Electrical Distribution Fundamentals Design Guide Data Bulletin-Schneider Electric
Electrical Distribution Fundamentals Design Guide Data Bulletin-Schneider Electric
Electrical Distribution Fundamentals Design Guide Data Bulletin-Schneider Electric
Guide
Data Bulletin
0100DB2301
03/2024
www.se.com
Legal Information
The information provided in this document contains general descriptions, technical
characteristics and/or recommendations related to products/solutions.
This document is not intended as a substitute for a detailed study or operational and
site-specific development or schematic plan. It is not to be used for determining
suitability or reliability of the products/solutions for specific user applications. It is the
duty of any such user to perform or have any professional expert of its choice
(integrator, specifier or the like) perform the appropriate and comprehensive risk
analysis, evaluation and testing of the products/solutions with respect to the relevant
specific application or use thereof.
The Schneider Electric brand and any trademarks of Schneider Electric SE and its
subsidiaries referred to in this document are the property of Schneider Electric SE or
its subsidiaries. All other brands may be trademarks of their respective owner.
This document and its content are protected under applicable copyright laws and
provided for informative use only. No part of this document may be reproduced or
transmitted in any form or by any means (electronic, mechanical, photocopying,
recording, or otherwise), for any purpose, without the prior written permission of
Schneider Electric.
Schneider Electric does not grant any right or license for commercial use of the
document or its content, except for a non-exclusive and personal license to consult it
on an "as is" basis.
Schneider Electric reserves the right to make changes or updates with respect to or in
the content of this document or the format thereof, at any time without notice.
To the extent permitted by applicable law, no responsibility or liability is
assumed by Schneider Electric and its subsidiaries for any errors or omissions
in the informational content of this document, as well as any non-intended use
or misuse of the content thereof.
Table of Contents
Engineering Fundamentals .............................................................................7
Introduction and Basic Fundamentals ................................................................7
Introduction................................................................................................7
Purpose of this Guide .................................................................................7
Applications of Electrical Power in Industrial and Commercial
Facilities ....................................................................................................7
Basic Design Philosophy.............................................................................8
Electric Power Fundamentals..........................................................................10
Introduction..............................................................................................10
Basic Concepts ........................................................................................10
AC Power ................................................................................................14
Transformers ...........................................................................................17
Basic Electrical Formulae..........................................................................23
Load Planning................................................................................................26
Basic Principles........................................................................................26
NEC Basic Branch Circuit Requirements ....................................................28
NEC Basic Feeder Circuit Sizing Requirements ..........................................31
System Voltage Considerations.......................................................................33
Basic Principles........................................................................................33
Voltage Drop Considerations .....................................................................34
System Arrangements ....................................................................................36
Introduction..............................................................................................36
Radial System..........................................................................................37
Radial System with Primary Selectivity .......................................................37
Expanded Radial Systems ........................................................................38
Loop Systems ..........................................................................................40
Secondary Selective Systems ...................................................................41
Primary Selective Systems........................................................................43
Secondary Spot-network Systems .............................................................43
Ring Bus Systems ....................................................................................44
Composite Systems..................................................................................45
Summary .................................................................................................48
System Grounding .........................................................................................50
Introduction..............................................................................................50
Solidly-grounded Systems.........................................................................51
Ungrounded Systems ...............................................................................54
High-resistance Grounded Systems...........................................................59
Reactance Grounding ...............................................................................63
Low-resistance Grounded Systems ...........................................................63
Creating an Artificial Neutral in an Ungrounded System ...............................64
NEC Systems Grounding Requirements.....................................................65
System Protection ..........................................................................................69
Introduction..............................................................................................70
Characterization of Power System Faults ...................................................70
Methods Of Reducing Short-circuit Current.................................................76
Engineering Fundamentals
Introduction and Basic Fundamentals
Introduction
With the increasing sophistication of modern power systems, it is easy to overlook the
fact that the basic function of a power distribution system has been the same for over
100 years: the safe, reliable distribution of power from a source to the connected
loads. Although this basic function has not changed, the complexity of the loads
themselves, along with today’s reliability and efficiency requirements, makes its
realization more complex.
This guide discusses the main considerations that must be taken into account to
obtain an optimal system design. Because the characteristics of each load, process,
or other issue, are unique, each design is unique to match the requirements imposed.
• Enhanced space economy: The power system footprint is optimized for the
available space.
• Enhanced simplicity: The power system is easy to understand and operate for a
qualified person.
• Reduced cost: The power system costs; both first cost and operating cost, are
low.
• Enhanced power quality: The power system currents and voltages are
sinusoidal, without large harmonics present. System voltage magnitudes do not
change appreciably.
• Enhanced transparency: The power system data at all levels is simply acquired
and interpreted, and the power system is simply interfaced with other building
systems. Enhanced control of the system is also possible.
While it is the goal of every power system design to meet the above basic
considerations, no system design can yield all the enhanced characteristics listed.
The relationship between the considerations listed is shown in Power System Design
Consideration Heuristics, page 9.
Some of the enhanced characteristics mentioned are mutually exclusive, and to
obtain a combination of several enhanced characteristics requires a significant
increase in cost. The design engineer, therefore, must consider the balance between
the performance requirements of the system and the cost, while not compromising the
basic safety elements, functionality, and code compliance.
Enhanced
Enhanced Enhanced Enhanced Enhanced
Power
Maintainability Reliability Flexability Transparency
Quality
Enhanced
Safety
Enhanced
Simplicity Basic Safety, Functionality,
and Code Compliance
Enhanced
Space
Economy
Increasing Cost
Introduction
The fundamentals of electric power and electrical systems have remained unchanged
for many years, though technologies employed in electrical power systems have
advanced greatly. Anyone responsible for any part of an electrical system relies on the
fundamental relationships presented. An understanding of the fundamentals of
electric power is vital to successful power system design. It is assumed that the reader
has a degree in electrical power engineering or electrical power engineering
technology, however the following discussion is presented as a review and reference
material for completeness.
Basic Concepts
Commercial electric power in the United States is generated and delivered as
alternating current, abbreviated as “AC”. AC power consists of sinusoidal voltages
and currents. Mathematically, an AC voltage or current can be expressed as follows:
Where:
• v(t) is an AC voltage
• i(t) is an AC current
• Vmax is the voltage amplitude
• Imax is the current amplitude
• f is the system frequency
• Фv is the voltage phase shift in degrees
• Фt is the current phase shift in degrees
• t is the time in ms
All angles are measured in degrees.
AC currents and voltages are economical to generate and, further, the magnitudes of
the currents or voltages can be stepped up or down using transformers.
Three-phase AC power is the standard in the United States due to its convenience of
generation. Three-phase (abbreviated “3Ф “) power is characterized by three different
phases, each with a phase shift 120° from the other two phases. The three phases are
typically referred to as “A”, “B”, and “C”. Further, the standard frequency for the United
States is 60 Hz. Therefore, three-phase voltages in the United States can be
mathematically described as follows:
Where:
• va(t) is the A-phase voltage
• vb(t) is the B-phase voltage
• vc(t) is the C-phase voltage
The voltages from (2-3) – (2-5) are shown graphically in Graphical Representation of
3Ф Voltages, page 11.
VAN VCN
VBN VBN
VCN VAN
Voltage
Voltage
Time
Time
The peaks of the voltage waveforms are 120° (5.5 ms at 60 Hz) apart. The peak of
phase A occurs before the peak of phase B, which in turn occurs before the peak of
phase C. This is referred to as an ABC phase sequence or ABC phase rotation. If any
two-phase labels are swapped, the result is a CBA phase rotation. Both are
encountered in practice. Also, the definition of time = 0 is arbitrary due to the periodic
nature of the waveforms.
Because the full mathematical representation of AC voltages and currents is not
practical, a shorthand notation is usually used. This shorthand notation treats the
sinusoids as complex quantities based upon the following mathematical relationship:
Where:
• Re is the real part of the complex quantities
The voltage quantities from (2-3) – (2-5) can therefore be rewritten as follows:
To further develop this shorthand notation, it must be recognized that the use of the
RMS (root-mean-square) quantity, rather than the amplitude, is advantageous in
power calculations (discussed below). The RMS quantity for a periodic function f(t) is
defined as follows:
(2–10)
T
1 2
Frms =
T ∫f
0
( t ( dt
Where:
• Frms is the RMS value of the periodic function f(t)
• T is the period of f(t)
Using (2-10), the RMS value of each of the sinusoidal voltages from (2-3) – (2-5) are
calculated as:
Because the RMS value is so useful in the calculation of power-related quantities, any
time an AC voltage or current value is given it is assumed to be an RMS value unless
otherwise stated.
Assuming only the real part of ejϴ is kept, the voltages from (2-7) – (2-9) can be written
as complex quantities known as phasors:
The phasor quantities in (2-15) – (2-17) can be treated as complex quantities for the
purposes of manipulation and calculation, but with the understanding that, if required,
the basic time-domain voltage relationships of (2-3) – (2-5) can easily be obtained.
The phasors can be plotted, as shown in Plot for Phasors per (2-15) – (2-17), page 13.
In most instances the Re and Im axes are omitted since the definition of time zero (and
thus angle zero) is arbitrary; the important information conveyed is the angular
relationships between the phasors themselves. The real part of a phasor is its
projection on the Re axis; if the phasors are imagined to rotate in a counter-clockwise
direction about the 0,0 point it can be seen that the peak of va(t), represented by the
tip of phasor V̄a crossing the Re axis, occurs first, followed by the peak of vb(t),
followed in turn by the peak of vc(t). Thus, for angles defined as positive in the
counter-clockwise direction the ABC phase sequence is indicated by a counter-
clockwise phasor rotation. If angles are defined as positive in the clockwise direction a
clockwise phasor rotation indicates an ABC phase sequence. Both are encountered in
practice. In this guide all angles in phasor diagrams are assumed to be positive in the
counterclockwise direction.
A very general representation of a 3Ф system is shown in General 3Ф System
Representation, page 14.
In Figure 4 the three phases A, B, and C have been labeled, along with the neutral (N)
and ground (G). The neutral is optional, however the ground always exists. The AC
voltages V̄a1, V̄b1, and V̄c1 per the discussion above could represent phase-to-phase
voltages (V̄ab1, V̄bc1, V̄ca1), phase-to-neutral voltage. (V̄an1, V̄bn1, V̄cn1), or phase-to-
ground voltages (V̄ag1, V̄bg1, V̄cg1). The existence of the neutral, and the relationship
between the phases and ground, is dependent upon the system grounding and is
discussed in System Grounding, page 50. A ground current is not defined; this is
because the ground is not intended to carry load current, only ground fault current as
discussed in subsequent sections of this guide. In practice, when 3Ф voltages are
discussed, they are assumed to be phase-to-phase voltages unless otherwise noted.
AC Power
With the basic concepts per above, AC electrical power can be described.
Consider the following DC circuit element:
For the circuit element of DC Circuit Element for Power Calculation, page 14 the
following is true:
P= Vdc·Idc (2–18)
Where:
• Vdc is the DC voltage across the circuit element under consideration, with polarity
as shown.
• Idc is the DC current through the circuit element under consideration, considered
positive for the direction shown.
• P is the power generated by, or dissipated through, the circuit element under
consideration.
The sign of P in (2-18) depends on the direction of current flow with respect to the DC
voltage. A positive value for P indicates power dissipated, while a negative value for P
indicates power generated. DC power is measured in Watts, where one Watt is 1 V x 1
A.
