub.d]'.)
Large motors. Based on Table 4-1, Chapter 4 of the 1993 IEEE Red Book, the unadjusted
per-unit reactances for the first-cycle and interrupting networks are 1.0 [X.sub.d]" and 1.5
[X.sub.d]" respectively.
Equation 9 is used to adjust the first-cycle reactance, where the 3-phase kVA rating is
approximately equal to the horsepower (hp) rating for induction motors and 0.8 power
factor (PF) synchronous motors. The corresponding adjusted first-cycle resistance is
calculated by using Equation 8. The following equations are then used to calculate the
adjusted interrupting reactance and resistance from the corresponding first-cycle values.
[X.sub.[M.sup.I]] = 1.5 x [X.sub.[M.sup.f]] (eq. 10a)
[R.sub.[M.sup.I]] = 1.5 x [R.sub.[M.sup.f]] (eq. 10b)
The results of these calculations also are shown in Table 2.

Small horsepower induction motors. The unadjusted per-unit reactances of the small
horsepower (less than 250 hp) induction motors shown in Fig. 1 for the first-cycle
[TABULAR DATA FOR TABLE 3 OMITTED] and interrupting networks are taken from the
footnotes of Table 4-2 in the 1993 IEEE Red Book. Specifically, you should refer to the
footnotes of the row entitled "All others, 50 HP and above" for the 150 hp induction motor.
Also refer to the footnotes of the row entitled "All smaller than 50 HP" for the group of
small-horsepower induction motors whose ratings are less than 50 hp. The following
equation is then used to adjust both the per-unit first-cycle and interrupting reactances,
where the 3-phase kVA rating is approximately equal to the hp rating for an individual
induction motor or the sum total of hp ratings for a group of motors.
[X.sub.M] = (unadjusted [X.sub.M]) x (base MVA [divided by] MVA rating) x [(kV rating
[divided by] base kV).sup.2] (eq. 11)
The corresponding adjusted first-cycle and interrupting resistances are calculated by using
Equation 8. The results of the calculations are listed in Table 3.
Reducing first-cycle and contact-parting networks to equivalent reactance and resistance
values results in simplified netwrork calculations of fault currents and short-circuit X/R
ratios.
In Part 2 of this series of articles (December 1995 issue), we used a simple example to show
the data preparation steps used in finding the appropriate per-unit reactances and
resistances of power system apparatus for the first-cycle (momentary) and contact-parting
(interrupting) networks. In this last installment, we'll explain how to use this data in
calculating short- circuit current [TABULAR DATA FOR TABLE 1 OMITTED] [TABULAR
DATA FOR TABLE 2 OMITTED] [TABULAR DATA FOR TABLE 3 OMITTED]
(symmetrical rms current). Specifically, we'll explain how to construct the networks from a
one-line diagram, locate the fault connection on the networks, and use basic AC circuit
theory to reduce these networks to equivalent reactance and resistance values. These values,
in turn, will then be used to calculate the first-cycle (momentary) and contact-parting
(interrupting) symmetrical rms currents and short-circuit X/R ratios. Parts 4 and 5 in
future issues will discuss how to use these results in selecting circuit breakers and fuses
from manufacturers' tables.
Simplified reactance and resistance networks