With AC voltages and currents the expression for power is more complex. Assume
that one phase is taken under consideration, with an AC current and voltage as
defined by (1-1) and (1-2) respectively. The expression for the instantaneous power,
after some manipulation, is:
Recall that Vmaxcan be expressed in terms of Vrms per (2-11); substituting Vrmsper (2-
11) into (2-20) yields:
However, the absolute value of the product VrmsIrms cos(Фv - ФI) is always less than
VrmsIrms unless (ᶲv - ФI) = 0. Further, if (ᶲv - ᶲI) = ±90º, as is the case with a purely
inductive or capacitive load. VrmsIrms cos(ᶲv - ФI) = 0. Because energy is required to
force current to flow, and energy is always conserved, AC power must have another
component. This component is most easily defined if AC power is treated as a
complex quantity. To do this, Complex Power S̄ is defined as follows:
S̄ = V̄ ·Ī* (2–22)
The quantities V̄ and Ī are the AC current and voltage in their complex forms per (2-
15) above, with the * operator denoting the complex conjugate, or angle negation, of
the current. This conjugation of the current is done to yield the correct value for the
power angle as described below. Real Power P and Reactive Power Q are defined as
follows:
S̄ = P + jQ (2–23)
P = Re{S̄} = VrmsIrmscos(Фv - Фi) (2–24)
Q = Im{S̄} = VrmsIrmssin(Фv - Фi) (2–25)
If the voltage magnitudes and power factor angles for each phase are equal, the
power quantities per phase can be represented as S̄1Ф, S 1Ф, P1Ф, Q1Ф; equations (2-
27) – (2-30) can then be simplified as:
Transformers
Transformers are vital components for AC power systems. They are used to change
the voltage and current magnitudes to suit the application.
The Ideal Transformer
Transformers are relatively simple devices that utilize Faraday’s law of
electromagnetic induction. In its simplest form, this law can be written:
ξ = -N dψ/dt (2–35)
Where ξ is the voltage induced in a coil of N turns that is linked by a magnetic flux Ψ.
In turn, the magnetic flux Ψ for a coil of N turns through which a current I passes and
linked by a magnetic path with reluctance ℜ can be expressed as:
Ψ = NI/ℜ (2–36)
Consider the simple transformer shown in Basic Transformer Model, page 17:
Equations (2-38) and (2-42) are the basic equations for a single-phase transformer.
The voltage ratio (V1/V2) is equal to the turns ratio (N1/ N2), and the current ratio is
equal to the inverse of the turns ratio. By re-writing (2-38) in terms of the turns ratio
(N1/N2) and substituting into (2-42), the following is obtained:
This is to be expected, since the apparent flowing into the transformer should ideally
equal the apparent power flowing out of the transformer.
The usefulness of the transformer lies in the fact that it can adjust the voltage and
current to the application. For example, on a transmission line it is advantageous to
keep the voltage high to be able to transmit the power with as small a current as
possible, to minimize line losses and voltage drop. At utilization equipment, it is
advantageous to work with low voltages that are more conducive to equipment design
and personnel safety.
Another important aspect of the transformer is that it changes the impedance of the
circuit. For example, if an impedance Z̄2 is connected to winding 2 of the ideal
transformer in Practical Transformer Model, page 19 it can be stated that:
Using (2-38) and (2-42), (2-44) can be written in terms V̄1 and Ī1:
By definition:
The impedance through the transformer is the load impedance at the transformer
output winding multiplied by the square of the turns ratio.
The resistance Rc represents the core losses due to hysteresis, and inductance Lc
represents the magnetizing inductance. Resistances R1 and R2 represent the winding
resistances of winding 1 and winding 2, respectively. Inductances L1 and L2 represent
the leakage inductances of windings 1 and 2, respectively. For quick calculations, the
core losses and magnetizing inductance are often ignored, and the model is treated
as an impedance in series with an ideal transformer.
To denote the proper polarity, the circuit representation for a transformer includes
polarity marks as shown in Standard Transformer Symbolic Representation, page 19.
If the current for one winding flows into its terminal with the polarity mark, the current
for the other winding flow out of its terminal with the polarity mark. In addition, the
ANSI polarity markings per IEEE Standard Terminal Markings and Connections for
Distribution and Power Transformers2 are shown; “H” denotes the higher voltage
winding, and “X” denotes the lower voltage winding.
3Ф Transformer Connections
To be useful in 3Ф systems transformers must be connected for use with 3Ф voltage.
This is accomplished using 3Ф transformer connections.
The wye-wye connection is shown in Wye-Wye Transformer Connection, page 20.
This could be a bank of three single-phase transformers or one 3Ф transformer which
consists of all three sets of windings on a common ferromagnetic core. Polarity
markings for three single-phase transformer connections are shown at the individual
transformers, and polarity markings for a 3Ф transformer are shown next to the A, B,
C, and N terminals.
For both the primary and secondary windings the magnitude of the line-to-line voltage
is equal to the magnitude of the line-to-neutral voltage multiplied by √3. For
convenience, the transformer turns ratio is taken as 1:1 on the phasor diagram.
Where:
• Vdc is the DC voltage across the circuit element under consideration.
• Idcis the current through the circuit element under consideration. Idcis considered
positive if it flows from the circuit element terminal at the higher voltage to the
terminal at the lower voltage.
• Rdcis the DC resistance of the circuit element under consideration, measured in
Ohms.
• P is the power dissipated or generated by the circuit element.
P is the power dissipated or generated by the circuit element. A positive power from
(2-49) indicates power dissipated by the circuit element, and a negative value
indicates power generated by the circuit element. The sign of P in (2-50) is lost due to
the squaring of the current or voltage.
Passive Energy Storage Elements
Capacitors store energy in the form of voltage, with governing equations:
ic = Cdvc/dt (2–51)
E = i/2Cvc2 (2–52)
Where:
• vc is the voltage across the capacitor.
• Ic is the current through the capacitor, considered positive if it flows toward the
terminal from which vc is referenced.
• C is the capacitance value of the capacitor, measured in Farads.
• E is the energy stored in the capacitor.
Inductors store energy in the form of current, with governing equations:
vi = LdiI/dt (2–53)
E = 1/2Li2 (2–54)
Where:
• vI is the voltage across the inductor.
• II is the current through the inductor, considered positive if it flows toward the
terminal from which vI is referenced.
• L is the inductance value of the inductor, measured in Henries.
• E is the energy stored in the inductor.
AC Voltages and Currents, Time-domain Form
Single-phase AC voltage and current can be expressed as follows:
Where:
• vtis an AC voltage.
• II is an AC current.
• Vmax is the voltage amplitude.
• Imax is the current amplitude.
• f is the system frequency.
• Фvis the voltage phase shift in degrees.
• Фi is the current phase shift in degrees.
• t is the time in milliseconds.
All angles are measured in degrees.
If the frequency is 60 Hz, 3Ф voltages can be written as:
Where:
• Va(t) is the A-phase voltage.
• Vb(t) is the B-phase voltage.
• Vc(t) is the C-phase voltage.
The RMS value of a perfectly sinusoidal AC voltage or current is:
V̄ = Vrms∠ Фv (2–65)
Ī̄ = Īrms∠ ФI (2–66)
Where:
• V̄ is the AC voltage in frequency-domain form.
• Ī is the AC current in frequency-domain form
• Фv is the voltage phase shift.
• ФI is the current phase shift.
Capacitor and Inductor impedances in frequency domain form are:
Where:
• Z̄cis the impedance of the capacitor.
• Z̄lis the impedance of the inductor.
• j = √(-1).
• f is the system frequency.
• C is the capacitance of the capacitor .
• L is the inductance of the inductor.
AC voltage and current for an impedance Z̄ are related as follows:
V̄ = Ī·Z̄ (2–69)
S̄ = P + jQ (2–71)
P = Re{S̄} = VrmsIrmscos(Фv - Фi) (2–72)
Q = Im{S̄} = VrmsIrmssin(Фv - Фi) (2–73)
S = |S̄| = √(P2 + Q2 ) (2–74)
For 3Ф circuits the total power is the sum of the power in each phase, that is,
If the current and voltage magnitudes and angles are equal for each phase, (2-76) –
(2-78) can be simplified as follows by considering the power quantities per phase to
be S̄1Ф ,S1Ф, P1Ф, and Q1Ф:
Basic Transformers
The input voltage and current V1 and I1 and the output voltage and current V2 and I2
for an ideal single-phase transformer are related as follows:
The impedance Z̄1at the input terminals of a transformer is related to the load
impedance Z̄2connected to the transformer output terminals by the following equation:
References
Because the subject matter for this section is basic and general to the subject of
electrical engineering, it is included in most undergraduate textbooks on basic circuit
analysis and electric machines. Where material is considered so basic as to be
axiomatic no attempt has been made to cite a particular source for it.
Load Planning
Abstract: Understanding the loads connected to an electrical system is an essential
consideration when designing or operating said system. Determining the size of the
equipment required, including fault interrupting devices, bus bars, conductors, and
similar, is not just a summation of connected load nameplates. Herein, considerations
and practices are presented to facilitate load planning to ensure adequate sizing is
accomplished while not over-sizing and increasing electrical system infrastructure
costs.
Additional information on motor characteristics, motor load, sizing requirements and
motor protection are found in the Fundamentals of Motor Control Design Guide.
Basic Principles
The most vital, but often the last to be acquired, pieces of information for power
system design are the load details. An important concept in load planning is that due
to non-coincident timing, some equipment operating at less than rated load, and some
equipment operating intermittently rather than continuously, the total demand upon
the power source is always less than the total connected load (see IEEE Standard
Terminal Markings and Connections for Distribution and Power Transformers5). This
concept is known as load diversity. The following standard definitions are given in
IEEE Standard Terminal Markings and Connections for Distribution and Power
Transformers5 and Electric Power Distribution System Design, New York6and are
tools to quantify it:
Demand: The electric load at the receiving terminals averaged over a specified
demand interval. of time, usually 15 min., 30 min., or one hour based upon the
particular utility’s demand interval. Demand may be expressed in amperes,
kiloamperes, kilowatts, kilovars, or kilovolt amperes.
Demand Interval: The period over which the load is averaged, usually 15 minutes, 30
minutes, or one hour.
Peak Load: The maximum load consumed or produced by a group of units in a stated
period of time. It may be the maximum instantaneous load or the maximum average
load over a designated period of time.
Maximum Demand: The greatest of all demands that have occurred during a
specified period of time such as one-quarter, one-half, or one hour. For utility billing
purposes the period of time is generally one month.
Coincident Demand: Any demand that occurs simultaneously with any other
demand.
Demand Factor: The ratio of the maximum coincident demand of a system, or part of
a system, to the total connected load of the system, or part of the system, under
consideration, that is:
Diversity Factor: The ratio of the sum of the individual maximum demands of the
various subdivisions of a system to the maximum demand of the whole system, that
is:
(3–2)
Where:
• Di = maximum demand of load i, regardless of time of occurrence.
• Dg = coincident maximum demand of the group of n loads.
Using (1), the relationship between the diversity factor and the demand factor is:
(3–3)
Where:
• TCLi = total connected load of load group i.
• DFi = the demand factor of load group i.
Load Factor: The ratio of the average load over a designated period of time to the
peak load occurring in that period, that is:
If T is the designated period of time, an alternate formula for the load factor may be
obtained by manipulating (3-4) as follows:
These quantities must be used with each type of load to develop a realistic picture of
the actual load requirements if the economical sizing of equipment is to be achieved.
Further, they are important to the utility rate structure (and thus the utility bill).
As stated in Electric Power Distribution System Design, New York7, the following must
be considered in this process:
• Load Development/Build-up Schedule: Peak load requirements, temporary/
construction power requirements, and timing.
• Load Profile: Load magnitude and power factor variations expected during low-
load, average load, and peak load conditions.
• Expected Daily and Annual Load Factor.
• Large motor starting requirements.
• Special or unusual loads such as resistance welding, arc welding, induction
melting, induction heating.
• Harmonic-generating loads such as variable-frequency drives, arc discharge
lighting.
• Forecasted load growth over time.
Reference IEEE Recommended Practice for Electric Power Systems in Commercial
Buildings8: and individual engineering experience on previous projects are both useful
in determining demand factors for different types of loads. In addition, the National
Electrical Code® 9 gives minimum requirements for the computation of branch circuit,
feeder, and service loads.
Unit Load
Type of Occupancy
VA/ m2 VA/ ft.2
Automotive facility 16 1.5
Convention center 15 1.4
Courthouse 15 1.4
Dormitory 16 1.5
Exercise center 15 1.4
Fire station 14 1.3
Gymnasiuma 18 1.7
Health care clinic 17 1.6
Hospital 17 1.6
Hotels and motels, including apartment houses without 18 1.7
provisions for cooking by tenantsb
Library 16 1.5
Manufacturing facilityc 24 2.2
Motion picture theater 17 1.6
Museum 17 1.6
Officed 14 1.3
Parking garagee 3 0.3
Penitentiary 13 1.2
Performing arts theater 16 1.5
Police station 14 1.3
Post office 17 1.6
Religious facility 24 2.2
Restaurantf 16 1.5
hg
Retail, , 20 1.9
School/university 33 3
Sports arena 33 3
Town hall 15 1.4
Transportation 13 1.2
Warehouse 13 1.2
Workshop 18 1.7
NOTE: The 125% multiplier for a continuous load as specified in 210.20(A) is included when using
the unit loads in this table for calculating the minimum lighting load for a specified occupancy.
a Armories and auditoriums are considered gymnasium-type occupancies.
b Lodge rooms are similar to hotels and motels.
Motor Loads (Article 220.14(C)): Motor loads must be calculated in accordance with
Articles 430.22, 430.24, and 440.6, summarized as follows:
• The full load current rating for a single motor used in a continuous duty
application is 125% of the motor’s full-load current rating as determined by Article
430.6, which refers to horsepower/ ampacity Tables 430.247, 430.248, 430.249,
or 430.250 as appropriate (Article 430.22).
• The load calculation for several motors, or a motor(s) and other loads, is 125% of
the full load current rating of the highest rated motor per a.) in Table 1 plus the
sum of the full-load current ratings of all the other motors in the group, plus the
ampacity required for the other loads (Article 430.24).
• For hermetic refrigerant motor compressors or multi-motor equipment employed
as part of air conditioning or refrigerating equipment, use the equipment
nameplate rated load current instead of the motor horsepower rating (Article
440.6).
Luminaires (lighting fixtures) (Article 220.14(D)): Calculate an outlet supplying
luminaire(s) based on the maximum VA rating of the equipment and lamps for which
the luminaire(s) is rated.
Heavy-duty Lampholders (Article 220.14(E)): Calculate loads for heavy-duty
lampholders at a minimum of 600 VA.
Sign and Outline Lighting (Article 220.14(F)): Calculate sign and outline lighting
loads at a minimum of 1200 VA for each required branch circuit specified in article
600.5(A).
Show Windows (Article 220.14(G)): Calculate show windows in accordance with
either:
• The unit load per outlet as required in other provisions of article 220.14.
• 200 VA per 300 millimeters (one foot) of show window.
Loads for Fixed Multi-outlet Assemblies: Calculate these in other than dwelling
units or the guest rooms and guest suites of hotels or motels as follows (Article 220.14
(H)):
• Where simultaneous use of appliances is unlikely, each 1.5 meters (five feet.) or
fraction thereof of each separate and continuous length, must be considered as
one outlet of 180 VA.
• Where simultaneous use of appliances is unlikely, each 300 millimeters (1 foot) or
fraction thereof must be considered as an outlet of 180 VA.
Receptacle Outlets (Articles 220.14(I), 220.14(J), 220.14(K), 220.14(L)): Loads for
these are calculated as follows:
• Dwelling occupancies (Article 220.14(J)): In one-family, two-family, and
multifamily dwellings, the minimum unit load shall be not less than 33 VA/m2 (3
VA/ft2). The lighting and receptacle outlets specified in 220.14(J)(1), (J)(2), and
(J)(3) are included in the minimum unit load. No additional load calculations are
required for such outlets. The minimum lighting load is determined using the
minimum unit load and the floor area as determined in 220.11 for dwelling
occupancies. Motors rated less than 1⁄8 hp and connected to a lighting circuit are
considered part of the minimum lighting load.
(1) All general-use receptacle outlets of 20 A rating or less, including receptacles
connected to the circuits in 210.11(C)(3) and 210.11(C)(4).
(2) The receptacle outlets specified in 210.52(E) and (G).
(3) The lighting outlets specified in 210.70.
• Office buildings (Article 220.14(K)): Calculate receptacle outlets to be the larger
of either the calculated value per c.) below or 11 VA/m2 (one VA/per square foot).
• All other receptacle outlets (Article 220.14(I)): Calculate each receptacle on one
yoke as 180 VA. Calculate a multiple receptacle consisting of four or more
receptacles at 90 VA per receptacle.
Sufficient Branch Circuits: Incorporate sufficient branch circuits into the system
design to serve the loads per Article 220.10 (summarized 1.) – 8.) above), along with
branch circuits for any specific loads not covered in Article 220.10. Determine the total
number of branch circuits from the calculated load and the size or rating of the branch
circuits used. Evenly proportion the load among the branch circuits (Article 210.11(C)).
In addition, Article 210.11(C) requires several dedicated branch circuits as follows for
dwelling units:
• Two or more 20 A small-appliance branch circuits (Article 210.11(C)(1)).
• One or more 20 A laundry branch circuits (Article 210.11(C)(2)).
• One or more bathroom branch circuits (Article 210.11(C)(3)).
Continuous Loads (Article 210.20): The branch circuit overcurrent protection must
be at least the sum of the non-continuous load + 125% of the continuous load, unless
the overcurrent device is 100%-rated. Because the overcurrent protection rating
determines the rating of the branch circuit (Article 210.3), the branch circuit must be
sized for the non-continuous load + 125% of the continuous load. In load calculations,
continuous loads should therefore be multiplied by 1.25 unless the circuit overcurrent
device is 100% rated. Motor loads are not included in this calculation as the 125%
factor is already included in the applicable sizing per above.
Type of Occupancy Portion of Lighting Load to Which Demand Demand Factor (%)
Factor Applies VA/
Hotels and motels, including apartment houses without provision First 20,000 or less at 60
for cooking by tenants a From 20,001 to 100,000 at 50
Remainder over 100,000 at 35
• Show Window or Track Lighting (Article 220.43): Show windows must use a
calculated value of 660 VA per linear meter (200 VA per linear foot), measured
horizontally along its base. Track lighting in other than dwelling units must be
calculated at an 150 VA per 660 millimeters (two feet) of lighting track or fraction
thereof. Where multi-circuit track is installed, the load is divided equally between
the track circuits.
• Receptacles in Other than Dwelling Units (Article 220.44): Demand factors for
non-dwelling receptacle loads are given in Table 3 (NEC Table 220.44) and Table
2 (NEC Table 220.42)..
Portion of Lighting Load to Which Demand Factor Applies VA Demand Factor (%)
Motors (Article 220.50): Motor loads shall be calculated in accordance with 430.24,
430.25, and 430.26. The feeder demands for these are calculated as follows:
• The load calculation for several motors, or a motor(s) and other loads, is 125% of
the full load current rating of the highest rated motor per the second bullet of
Motor Loads (Article 220.14(C)) in NEC Basic Branch Circuit Requirements,
page 28 plus the sum of the full-load current ratings of all the other motors in the
group, plus the ampacity required for the other loads (Article 430.24).
• Base the load calculation for factory-wired multimotor and combination-load
equipment on the minimum circuit ampacity marked on the equipment (Article
430.25) instead of the motor horsepower rating.
• Where allowed by the authority having jurisdiction, feeder demand factors may be
applied based on the duty cycles of the motors. No demand factors are given in
the NEC for this situation.
Fixed Electric Space Heating (Article 220.51): Calculate the feeder loads for these
at 100% of the connected load. However, in no case shall a feeder or service load
current rating be less than the rating of the largest branch circuit supplied.
Noncoincident Loads (Article 220.60): While unlikely that two or more non-
coincident loads will be in use simultaneously, it is permissible to use only the largest
loads for use at one time in calculating the feeder demand.
Feeder neutral load (Article 220.61): The feeder neutral load is defined as the
maximum load imbalance on the feeder. The maximum load imbalance for three-
phase four-wire systems is the maximum net calculated load between the neutral and
any one ungrounded conductor. A service or feeder supplying the following loads such
as a feeder or service supplying household electric ranges, wall-mounted ovens,
counter-mounted cooking units, and electric dryers and the unbalanced load in excess
of 200 A shall be permitted to have an additional demand factor of 70%. Refer to NEC
article 220.61 for neutral reductions in systems other than three-phase, four-wire
systems. This demand factor does not apply to non-linear loads; in fact, it may be
necessary to oversize the neutral due to the current flow from non-linear load triple
harmonics.
Continuous Loads (Article 215.3): The rating of the overcurrent protection for a
feeder circuit must be at least the sum of the non-continuous load + 125% of the
continuous load, unless the overcurrent device is 100%-rated. Because the rating of
the overcurrent protection determines the rating of the branch circuit (Article 210.3),
size the branch circuit for the non-continuous load + 125% of the continuous load. In
the final feeder circuit load calculation, the continuous portion of the load should
therefore be multiplied by 1.25, unless the overcurrent device for the circuit is 100%-
rated. Motor loads are not included in this calculation as the 125% factor is already
included in the applicable sizing per above.
Additional calculation data is given in NEC Article 220 for dwelling units, restaurants,
schools, and farms. This data is not repeated here.
As this guide only presents the basic NEC requirements for load calculations, it is
imperative to refer to the NEC itself when in doubt about a specific load sizing
application. Computer programs are commercially available to automate the
calculation of feeder and branch circuit loads per the NEC methodology described
above.
Basic Principles
The selection of system voltages is crucial to successful power system design.
Reference American National Standard Preferred Voltage Ratings for Electric Power
Systems and Equipment (60Hz)12 lists the standard voltages for the United States and
their ranges. The nominal voltages from American National Standard Preferred
Voltage Ratings for Electric Power Systems and Equipment (60Hz)12 are given in
Standard Nominal Three-phase System Voltages per ANSI C84.1-1989, page 33.
ANSI C84.1-1989 divides system voltages into “voltage classes”. Voltages 600 V and
below are referred to as “low voltage”, voltages from 600 V - 69 kV are referred to as
“medium voltage”, voltages from 69 kV - 230 kV are referred to as “high voltage”, and
voltages 230 kV - 1,100 kV are referred to as “extra high voltage”, with 1,100 kV also
referred to as “ultra-high voltage”. The emphasis of this guide is on low- and medium-
voltage distribution systems.
The choice of service voltage is limited to those voltages that the serving utility
provides. In most cases, only one choice of electrical utility is available, and thus only
one choice of service voltage. As the power requirements increase, so to does the
likelihood that the utility will require a higher service voltage for a given installation. In
some cases, a choice may be given by the utility as to the service voltage desired, in
which case an analysis of the various options is required to arrive at the correct
choice. In general, the higher the service voltage the more expensive the equipment
required to accommodate it is. Maintenance and installation costs also increase with
increasing service voltage. However, equipment such as large motors may require a
service voltage of 4160 V or higher, and, further, service reliability tends to increase at
higher service voltages.
Another factor to consider regarding service voltage, is the voltage regulation of the
utility system. Voltages defined by the utility as “distribution” should, in most cases,
have adequate voltage regulation for the loads served. Voltages defined as
“subtransmission” or “transmission”, however, often require the use of voltage
regulators or load-tap changing transformers at the service equipment to give
adequate voltage regulation. This situation typically only occurs for service voltages
above 34.5 kV, however it can occur on voltages between 20 kV and 34.5 kV. When in
doubt, consult the serving utility.