Fig. 1, on page 52, is a single-line diagram of a simplified industrial power system, the same
used in our reader's quiz in Part 2.
Fig. 2, on page 54, presents the steps used to construct four types of simplified networks
from the adjusted per-unit reactances and resistances of Tables 1, 2, and 3: first-cycle
(momentary) reactance network, first-cycle (momentary) resistance network, contactparting (interrupting) reactance network, and contact-parting (interrupting) resistance
network.
In Step A of Fig. 2, we begin the construction of the simplified network by drawing upper
and lower busses. An unlabeled sinusoidal AC voltage source is connected across the busses.
Eventually ([ILLUSTRATION FOR FIGURE 3 OMITTED], on page 58), this voltage source
is labeled the prefault phase-to-neutral voltage at the faulted bus.
Step B is the key step in the construction of a simplified network. For every source of shortcircuit current (i.e., machines and utility), we connect one terminal of its reactance to the
upper bus of the network. We connect the other terminal of the reactance to a new bus
whose label is the bus number from the one-line diagram.
Finally in Step C, we interconnect the remaining reactances of passive apparatus (i.e.,
transformers, feeders, etc.) to match the one-line diagram.
Fault connection on the simplified network
In Fig. 3, we see the three-phase fault location simulated by connecting a jumper from the
faulted bus to the lower reference bus. This diagram shows this procedure for three-phase
faults at Busses 2 and 4. Since the prefault bus voltage magnitudes throughout the power
system are very close to 1.0 per-unit under normal load conditions, we can make a
conservative assumption and select the bus voltage of the sinusoidal AC voltage source to be
greater than or equal to 1.0 per-unit. Also, we set the phase angle of the source at 0
[degrees]; this will serve as the phase angle reference for the power system.
Reduction of simplified network
An important advantage of the simplified reactance (or resistance) network is the ease with
which you can reduce the network to a single equivalent reactance [X.sub.EQ] (or resistance
[R.sub.EQ]). All you need to do is apply the equations shown in Table 4, on page 58, to
combine series and parallel arrangements of reactances. (These equations are the same for
series and parallel arrangements of resistances.) Reactances in series "see" the same current

and are combined by addition (e.g., [X.sub.M4] + [X.sub.T2] and [X.sub.U] + [X.sub.T1]).
Parallel reactances (indicated by double slashes,; e.g. [X.sub.M2]//[X.sub.M3]) have
common bus voltage.
Fig. 4, on page 60, summarizes the procedure used to reduce the simplified reactance
network to an equivalent reactance ([X.sub.EQ]) for a three- phase fault at Bus 4. Note that
in the top network, the three left-most branches are connected in parallel and are combined
into a single reactance in the middle network. The trick is to notice that this single reactance
is in series with [X.sub.T3], and that both these reactances are combined by addition in the
bottom network. Finally, these two reactances in the bottom network are connected in
parallel, and are combined to yield [X.sub.EQ].
In practice, it's easier to use numerical values instead of symbols to reduce the simplified
networks. Symbols are used in Figs. 4 and 5 only to show how to use the equations to do the
reductions. Furthermore, the reduction procedure is identical for any of the four types of
simplified networks.
Table 5, on page 61, lists the numerical results of the equivalent first-cycle (momentary) and
contact-parting (interrupting) reactances and resistances for the three-phase faults at the
major busses of the Fig. 1 one-line. Note that the equivalent contact-parting (interrupting)
reactance and resistance are not calculated at low voltage Bus 3; this is because only the
first-cycle (momentary) results are needed in selecting the short-circuit capabilities of lowvoltage circuit breakers and fuses. Parts 4 and 5 of this series (in future issues) will provide
further information a bout this subject.
Finally the equations used to calculate the first-cycle (momentary) and contact-parting
(interrupting) symmetrical rms currents and short-circuit X/R ratios are given in Table 4.
Final cautionary notes
Once again, you're cautioned not to select circuit breakers and fuses from a manufacturer's
table solely on the basis of symmetrical rms current. There are circumstances when
symmetrical rms current must be adjusted by a multiplying factor to account for
asymmetrical limitations. (These limitations will be discussed in Parts 4 and 5 of this
series.)
We strongly urge you to check the results noted in Table 5. Important note: Just because
you successfully performed these calculations does not certify or qualify you to do these

calculations in your practice. To employ or promote yourself on this basis alone is a serious
breach of professional ethics.
[TABULAR DATA FOR TABLE 5 OMITTED]
SUGGESTED READING
EC&M articles: "How to Perform Short-Circuit Calculations - Parts 1 and 2," June and
December 1995 issues; "The How and Why of Short-Circuit Studies," August 1993 issue.
EC&M books: Practical Guide to Applying Low-Voltage Fuses, Second Edition. To order, call
1-800-543-7771.
IEEE publications: ANSI/IEEE 141-1993, Recommended Practice for Electric Power
Distribution for Industrial Plants (IEEE Red Book); Fault Calculations of Industrial and
Commercial Power Systems, Frank F. Mercede. (IEEE Order Number HL 0459-8). To
order, call 1-800-678-IEEE.
Short-circuit currents represent a tremendous amount of destructive energy, which can be
released through electrical systems under fault conditions. Baseline short-circuit studies
should be performed when the facility electrical system is first designed, and then updated
when a major modification or renovation takes place but no less frequently than every
five years. Major changes would be considered a change in feed by the electric utility, a
change in the primary or secondary system configuration within the facility, a change in
transformer size or impedance, a change in conductor lengths or sizes, or a change in the
motors that are energized by the system.
Every electrical system confines electrical current flow to selected paths by surrounding the
conductors with insulators of various types. Short-circuit current is the flow of electrical
energy that results when the insulation barrier fails and allows current to flow in a shorter
path than the intended circuit.