The utilization voltage is determined by the requirements of the served loads. For
most industrial and commercial facilities this is 480Y/277 V, although 208Y/120 V is
also required for convenience receptacles and small machinery. Large motors may
require 4160 V or higher. Distribution within a facility may be 480Y/277 V or, for large
distribution systems, medium-voltage distribution may be required. Medium-voltage
distribution implies a medium-voltage (or higher) service voltage, and results in higher
costs of equipment, installation, and maintenance than low-voltage distribution.
However, this must be considered along with the fact that medium-voltage distribution
generally results in smaller conductor sizes and makes control of voltage drop easier.
Power equipment ampacity limitations impose practical limits on the available service
voltage to serve a given load requirement for a single service, as shown in Equipment
Design Limits to Service Voltage vs. Load Requirements, for a Single Service, page
35.
Where:
• V̄drop is the voltage drop along a length of conductor or across a piece of
equipment in V,
• Īl is the load current in A, and
• Z̄cond is the conductor or equipment impedance, in Ohms.
Thus, the larger the load current and larger the conductor impedance, the larger the
voltage drop. Unbalanced loads, of course, give an unbalanced voltage drop, which
leads to an unbalanced voltage at the utilization equipment.
Section 210.19(A) – branch circuits – Informational Note N° 3 recommends the
voltage drop at the farthest outlet of power, heating, and lighting, or combination of
such loads, to three percent of the applied voltage. Alternatively, the maximum
combined voltage drop on the feeder and branch circuits to the farthest outlet should
be five percent.
Section 215.2(A)(1) – feeders – Informational Note N° 2 has the same
recommendations for feeders.
Those statements mean that the feeder could have a one percent voltage drop if the
branch circuit had no more than four percent. Also, limiting the branch circuit voltage
drop to three percent allows a two percent drop in the feeder. These or any other
combinations of feeder and branch circuit voltage drops not exceeding a total of five
percent are adequate.
A voltage drop of five percent or less from the utility service to the most remotely-
located load is recommended by NEC article 210.19(A)(1), FPN No. 4. Because this is
a note only, it is not a requirement per se but is the commonly accepted guideline.
Table 5 - Equipment Design Limits to Service Voltage vs. Load Requirements, for a Single Service
208 1,800
480 Switchboard or Low-Voltage
5000 4,157
600 Power Switchgear
5,196
\
2,400 4,489
4,160 1080 7,782
4,800 8,979
6,900 8,605
8,320 10,376
12,000 14,965
12,470 720 15,551
Metal-Enclosed Switchgear,
13,200 16,461
w/Fuses
13,800 17,210
20,780 6,299
22,860 6,929
23,000 175 6,972
24,940 7,560
2,400 12,471
4,160 21,616
4,800 24,942
6,900 38,853
8,320 3000 43,232
12,000 62,354
12,470 64,796
13,200 Metal-Clad Switchgear 68,589
13,800 71,707
71,984
20,780
79,189
22,860
2000 79,674
23,000
24,940
86,395
Because conductor impedance increases with the length of the conductor, unless the
power source is close to the center of the load, the voltage varies across the system,
and, further, it can be more costly to maintain the maximum voltage drop across the
system to within five percent of the service voltage since larger conductors must be
used to offset longer conductor lengths.
Also, from equation (4-1) as load changes, so does the voltage drop. For a given
maximum load, a measure of this change at a given point is the voltage regulation,
defined as:
Where:
• V̄no load is the voltage, at a given point in the system, with no load current flowing
from that point to the load.
• V̄load is the voltage, at the same point in the system, with full load current flowing
from that point to the load.
Another source of concern when planning for voltage drop is the use of power-factor
correction capacitors. Because these serve to reduce the reactive component of the
load current, they also reduce the voltage drop per equation (4-1).
Both low and high voltage conditions, and voltage imbalance, have an adverse effect
on utilization equipment (see IEEE Recommended Practice for Electric Power
Distribution for Industrial Plants13 for additional information). Voltage drop must
therefore be considered during power system design to avoid future problems.
System Arrangements
Abstract: The electrical point of interconnection with a utility can vary in voltage level
whether it be secondary, primary, or transmission voltages. The reliability of an
electrical system is directly affected by the system arrangement and the voltage level
to which it is connected. Additionally, the need for redundancy or serviceability without
a complete shutdown are also considerations when evaluating various system
arrangements.
Introduction
The selection of system arrangement has a profound impact upon the reliability and
maintainability of the system. Several commonly used system topologies are
presented here, along with the pros and cons of each. The figures for each of these
assume that the distribution and utilization voltage are the same, and that the service
voltage differs from the distribution/utilization voltage. The symbology (low voltage
circuit breaker, low-voltage drawout circuit breaker, medium voltage switch, medium
voltage breaker) reflects the most commonly-used equipment for each arrangement.
The symbology used throughout this section is shown in Symbology, page 36.
Figure 15 - Symbology
Radial System
The radial system is the simplest system topology and is shown in Radial System,
page 37. It is the least expensive in terms of equipment first cost. However, it is also
the least desirable since it incorporates only one utility source and the loss of the utility
source, transformer, or the service or distribution equipment results in a loss of
service. Further, the loads must be shut down to perform maintenance on the system.
This arrangement is most used where the need for low first-cost, simplicity, and space
economy outweigh the need for enhanced reliability.
Typical equipment for this system arrangement is a single unit substation consisting of
a fused primary switch, a transformer of sufficient size to supply the loads, and a low-
voltage switchboard.
An automatic transfer scheme may optionally be provided between the two primary
switches to automatically switch from an offline utility source to an available source.
Most often metal-clad circuit breakers are used, rather than metal-enclosed switches.
More about typical equipment application guidelines follows in a subsequent section
of this guide.
Figure 18 - Expanded Radial System with One Utility Source and a Single
Primary Feeder
Figure 19 - Expanded Radial system with One Utility Source and Multiple
Primary Feeders
Expanded Radial System with Two Utility Sources and Multiple Primary Feeders,
page 40 shows an expanded radial system utilizing multiple substations and two utility
sources, again with metal-clad primary switchgear but with a duplex metal-enclosed
switchgear for utility source selection.
Figure 20 - Expanded Radial System with Two Utility Sources and Multiple
Primary Feeders
Of the arrangements discussed this far, the arrangement of Expanded Radial System
with Two Utility Sources and Multiple Primary Feeders, page 40 is the most
dependable; it does not rely on a single utility source for system availability, nor does
the loss of one transformer or feeder cause a loss of service to the entire facility.
However, the loss of a transformer or feeder results in the loss of service to a part of
the facility. More dependable system arrangements are required if this is to be
avoided.
Loop Systems
The loop system arrangement is one of several arrangements that allows one system
component, such as a transformer or feeder cable, to fail without causing a loss of
service to a part of the facility.
Primary Loop System, page 41 shows a primary loop arrangement. The advantages
of this arrangement over previously mentioned arrangements are that a loss of one
feeder cable does not cause one part of the facility to experience a loss of service and
that one feeder cable can be maintained without causing a loss of service (An outage
to part of the system is experienced after the loss of a feeder cable until the loop is
switched to accommodate the loss of the cable).
In Primary Loop System, page 41 metal-clad circuit breakers are used as the feeder
protective devices. Fused metal-enclosed-feeder switches could be utilized for this,
but take care if this is considered since the feeder fuses must be able to serve both
transformers and the feeder and transformer fuses have to coordinate for maximum
selectivity.
Composite Systems
The above system arrangements are the basic building blocks of power distribution
system topologies but are rarely used alone for a given system. To increase system
reliability, it is usually necessary to combine two or more of these arrangements. For
example, one commonly used arrangement is shown in Composite System: Primary
Loop/Secondary Selective, page 46.
A fault on a primary loop cable or the loss of one transformer can be accommodated
without loss of service to either load bus (but with an outage to part of the system until
the system is switched to accommodate the loss). In addition, a single section of the
primary loop or one transformer can be taken out of service while maintaining service
to the loads.
The system of Composite System: Primary Loop/Secondary Selective, page 46 can
be expanded by the addition of an additional utility source and a primary bus tie
breaker to form an even more dependable system, as shown in Composite System:
Primary Selective/Primary Loop/Secondary Selective, page 47. With this
arrangement, the loss of a single utility source, a single primary circuit breaker, a
single loop feeder cable, or a single transformer can be accommodated without loss of
service. And any one primary circuit breaker, any one section of the primary
distribution loop, or any one transformer can be taken out of service without loss of
service to the loads. However, the cost of a second utility service and two additional
metal-clad breakers must be considered.
Summary
Various system arrangements are presented in this section, starting with the least
complex and progressing to a very complex, robust system arrangement. In general,
as reliability increases so does complexity and cost. Economic considerations usually
dictate how complex a system arrangement can be used, and thus have a great deal
of impact on system reliability. Power System Arrangements Summary for the Basic
Arrangements in this Section, page 49 and Power System Arrangements Summary
for the Composite Arrangements in this Section, page 50 show the features of each
system arrangement given in this section.
The formulas given in these tables are for the systems as shown in the earlier figures.
They hold true for expanded versions of these system arrangements where the
expansion is made symmetrically with respect to the configuration shown. They do not
hold true when modifications are made to the system arrangements with respect to
symmetry, with altered numbers of switching/protective devices, or for concurrent loss
of different types of system components. When in doubt regarding a system which is
derived from, but not identical, to the systems shown in the earlier figures, double-
check these numbers.
From a maintenance perspective, the number of system elements that can be taken
down for maintenance is the same as the number that can fail while maintaining
service to the loads.
These tables do not attempt to address concurrent losses of different types of system
components, nor are they always mean a loss of service to a particular load after a
component loss while the system is being switched to an alternate configuration.
However, they are a guide to the relative strengths and weaknesses of each of the
system arrangements presented.
Table 6 - Power System Arrangements Summary for the Basic Arrangements in this Section
Arrangement Utility Losses Primary Breaker Primary Feeder Transformer Secondary Main/ Cost
Allowed Losses Allowed Losses Allowed Losses Allowed Tie Breaker
Losses Allowed
Radial 0 0 0 0 0 $
Expanded Radial, 0 0 0 0 0 $$
Single Primary
Feeder
Expanded Radial, 0 0 0 0 0 $$
Multiple Primary
Feeders
Expanded Radial, u-1a, page 49 0 0 0 0 $$+
Multiple Utility
Sources, Multiple
Primary Feeders
Secondary- 0 0 0 1 1 $$$
Selective System
Secondary Spot u-1 a,c,e,f PB-1 a,c,e,f F-1 a, page 49,c,Δ,f T-1 SB-1 a,c,e,f $$$$$
Network a,c,e,f
Notes:
U = Number of Utility Sources a Assumes that each utility source has sufficient capacity to supply the entire
system.
PB = Number of Primary Circuit Breakers
b Assumes that all secondary circuit breakers, including feeder breakers, are
SF = Number of Primary Feeders interchangeable.
T = Number of Transformers c Assumes that each primary main and bus tie (if applicable) circuit breakers has
sufficient capacity to supply the entire system.
SB = Number of Secondary Main and Tie Circuit Breakers
d Assumes that all primary circuit breakers, including feeder breakers, are
$ = Relative Cost, with $=Least Expensive interchangeable.
e Assumes that each primary feeder has sufficient capacity to supply the entire
system.
f Assumes that each transformer, secondary main and bus tie (if applicable)
circuit breaker have sufficient capacity to supply the entire system.
g Assumes that the ring bus has sufficient capacity to supply the entire system.
Table 7 - Power System Arrangements Summary for the Composite Arrangements in this Section
Arrangement Utility Losses Primary Breaker Primary Feeder Transformer Secondary Main/ Cost
Allowed Losses Allowed Losses Allowed Losses Allowed Tie Breaker
Losses Allowed
Primary Double u-1 a,c PB-F/2-ua,c,d F/2 T-1 f T-1 f,b $$$$$$$$$$
Selective /
Secondary
Selective
Primary Ring Bus u-1 a,c,e,f PB-F/2-u+1 a,c,d,g F/2 T-1 f T–1 f,b $$$$$$$$$$
/ Primary
Selective/
Secondary
Selective
Notes:
U = Number of Utility Sources a Assumes that each utility source has sufficient capacity to supply the entire
system.
PB = Number of Primary Circuit Breakers
b Assumes that all secondary circuit breakers, including feeder breakers, are
SF = Number of Primary Feeders interchangeable.
T = Number of Transformers c Assumes that each primary main and bus tie (if applicable) circuit breakers has
sufficient capacity to supply the entire system.
SB = Number of Secondary Main and Tie Circuit Breakers
d Assumes that all primary circuit breakers, including feeder breakers, are
$ = Relative Cost, with $=Least Expensive interchangeable.
e Assumes that each primary feeder has sufficient capacity to supply the entire
system.
f Assumes that each transformer, secondary main and bus tie (if applicable)
circuit breaker have sufficient capacity to supply the entire system.
g Assumes that the ring bus has sufficient capacity to supply the entire system.
System Grounding
Abstract: System grounding considerations affect many aspects of an electrical
system. Knowledge of the various types of system grounding and performance
characteristics is critical when designing or operating an electrical system. The
voltage, system arrangement, loads connected, and continuity of service drive
grounding requirements and design choices.
Introduction
The topic of system grounding is extremely important, as it affects the susceptibility of
the system to voltage transients, determines the types of loads the system can
accommodate, and helps to determine the system protection requirements.
The system grounding arrangement is determined by the grounding of the power
source. For commercial and industrial systems, the types of power sources generally
fall into four broad categories:
• Utility Service: The system grounding is usually determined by the secondary
winding configuration of the upstream utility substation transformer.
• Generator: The system grounding is determined by the stator winding
configuration.
• Transformer: The system grounding on the system fed by the transformer is
determined by the transformer secondary winding configuration.
• Static Power Converter: For devices such as rectifiers and inverters, the system
grounding is determined by the grounding of the output stage of the converter.
All categories fall under the NEC definition for a “separately-derived system”. The
recognition of a separately derived system is important when applying NEC
requirements to system grounding, as discussed below.
All the power sources mentioned above, except Static Power Converter, are
magnetically operated devices with windings. To understand the system voltage
relationships with respect to system grounding, it must be recognized that there are
two common ways of connecting device windings: wye and delta. These two
arrangements, with their system voltage relationships, are shown in Wye and Delta
Winding Configurations and System Voltage Relationships, page 51. As can be seen
from the figure, in the wye-connected arrangement there are four terminals, with the
phase-to-neutral voltage for each phase set by the winding voltage and the resulting
phase-to-phase voltage set by the vector relationships between the voltages. The
delta configuration has only three terminals, with the phase-to-phase voltage set by
the winding voltages and the neutral terminal not defined.
Neither of these arrangements is inherently associated with any system grounding
arrangement, although some arrangements more commonly used, for reasons that
are explained further below.
Wye
Delta
Solidly-grounded Systems
The solidly-grounded system is the most common system arrangement, and one of
the most versatile. The most commonly used configuration is the solidly-grounded
wye, because it supports single-phase phase-to-neutral loads.
The solidly-grounded wye system arrangement can be shown by considering the
neutral terminal from the wye system arrangement in Wye and Delta Winding
Configurations and System Voltage Relationships, page 51 to be grounded. This is
shown in Solidly-grounded Wye System Arrangement and Voltage Relationships,
page 52.
The delta arrangement can be configured in another manner, however, that does have
merits as a solidly-grounded system. This arrangement is shown in Figure 34. While
the arrangement of Center-Tap-grounded Delta System Arrangement and Voltage
Relationships, page 53 may not appear at first glance to have merit, this system is
suitable both for three-phase and single-phase loads, so long as the single-phase and
three-phase load cables are kept separate from each other. This is commonly used for
small services which require both 240 Vac three-phase and 120/240 Vac single-
phase. The phase A voltage to ground is 173% of the phase B and C voltages to
ground. This arrangement requires the BC winding to have a center tap.
Winding Circuit
Configuration Model
Ungrounded Systems
This system grounding arrangement is at the other end of the spectrum from solidly-
grounded systems. An ungrounded system is a system where there is no intentional
connection of the system to ground.
In Ungrounded Delta System with a Ground-current Trip on One Phase, page 57, the
effects of a single phase to ground current trip are shown. The equations in
Ungrounded Delta System Winding Arrangement and Voltage Relationships, page 55
are not immediately practical for use, however, if the fault impedance is assumed to
be zero and the system capacitive charging impedance is assumed to be much larger
than the phase impedances, these equations reduce into a workable form.
Ungrounded Delta System with a Ground-current Trip on One Phase, page 57 shows
the resulting equations and shows the current and voltage phase relationships.
As seen from Ungrounded Delta System: Simplified Ground-current Trip Voltage and
Current Relationships, page 58, the net result of a ground current trip on one phase of
an ungrounded delta system is a change in the system phase-to-ground voltages. The
phase-to-ground voltage on the faulted phase is zero, and the phase-to-ground
voltage on the unfaulted phases are 173% of their nominal values. This has
implications for power equipment; the phase-to-ground voltage rating for equipment
on an ungrounded system must be at least equal the phase-to-phase voltage rating.
This also has implications for the methods used for ground detection, as explained
later in this guide.
Winding Circuit
Configuration Model
Figure 38 - Ungrounded Delta System: Simplified Ground-current Trip Voltage and Current Relationships
Ground
Fault
The ground currents with one phase is faulted to ground are essentially negligible.
Because of this fact, from an operational standpoint ungrounded systems have the
advantage of being able to remain in service if one phase is faulted to ground.
However, suitable ground detection must be provided to alarm this condition (and is
required in most cases by the NEC, see The National Electrical Code16 as described
below). In some older facilities, it has been reported that this type of system has
remained in place for 40 years or more with one phase grounded. This condition is not
unsafe in and of itself (other than due to the increased phase-to-ground voltage on the
unfaulted phases), however, if a ground current trip occurs on one of the ungrounded
phases the result is a phase-to-phase fault with its characteristic large fault current
magnitude.
Another important consideration for an ungrounded system is its susceptibility to large
transient overvoltages. These can result from a resonant or near-resonant condition
during ground current trips, or from arcing (see IEEE Recommended Practice for
Grounding of Industrial and Commercial Power Systems17). A resonant ground
current trip condition occurs when the inductive reactance of the ground-current trip
path approximately equals the system capacitive reactance to ground. Arcing
introduces the phenomenon of current-chopping, which can cause excessive
overvoltages due to the system capacitance to ground.
The ground detection mentioned above can be accomplished using voltage
transformers connected in wye-broken delta, as illustrated in A Ground Detection
Method for Ungrounded Systems, page 59.
16. NFPA 70, The National Fire Protection Association, Inc., 2005 Edition.
17. IEEE Std. 142-1991, December 1991.
In A Ground Detection Method for Ungrounded Systems, page 59, three ground
detection lights “LTA”, “LTB”, and “LTC” are connected so that they indicate the A, B,
and C phase-to-ground voltages, respectively. A master ground detection light “LTM”
indicates a ground current trip on any phase. With no ground current trip on the
system, “LTA”, “LTB”, and “LTB” glow dimly. If a ground current trip occurs on one
phase, the light for that phase is extinguished and “LTM” glows brightly along with the
lights for the other two phases. Control relays may be substituted for the lights if
necessary. Resistor “R” is connected across the broken-delta voltage transformer
secondaries to minimize the possibility of ferroresonance. Most ground detection
schemes for ungrounded systems use this system or a variant thereof.
The ground detection per A Ground Detection Method for Ungrounded Systems, page
59 indicate on which phase the ground current trip occurs, but not where in the system
the ground current trip occurs. This, along with the disadvantages of ungrounded
systems due to susceptibility to voltage transients, was the main impetus for the
development of other ground system arrangements.
Modern power systems are rarely ungrounded due to the advent of high-resistance
grounded systems as discussed below. However, older ungrounded systems are
occasionally encountered.
The resistor is sized to be less than or equal to the magnitude of the system charging
capacitance to ground. If the resistor is thus sized, the high-resistance grounded
system is usually not susceptible to the large transient overvoltages that an
ungrounded system can experience. The ground resistor is usually provided with taps
to allow field adjustment of the resistance during commissioning.
If no ground current trip current is present, the phasor diagram for the system is the
same as for a solidly-grounded wye system, as shown in High-resistance Grounded
System with No Ground Present, page 60. However, if a ground current trip occurs on
one phase the system response is as shown in High-resistance Grounded System
with a Ground-current Trip on One Phase, page 61. As seen from High-resistance
Grounded System with a Ground-current Trip on One Phase, page 61, the ground
fault current is limited by the grounding resistor. If the approximation is made that that
Z̄A and Z̄F are very small compared to the ground resistor resistance value R, which is
a good approximation if the fault is a bolted ground fault, then the ground fault current
is approximately equal to the phase-to-neutral voltage of the faulted phase divided by
R. The faulted phase voltage to ground in that case would be zero and the unfaulted
phase voltages to ground would be 173% of their values without a ground current trip
present. This is the same phenomenon exhibited by the ungrounded system
arrangement, except that the ground current fault current is larger and approximately
in-phase with the phase-to-neutral voltage on the faulted phase. The limitation of the
ground current trip current to such a low level, along with the absence of a solidly-
grounded system neutral, has the effect of making this system ground arrangement
unsuitable for single-phase line-to-neutral loads.
Winding Circuit
Configuration Model
Ground
Fault
The ground current trip current is not large enough to force its removal by taking the
system off-line. Therefore, the high-resistance grounded system has the same
operational advantage in this respect as the ungrounded system. However, in addition
to the improved voltage transient response as discussed above, the high-resistance
grounded system has the advantage of allowing the location of a ground current trip to
be tracked.
A typical ground detection system for a high-resistance grounded system is illustrated
in Pulsing Ground Detection System, page 62. The ground resistor is shown with a tap
between two resistor sections R1 and R2. When a ground current trip occurs, relay 59
(the ANSI standard for an overvoltage relay, as discussed later in this guide) detects
the increased voltage across the resistor. It sends a signal to the control circuitry to
initiate a ground current trip alarm by energizing the “alarm” indicator. When the
operator turns the pulse control selector to the “ON” position, the control circuit causes
pulsing contact P to close and re-open approximately once per second. When P
closes R2 is shorted and the “pulse” indicator is energized. R1 and R2 are sized so
that approximately five to seven times the resistor continuous ground current trip
current flows when R2 is shorted. The result is a pulsing ground current trip current
that can be detected using a clamp-on ammeter (an analog ammeter is most
convenient). By tracing the circuit with the ammeter, the ground current trip location
can be determined. Once the ground current trip has been removed from the system
pressing the “alarm reset” button de-energizes the “alarm” indicator.
This type of system is known as a pulsing ground detection system and is very
effective in locating ground current trips but is generally more expensive than the
ungrounded system ground current trip indicator in High-resistance Grounded System
with No Ground Present, page 60.