In normal operations, as shown in Fig. 1, the impedance of the electrical load limits the
current flow to relatively small values. However, a short-circuit path bypasses the normal
current-limiting load impedance, resulting in excessively high current values that are
restricted only by limitations of the power system itself, and by the impedances of the
conductive elements that still remain in the path between the power source and the shortcircuit point (Fig. 2).
Using basic Ohm's Law (E = I Z or I = E Z) as a guide, it's obvious that if the voltage
remains constant and the impedance suddenly decreases, approaching zero, then the
current must simultaneously increase, approaching infinity, to satisfy Ohm's Law.

There are three basic sources of short-circuit current: the electric utility, motors, and on-site
generators. Obviously, the largest source is the electric utility, although the high- and
medium-voltage lines leading to the facility do have finite impedances, as does the utility
service transformer. The second largest source is from motors within a facility.
With today's high fault currents, it's more important than ever to protect electrical
equipment from extremely high current levels. Otherwise, the equipment will explode as it
attempts to interrupt the fault. But for many, fault current calculations have always been
difficult to get a handle on, until now.
Here's a new method to calculate short-circuit currents, one we like to call the Easy Way
kVA Method. You can use in it in place of the abstract per-unit method of short-circuit
calculations from the past. With the kVA method, you can easily visualize what currents will
flow and where they will flow, and you can calculate them using an inexpensive handheld
calculator in moments, regardless of the complexity of the electrical power system.
This method is simple because there are no awkward base changes to make, because kVAs
are the same on both the primary and secondary sides of every transformer. Best of all, you
only need one calculation to determine the short-circuit values at every point within the
entire electrical power system. With the old per-unit method, you needed a separate
calculation for each point in the system.

You can obtain short-circuit kVA values from your electrical utility company, but shortcircuit power is also protected by generators and motors. The kVA produced by a motor is
equal to its starting inrush current. Likewise, the kVA produced by a generator is equal to its
kVA nameplate rating divided by its nameplate subtransient reactance rating (Xd).
For example, suppose we have a 1,000kVA generator with a subtransient rating of 0.15. It
would instantaneously produce 6,667kVA (1,000 0.15). Or, suppose we have a 100-hp
motor with subtransient rating of 0.17. It would instantaneously produce 588kVA (100
0.17).
Now suppose this motor and generator connects to the same bus. Then, the short-circuit
power available at that bus is the sum 6,667kVA plus 588kVA, or 7,255kVA. If the electrical
utility is rated to deliver 100,000kVA to this same bus, then the total short-circuit power
available at that bus is 107,255kVA.
Using the kVA method also greatly simplifies the calculation of short-circuit power
attenuation (or holdback) provided by reactors, transformers, and conductors. For example,
a 2,000kVA, 7% impedance transformer will pass through its windings a maximum of
28,571kVA of power (2,000 0.07), if infinite power flows to one side of its windings. If
instead of an infinite current source, the above bus connects to this transformer, then the
amount of power that will be let through the transformer is the reciprocal of the sum of
the reciprocals of the two, or 1 ( [1 107,254] + [1 28,571] ), or 22,561kVA. You can
determine transformer impedance, reactor impedance, or cable size with the kVA method
quickly enough to make what-if calculations.
Comparisons over several years have found results of the kVA method to be accurate within
3% of computer calculations using expensive software, so you can even use the kVA method
as a check on the input and output of a computer calculation. This is an excellent benefit
because standard engineering procedure requires you to check calculations using a different
method from the one originally used.

You might also like