Ground
Fault
Reactance Grounding
In industrial and commercial facilities, reactance grounding is commonly used in the
neutrals of generators. In most generators, solid grounding may permit the level of
ground-fault current available from the generator to exceed the three-phase value for
which its windings are braced (see IEEE Recommended Practice for Grounding of
Industrial and Commercial Power Systems18). For these cases, grounding of the
generator neutral through an air-core reactance is the standard solution for lowering
the ground current trip level. This reactance ideally limits the ground-fault current to
the three-phase available fault current and allows the system to operate with phase-
to-neutral loads.
The zig-zag transformer only passes ground current. Its typical implementation on an
ungrounded system, to convert the system to a high-resistance grounded system, is
shown in Zig-zag Grounding Transformer Implementation, page 65. The zig-zag
transformer distributes the ground current ĪG equally between the three phases. For all
practical purposes the system, from a grounding standpoint, behaves as a high-
resistance grounded system.
Qualified Person: One who has the skills and knowledge related to the construction
and operation of the electrical equipment and installations and has received safety
training on the hazards involved.
With these terms defined, several of the major components of the grounding system
can be illustrated by redrawing the system of NEC System Grounding Terms
Illustration, page 67 and labeling the components.
Unground Conductor
Unground Conductor
Grounded Conductor
Ungrounded Conductor
20. NFPA 70, The National Fire Protection Association, Inc., 2005 Edition.
21. IEEE Std. 142-1991, December 1991.
• Certain systems are permitted, but not required, to be solidly grounded. They are
listed as electric systems used exclusively to supply industrial electric furnaces
for melting, refining, tempering, and the like, separately derived systems used
exclusively for rectifiers that supply only adjustable-speed industrial drives, and
separately derived systems supplied by transformers that have a primary voltage
rating less than 1000 V provided that certain conditions are met [Article 250.21].
• If a system 50 - 1000 Vac is not solidly grounded, ground detectors must be
installed on the system unless the voltage to ground is less than 120 V [Article
250.21].
• Certain systems cannot be grounded. They are listed as circuits for electric
cranes operating over combustible fibers in Class III locations as provided in
Article 503.155, circuits within hazardous (classified) anesthetizing locations and
other isolated power systems in health care facilities as provided in 517.61 and
517.160, circuits for equipment within electrolytic cell working zone as provided in
Article 668, and secondary circuits of lighting systems as provided in 411.5(A)
[Article 250.22]. Some of the requirements for hazardous locations and health
care facilities are covered in Electrical Energy Management, page 220.
• For solidly-grounded systems, an unspliced main bonding jumper must be used
to connect the equipment grounding conductor(s) and the service disconnect
enclosure to the grounded conductor within the enclosure for each utility service
disconnect [Article 250.24(B)].
• For solidly-grounded systems, an unspliced system bonding jumper must be
used to connect the equipment grounding conductor of a separately derived
system to the grounded conductor. This connection must be made at any single
point on the separately derived system from the source to the first system
disconnecting means or overcurrent device [250.30(A)(1)].
• A grounding connection on the load side of the main bonding or system bonding
jumper on a solidly-grounded system is not permitted [Articles 240.24(A)(5),
250.30(A)]. The reasons for this are explained in below and in Arc Flash Hazard,
page 150.
• Ground fault protection of equipment must be provided for solidly grounded wye
electrical services, feeder disconnects on solidly-grounded wye systems, and
building or structure disconnects on solidly-grounded wye systems under the
following conditions:
1. The voltage is greater than 150 V to ground but does not exceed 600 V
phase-to-phase.
2. The utility service, feeder, or building or structure disconnect is rated 1000 A
or more.
3. The disconnect in question does not supply a fire pump or continuous
industrial process.[Articles 215.10, 230.95, 240.13].
• Where ground current trip protection is required per Article 215.10 or 230.95 for a
health care facility, an additional step of ground current trip protection is required
in the next downstream device toward the load, except for circuits on the load
side of an essential electrical system transfer switch and between on-site
generating units for the essential electrical system and the essential electrical
system transfer switches [Article 517.17]. Specific requirements for health-care
systems are described in Emergency Power Distribution Equipment, page 184.
• The alternate source for an emergency or legally-required standby system is not
required to have ground current trip protection. For an emergency system,
ground-fault indication is required [Articles 700.26, 701.17]. Emergency Power
Distribution Equipment, page 184 describes the requirements for Emergency and
Standby Power Systems.
• All electrical equipment, wiring, and other electrically conductive material must be
installed in a manner that creates a permanent, low-impedance path facilitating
the operation of the overcurrent device. This circuit must have the ability to carry
the ground current trip current imposed upon it. [Article 250.4(A)(5)]. The intent of
this requirement is to allow ground current trip current magnitudes to be sufficient
for the ground current trip protection/detection to detect (and for ground current
trip protection to clear) the fault, and for a ground current trip not to cause
damage to the grounding system.
• High-impedance grounded systems may be utilized on AC systems of 480 - 1000
V where:
1. Conditions of maintenance and supervision so that only qualified persons
access the installation.
2. Continuity of power is required.
3. Ground detectors are installed on the system.
4. Line-to-neutral loads are not served. [Article 250.36]
• For systems over 1000 V:
1. The system neutral for solidly-grounded systems may be a single point
grounded or multigrounded neutral. Additional requirements for each of
these arrangements apply [Article 250.184].
2. The system neutral derived from a grounding transformer may be used for
grounding [Article 250.182].
3. The minimum insulation level for the neutral of a solidly-grounded system is
600 V. A bare neutral is permissible under certain conditions [Article 250.184
(A) (1)].
4. Impedance grounded neutral systems may be used where conditions 1, 3,
and 4 for the use of high-impedance grounding on systems 480-1000 V
above are met [Article 250.186].
5. The neutral conductor must be identified and fully insulated with the same
phase insulation as the phase conductors [Article 250.186 (B)].
• Zig-zag grounding transformers must not be installed on the load side of any
system grounding connection [Article 450.5].
• When a grounding transformer is used to provide the grounding for a three-phase
four-wire system, the grounding transformer must not be provided with
overcurrent protection independent of the main switch and common-trip
overcurrent protection for the three-phase, four-wire system [Article 450.5 (A)
(1)]. An overcurrent sensing device must be provided that causes the main switch
or common-trip overcurrent protection to open if the load on the grounding
transformer exceeds 125% of its continuous current rating [Article 450.5 (A) (2)].
Again, these points are not intended to be an all-inclusive reference for NEC
grounding requirements. They do, however, summarize many of the major
requirements. When in doubt, consult the NEC.
System Protection
Abstract: To protect personnel, equipment, and maintain continuity of service for an
electrical system, protection or fault interrupting devices are required. Adequate
system designs allow for the system to withstand and isolate faults while not causing
additional damage and/or outages. System protection is paramount and must be
understood by all persons interacting or responsible for electrical systems.
Introduction
An important consideration in power system design is system protection. Without
system protection, the power system itself, which is intended to be of benefit to the
facility in question, would itself become a hazard.
The major concern for system protection is protection against the effects of
destructive, abnormally high currents. These abnormal currents, if left unchecked,
could cause fires or explosions resulting in risk to personnel and damage to
equipment. Other concerns, such as transient overvoltages, are also considered when
designing power system protection although they are generally considered only after
protection against abnormal currents has been designed.
The Thevenin impedance for a power system at a given point in the system is referred
to as the short-circuit impedance. In the great majority of power systems, the short-
circuit impedance is predominately inductive, therefore one simplification that is often
made is to treat the impedance purely as inductance. This has the effect of causing
the fault current to lag the system line-to-neutral voltage by 90º. If the system is an
ungrounded delta system, the equivalent line-to-neutral voltage is obtained by
performing a delta-wye conversion of the source voltage.
The phase-to-phase fault value is calculated from the three-phase fault value, if it is
remembered that the line-to-line voltage magnitude is equal to the line-to-neutral
voltage magnitude multiplied by √3 and that there will be twice the impedance in the
circuit since the return path must be considered. These two facts, taken together,
allow computation of the line-to-line fault current magnitude |Ī(fl-l)| as:
3 I f 3ø (7–1)
I fI I =
2
This, however, is as far as this simplified analysis method take us. To further
characterize fault currents, a method for calculating unbalanced faults must be used.
The universally-accepted method for this is a method known as the method of
symmetrical components.
In the method of symmetrical components, unbalanced currents and voltages are
broken into three distinct components: positive sequence, negative sequence, and
zero sequence. These sequence components are thought of as independent sets of
rotating balanced phasors. The positive sequence set rotates in the standard A-B-C
phase rotation. The negative sequence set rotates in the negative or C-B-A phase
rotation. In the zero sequence, set all three phase components are in phase with one
another. The positive, negative and zero sequence components are further simplified
by referring only to the A-phase phasor of each set; these are referred to as V̄1for the
positive sequence set, V̄2for the negative sequence set and V̄ 0 for the zero-sequence
set. For a given set of phase voltages V̄a, V̄b , V̄c, the sequence components are
related to the phase voltages as follows:
Where:
a = 1∠120∘
The system may be separated into positive, negative, and zero-sequence networks
depending upon the fault type and the resulting sequence quantities, then combined
per (7-5), (7-6), and (7-7) to yield the phase values.
Modern short-circuit analysis is performed using computers. Even large systems are
quickly analyzed via short-circuit analysis software. Even so, some heuristic benefit
can be gained by knowing how the method of symmetrical components works. For
example, certain protective relays are often set in terms of negative-sequence values
and ground currents are often referred to as zero-sequence quantities in the literature.
Another factor that must be considered is the existence of DC quantities in fault
currents. Because of the system inductance, the current cannot change
instantaneously, therefore upon initiation of a fault the system must go through a
transient condition that bridges the gap between the faulted and unfaulted conditions.
This transition involves DC currents. For a generic single-phase AC circuit with an
open-circuit voltage v(t)=Vmsin(( ωt+θ)) and a short-circuit impedance consisting of
resistance R and inductance L, the fault current for a fault initiated at time t = 0 can be
expressed as in Electrical Transients in Power Systems22:
Where:
ϕ = tan-1(ωL/R)
The angle ϕ can be recognized to be the angle of the Thevenin impedance. Several
key points can be taken from (7-8):
• When the fault occurs such that (θ-ϕ) = 0 no transient occurs. For a purely
inductive circuit this would mean that θ = 90º and thus the fault is initiated when
the voltage is at its peak.
• When the fault occurs such that (θ-ϕ) = 90º the maximum transient occurs. For a
purely inductive circuit this would mean that θ = 0º and thus the fault is initiated
when the voltage is zero.
• The time constant of the circuit is (L/R) and thus the higher the value of L/R the
longer the transient lasts. Instead of (L/R) power systems typically are defined in
terms of (X/R), where (X/R) is the ratio of the inductive reactance of the short-
circuit impedance to its resistance. Thus, the higher (X/R) or the “X/R ratio”, the
longer the short-circuit transient lasts. This has great implications on the rating of
equipment.
A typical plot of fault current on a distribution system with a low X/R ratio and closing
angle such that a small transient is produced is shown in Fault Current for System
with Low X/R Ratio and Small-transient Closing Angle, Normalized to a Steady-state
22. Alan Greenwood, New York, John Wiley and Sons Inc., 1971.
Magnitude of 1, page 73. In contrast with this is the plot shown in Fault Current for
System with Higher X/R Ratio and Closing Angle of 0, Normalized to a Steady-state
Magnitude of 1, page 74, which is the fault current for a system with a high X/R ratio
and closing angle of 0 such that there is a large transient.
Figure 48 - Fault Current for System with Low X/R Ratio and Small-transient Closing Angle, Normalized to a
Steady-state Magnitude of 1
Figure 49 - Fault Current for System with Higher X/R Ratio and Closing Angle of 0, Normalized to a Steady-
state Magnitude of 1
Steady-state Component of Waveform in Figure 49, page 75 shows only the steady-
state component of the waveform of Fault Current for System with Higher X/R Ratio
and Closing Angle of 0, Normalized to a Steady-state Magnitude of 1, page 74 and
Transient Component of Waveform in Figure 49, page 75 shows only the transient
component.
The fault current is often described in terms of its RMS Symmetrical and RMS
Asymmetrical values. The RMS symmetrical value is the RMS value considering the
steady-state component only. The RMS asymmetrical value is the RMS value over the
first cycle considering both the steady-state and transient components at the worst-
Asymmetry Factor = 1 + 2e ( )
2!
X
R
(7–9)
For example, this asymmetry factor for an X/R ratio of 25 is 1.6, meaning that the
approximate worst-case RMS asymmetrical value over the first cycle for the fault
current at an X/R ratio of 25 is no greater than the RMS symmetrical value multiplied
by 1.6.
For motors and generators, which have a high X/R ratios, calculations for the transient
performance during a fault are simplified by representing the short-circuit impedances
differently for different time periods after the fault initiation. The reactive component of
the short-circuit impedance for the first half-cycle into the fault is the subtransient
reactance (X”d). For the first several cycles into the fault the reactance is larger and is
termed the transient reactance (X’d . For the long-term fault, currents (up to 30 cycles
or so into the fault) the reactance is even larger and is termed the synchronous
reactance (Xd. The synchronous reactance is much larger than either the transient or
subtransient reactance and models the phenomenon of AC decrement; after the DC
component decays the AC component continues to decay, eventually reaching a
value that can be less than the generator rated load current.
In general, the closer the fault is to a generator or generators the higher the X/R ratio
and thus the larger the DC offset. The AC decrement of the fault from generator
sources is pronounced. Faults from most utility services are sufficiently far removed
from generation and have enough resistance in the distribution lines that there is less
DC offset and essentially no AC decrement. The fault current contribution from
induction motors has a high DC offset but also decays rapidly to zero over the first few
cycles since there is no applied field excitation. The fault current contribution from
synchronous motors has a large DC component and decays to zero but at a slower
rate than for induction motors due to the applied field excitation. For a given point in
the system, the fault current is the sum of the contributions from all these sources and
the DC offset, DC decay, and AC decrement are all dependent upon the relative
location of the fault with respect to these sources.
The existence of the transient is of vital importance to selecting the proper equipment
for system protection. Because standards for equipment short-circuit ratings vary
(more is stated regarding this in subsequent sections of this guide), even more speed
and efficiency is gained by using a computer for short circuit calculations; the various
equipment rating standards can be considered to produce accurate results for
comparison with the equipment ratings.
11 kV 11 kV
Grid 01 Grid 01
381.051 MVAsc 381.051 MVAsc
Bus 01 Bus 01
11 kV 11 kV
Transformer 01 Transformer 01
2 MVA 2 MVA
11/0.415 kV 11/0.415 kV
6.25 %Z 6.25 %Z
Bus 02
0.415 kV Bus 02
0.415 kV
Cable 01 Cable 01
AL 30 m 30 m
AL
3-1/C 300 3-1/C 300
Lighting transformers:
A lighting transformer is nothing but a 1:1 transformer. The transformer turns ratio
is the number of turns of the primary winding divided by the number of turns of the
secondary coil.
The trip current in bus 03 is 24.997 kA, and the cable length is 30 meters. However,
because the breaking current of a low voltage circuit breaker is 20 kA, a lighting
transformer has been installed between the cable and bus 03. The trip current is now
within 20 kA of the safe limit. The trip current on bus 03 is now 10.707 kA. The trip
current is reduced as shown in Lighting Transformer, page 78.
11 Kv 11 Kv
Grid 01 381.051 MVAsc Grid 2 381.051 MVAsc
Bus 01
11 kV
Bus 01
11 kV
Transformer 01
Transformer 01 2 MVA
2 MVA 11/0.415 kV
11/0.415 kV
6.25 %Z
Bus 02
6.25 %Z
0.415 kV
Bus 02 AL Cable 01
0.415 kV 30 m
3-1/C 300
Cable 01 Transformer 01
AL 30 m 500 kVA
3-1/C 300 0.415/0.415 kV
4 %Z
24.997 kA
Bus 03 24 10.707 kA
0.415 kV .99
7k Bus 03 10.7
A 07 kA
0.415 kV
There are numerous approaches available to minimize the trip current in a medium
voltage system. One option is to use a current-limiting reactor.
Current limiting reactor (CLR):
Although the CLR introduces impedance into the circuit degrading the voltage profile
during normal operation, it can be a cost-effective solution obviating the need for
upgrading system switchgear due to an increase in the trip level. A current limiting
reactor is a series reactor connected into the circuit for limiting trip current. Above the
result is a trip current limit the some values.
Applications of the CLR for limiting trip level are varied. One of the more attractive
applications is often the bus section application (see Application of CLR, page 79).
For this arrangement, the CLR is placed between two bus bars to connect them
together.
Disadvantages include:
• Voltage drop
• Increasing power loss
• Reduced grid power factor
• High space requirements
For the system bus 02, as shown in Add a Generator, page 80, the trip current is
19.318 kA under normal conditions. However, as the system power increases to 15
MW for internal purposes by employing one generator, the trip current has increased
to 24.672 kA, as shown in Add a Generator, page 80. How to reduce the fault current?
Whenever we have use some methods, that are given below.
11 kV 11 kV
Grid 01 Grid 01
3810.512 MVAsc 3810.512 MVAsc
Bus 01 Bus 01
110 kV 110 kV
11 kV
926.2 A
15 MW
Gen 01
Trafo 01 19 %Xd”
Trafo 01
45 MWA
45 MWA
110/11 kV
110/11 kV
12.5 %Z
12.5 %Z
The trip current on bus 02 is from 24.672 kA to 22.35 kA as a result of the above
configuration (see Figure 55). The trip current is reduced if the reactor impedance is
increased as shown in With Current Limiting Reactor, page 81.
110 kV 110 kV
Grid 01 Grid 01
3810.512 MVAsc 3810.512 MVAsc
Bus 01 11 kV Bus 01 11 kV
110 kV 926.2 A 110 kV 926.2 A
15 MW 15 MW
Gen 01 Gen 01
110 kV
110 kV Grid 01
Grid 01 3810.512 MVAsc
3810.512 MVAsc
11 kV
Bus 01 Bus 01 926.2 A
110 kV 11 kV 110 kV 15 MW
926.2 A Gen 01
15 MW
Gen 01
19 %Xd”
5.355 kA
Trafo 01 19 %Xd” Trafo 01
45 MWA 45 MWA
110/11 kV 110/11 kV Trafo 02
12.5 %Z 12.5 %Z 500 kVA
11/11 kV
4 %Z
Bus 02 19.318 kA 5.355 kA
Bus 02 19.318 kA 0.632 kA
11 kV
24 11 kV 19
.67 .8 74
2k k
A A
T 300 Vac 0–1200 A 200,000 A Yes UL 248-15-2000, CSA Similar to Class J, but
C22.2 NO 248.15-2000 dimension-ally smaller
600 Vdc 0–1200 A 200,000 A Yes
a * Because of their interchangeability with Class H fuses, class K-1, K-5, and K-9 fuses cannot be marked as “current-limiting”.IEEE
Recommended Practice for Protection and Coordination of Industrial Power Systems26
In some cases, the fuse average melting time only is given. This can be treated as the
fuse opening time with a tolerance of ±15%. The -15% boundary is the minimum
melting time and the +15% boundary is the total clearing time.
The time-current characteristic does not extend below .01 seconds. This is since
below .01 seconds the fuse is operating in its current-limiting region and the fuse I2t is
of increasing importance.
The time-current characteristic curves are used to demonstrate the coordination
between protective devices in series. The basic principle of system protection is that
for a given fault current ideally only the device nearest the fault opens, minimizing the
effect of the fault on the rest of the system. This principle is known as selective
coordination and is analyzed with the use of the device time-current characteristic
curves.
As an example, consider a 480 V system with two sets of fuses in series, with a
system available trip current of 30,000 A. Bus “A” uses a 400 A class J fuses which
supply, among others, bus “B”. Bus “B” uses a 100 A class J fuses. Coordination
between the 400 A and 100 A fuses is shown via the time-current curves of Fuse
Coordination Example, page 86, along with a one-line diagram of the part of the
system under consideration. Because the time bands for the two fuses do not overlap,
these are coordinated for all operating times above .01 seconds. It can be stated that
these two sets of fuses are coordinated through approximately 4200 A, since at 4200
A Fuse A has the potential to begin operating in its current-limiting region. Fuse B has
the potential to begin operating its current-limiting region at 1100 A. For currents
above approximately 4000 A, therefore, both sets of fuses have the potential to be
operating in the current-limiting region. When both sets of fuses are operating the
current-limiting region the time-current curves cannot be used to the determine
coordination between them. Instead, for a given fault current the minimum melting I2t
for Fuse A must be greater than the maximum clearing I2t for Fuse B. In practice,
instead of publishing I2t data fuse manufacturers typically publish ratio tables showing
the minimum ratios of fuses of a given type that coordinate with each other.
Low-voltage fuse AC interrupting ratings are based upon a maximum power factor of
.2, corresponding to a maximum X/R ratio of 4.899. To evaluate a low-voltage fuse’s
interrupting rating on a system with a higher X/R ratio the system symmetrical fault
current must be multiplied by a multiplying factor (see IEEE Recommended Practice
for Protection and Coordination of Industrial Power Systems27):
Where:
(X/R)actualis the actual system X/R.
(X/R)testis the test X/R.
The available symmetrical fault current multiplied by the multiplying factor per (7-10)
can be compared to the fuse interrupting rating.
The use of fuses requires a holder and a switching device in addition to the fuses
themselves. Because they are single-phase devices, a single blown fuse from a three-
phase set causes a loss of phase condition, which can lead to motor damage.
Replacing fuses typically requires opening equipment doors and/or removing cover
panels. Also, replacement fuses must be stocked to get a circuit with a blown fuse
back on-line quickly. For these reasons, the use of low-voltage fuses in modern power
systems is generally discouraged.
28. NFPA 70, The National Fire Protection Association, Inc., 2020 Edition.
29. IEEE Std. 242-2001, December 2001.
30. UL 489, Underwriter’s Laboratories Inc., October 24,2016.
Control voltage: The control voltage rating is the AC or DC voltage designated for
application to control devices intended to open or close a circuit breaker. In most
cases this only applies to accessories that are custom ordered, such as motor
operators.
Interrupting rating: This is the highest current at rated voltage that the circuit breaker
is intended to interrupt under standard test conditions.
Short-time or withstand rating: This characterizes the circuit-breaker’s ability to
withstand the effects of short-circuit current flow for a stated period. Molded-case
circuit breakers typically do not have a withstand rating, although some newer-design
breakers do.
Instantaneous override: A function of an electronic trip circuit breaker that causes
the instantaneous function to operate above a given level of current if the
instantaneous function characteristic has been disabled.
Current limiting circuit breaker: This is a circuit breaker which does not employ a
fusible element and, when operating in its current-limiting range, limits the let-through
I2t to a value less than the I2t of a ½-cycle wave of the symmetrical prospective
current.
HID: This is a marking that indicates that a circuit breaker has passed additional
endurance and temperature rise tests to assess its ability for use as the regular
switching device for high intensity discharge lighting. Per NEC 240.83 (D) a circuit
breaker, which is used as a switch in an HID lighting circuit must be marked as HID.
HID circuit breakers can also be used as switches in fluorescent lighting circuits. SWD
marked circuit breakers may not be used as switches for HID circuits.
SWD: This is a marking that indicates that a circuit breaker has passed additional
endurance and temperature rise tests to assess its ability for use as the regular
switching device fluorescent lighting. Per NEC 240.83 (D) a circuit breaker which is
used as a switch in an SWD lighting circuit must be marked as SWD or HID.
Frame: The term Frame is applied to a group of circuit breakers of similar
configuration. Frame size is expressed in amperes and corresponds to the largest
ampere rating available in that group.
Thermal-magnetic circuit breaker: This type of circuit breaker contains a thermal
element to trip the circuit breaker for overloads and a faster magnetic instantaneous
element to trip the circuit breaker for short circuits. On many larger thermal-magnetic
circuit breakers the instantaneous element is adjustable.
Electronic trip circuit breaker: An electronic circuit breaker contains a solid-state
adjustable trip unit. These circuit breakers are extremely flexible in coordination with
other devices.
Sensor: An electronic-trip circuit breaker’s sensor is usually an air-core current
transformer (CT) designed specifically to work with that circuit breaker’s trip unit. The
sensor size, in conjunction with the rating plug, determines the electronic-trip circuit
breaker’s continuous current rating.
Rating plug: An electronic trip circuit breaker’s rating plug can vary the circuit
breaker’s continuous current rating as a function of its sensor size.
Typical molded-case circuit breakers are shown in Molded-case Circuit Breakers,
page 89. In Molded-case Circuit Breakers, page 89 on the left is a thermal-magnetic
circuit breaker, and on the right is an electronic-trip circuit breaker. The thermal-
magnetic circuit breaker is designed for cable connections and the electronic circuit
breaker is designed for bus connections, but neither type is inherently suited for one
connection type over another. Note the prominently-mounted operating handle on
each circuit breaker.
Circuit breakers can be mounted in stand-alone enclosures, in switchboards, or in
panelboards (more information on switchboards and panelboards is given in a later
section of this guide).
A further reduction in the let-through energy for a fault in the region between two
electronic-trip circuit breakers can be accomplished through zone-selective
interlocking. This consists of wiring the two trip units such that if the downstream
circuit breaker senses the fault (typically this is based upon the short-time pickup) it
sends a restraining signal to the upstream circuit breaker. The upstream circuit
breaker continues to time out as specified on its characteristic curve, tripping if the
downstream device does not clear the fault. However, if the downstream device does
not sense the fault and the upstream devices does, the upstream device does not
have the restraining signal from the downstream device and trips with no intentional
delay. For example, if zone-selective interlocking were present in the system of
Typical Molded-case Circuit Breaker Coordination, page 92 and fault occurs on bus C
circuit breaker B senses the fault and send a restraining signal to circuit breaker A.
Circuit breaker A is coordinated with circuit breaker B, so circuit breaker B trips first. If
circuit breaker B does not clear the fault, circuit breaker A times out on its time-current
characteristic per Typical Molded-case Circuit Breaker Coordination, page 92 and
trips. If the fault occurs at bus B, circuit breaker B does not detect the fault and thus
does not send the restraining signal to circuit breaker A. Circuit breaker A senses the
fault and trips with no intentional delay, which is faster than dictated by its time-current
characteristic per Typical Molded-case Circuit Breaker Coordination, page 92. Care
must be used when applying zone-selective interlocking where there are multiple
sources of power and fault currents can flow in either direction through a circuit
breaker.
Table 9 - Typical Characteristics of Molded-case Circuit Breakers for Commercial and Industrial Applications
Frame Size (A) Number of Poles Interrupting Rating at AC Voltage (kA, RMS symmetrical)
800, 1000 3 42 30 22
65 50 25
200 100 65
1200 3 42 30 22
3 65 50 25
3 200 100 65
1600, 2000 3 65 50 42
3 125 100 65
3000, 4000 3 100 100 85
3 200 150 100
The continuous current rating is set by the sensor and rating plug sizes for a given
electronic-trip circuit breaker. This can be smaller than the frame size. As can be seen
from Typical Characteristics of Molded-case Circuit Breakers for Commercial and
Industrial Applications, page 93, more than one interrupting rating can be available for
a given frame size.
Molded-case circuit breakers are tested for interrupting capabilities with test X/R ratios
as shown in AC Test Circuit Characteristics for Molded-case Circuit Breakers, page 93
(see Molded-Case Circuit Breakers, Molded Case Switches and Circuit-Breaker
Enclosures,32). As with fuses, when a circuit breaker is applied in a circuit with an X/R
ratio larger than its test X/R, then multiply the available RMS symmetrical fault current
by the multiplying factor per equation (7-10) for comparison with the circuit breaker
interrupting rating.
Current-limiting circuit breakers are also available. Coordination between two current-
limiting circuit breakers when they are both operating in the current limiting range is
typically determined by test.
Low-voltage molded case circuit breakers are not maintainable devices. Loss of a
component generally requires replacement of the entire circuit breaker unless the
circuit breaker has been specifically designed for maintainability.
Magnetic-only circuit breakers which have only magnetic tripping capability are
available. These are often used as short-circuit protection for motor circuits
(discussed in more detail in Surge Protection, page 128). For this reason, these are
often referred to as motor circuit protectors.
Molded case switches are also available. These do not have a thermal element;
however, most have a magnetic element which opens the switch above a specified
current to shield the switch from damage due to lack of a short-time rating.
Molded-case circuit breakers are available with several different options, such as
stored-energy mechanisms, key interlocks, motor operators. Refer to specific
manufacturer’s literature for details.
Because the switching means is included with the device, molded-case circuit
breakers give inherent flexibility of operation. This allows circuits to be reclosed
without removing cover panels and exposing the operator to hazardous voltages.
Three-pole circuit breakers are used for three-phase circuits, alleviating the concern
for single-phasing. Circuit breakers are not one-time devices, eliminating the need to
store spares in the event of a fault. These characteristics make molded-case circuit
breakers very versatile protective devices.
rating (see IEEE Standard for Low-Voltage AC Power Circuit Breakers Used in
Enclosures35).
Table 11 - Preferred Ratings for Low-voltage AC Power Circuit Breakers with Instantaneous Direct-acting
Phase Trip Elements a
System Nominal Rated Maximum Insulation Three-phase Frame Size (A) Range of trip device
Voltage (V) Voltage (V) (dielectric) Short-circuit current ratings (A) c
Withstand (V) Current Rating
(V) (A) b
Table 11 - Preferred Ratings for Low-voltage AC Power Circuit Breakers with Instantaneous Direct-acting
Phase Trip Elements a (Continued)
b Ratings in this column are RMS symmetrical values for single-phase (two pole) circuit breakers and three-phase average RMS symmetrical
values of three-phase (three-pole) circuit breakers. When applied on systems where rated maximum voltage may appear across a single pole,
the short-circuit current ratings are 87% of these values. See 5.6 in IEEE Std C37.13-1990.
c The continuous-current-carrying capability of some circuit-breaker-trip-device combinations may be higher than the trip-device current rating.
See 10.1.3 in IEEE Std C37.13-1990.
IEEE Recommended Practice for Protection and Coordination of Industrial Power Systems36
Table 12 - Preferred Ratings for Low-voltage AC Power Circuit Breakers Without Instantaneous Direct-acting
Phase Trip Elements a
Rated Maximum Voltage Frame Size (A) Range of Trip Device Current Ratings (A) b
(V)
Setting of Short-time Delay Trip Element
Table 12 - Preferred Ratings for Low-voltage AC Power Circuit Breakers Without Instantaneous Direct-acting
Phase Trip Elements a (Continued)
IEEE Recommended Practice for Protection and Coordination of Industrial Power Systems37
As with molded-case circuit breakers, low-voltage power circuit breakers are tested at
a given power factor. The test power factor is 15% for unfused circuit breakers and
20% for fused circuit breakers. Short Circuit Multiplying Factors for Low-voltage
Power Circuit Breakers, page 97 shows the multiplying factors for both fused and
unfused circuit breakers for various short-circuit power factors. The multiplying factors
for unfused circuit breakers are calculated similarly to those for molded-case circuit
breakers, but those for fused circuit breakers are based upon RMS rather than peak
current and differ slightly from the multiplying factors obtained from equation (7-10)
(see IEEE Standard for Low-Voltage AC Power Circuit Breakers Used in
Enclosures38).
Table 13 - Short Circuit Multiplying Factors for Low-voltage Power Circuit Breakers
System Short-Circuit Power System X/R Ratio Multiplying Factor x RMS Multiplying Factor x RMS
Factor Symmetrical Short-Circuit Symmetrical Short-Circuit
Current, for Unfused Power Current, for Fused Power
Circuit Breakers Circuit Breakers
20 4.9 1.00 1.00
15 6.6 1.00 1.07
12 8.27 1.04 1.12
10 9.95 1.07 1.15
8.5 11.72 1.09 1.18
7 14.25 1.11 1.21
5 20.0 1.14 1.26
IEEE Standard for Low-Voltage AC Power Circuit Breakers Used in Enclosures38
Medium-voltage Fuses
The definition of fuse in Low Voltage Fuses, page 82 is equally applicable to medium-
voltage fuses. Recall from Basic Principles, page 33 that the medium-voltage level is
defined by ANSI C84 as containing standard system voltages from 2400 through
69,000 V, and that the high voltage level contains standard system voltages from 115
kV through 230 kV. The medium-voltage level, strictly, is defined by ANSI C84 as
greater than 1000 V and less than 100,000 V. Similarly, the high-voltage level is
defined as greater than 100,000 V through 230,000 V. Strictly speaking, high-voltage
fuse standards are used for both medium- and high-voltage fuses. However, the focus
of this section is on medium voltage fuses 1 kV through 38 kV.
The following standards apply to medium-voltage fuses (see IEEE Recommended
Practice for Protection and Coordination of Industrial Power Systems39):
• IEEE Std. C37.41-2016
• IEEE Std. C37.42-2016
• IEEE Std. C37.48-2020
Those definitions in Low Voltage Fuses, page 82 which do not specifically reference
low-voltage fuses are also valid for medium-voltage fuses. Generally, medium-voltage
fuses can be divided into two major categories : Current-limiting and expulsion.
Current-limiting fuses were defined in Low Voltage Fuses, page 82I, and the same
basic definition applies to medium-voltage fuses. Expulsion fuses are defined as
follows (see IEEE Recommended Practice for Protection and Coordination of
Industrial Power Systems39):
Expulsion fuse: A vented fuse in which the expulsion effect of the gases produced by
internal arcing, either alone or aided by other mechanisms, results in current
interruption.
In addition, medium-voltage fuses are further classified as power fuses or distribution
fuses as follows (see IEEE Recommended Practice for Protection and Coordination of
Industrial Power Systems39):
Power fuse: Defined by ANSI C37.42-1996 as having dielectric withstand (BIL)
strengths at power levels, applied primarily in stations and substations, with
mechanical construction basically adapted to station and substation mountings.
Distribution fuse: Defined by ANSI C37.42-1996 as having dielectric withstand (BIL)
strengths at distribution levels, applied primarily on distribution feeders and circuits,
and with operating voltage limits corresponding to distribution voltages. These are
further subdivided into distribution current limiting fuses and distribution fuse cutouts,
as described below.
Current-limiting fuses interrupt in less than one half cycle when subjected to currents
in their current-limiting range. This is an advantage as it limits the peak fault current to
a value less than the prospective fault current as described above for low-voltage
fuses. This provides current-limiting fuses with high interrupting ratings and allows
them to shield downstream devices with lower short-circuit ratings in some cases.
However, the same technologies that combine to give medium-voltage current-liming
fuses their current-limiting characteristics can also produce thermal issues when the
fuses are loaded at lower current levels. For this reason, the following definitions
apply to current-limiting fuses (see IEEE Recommended Practice for Protection and
Coordination of Industrial Power Systems39).
Backup current-limiting fuse: A fuse capable of interrupting all currents from its
maximum rated interrupting current down to its rated minimum interrupting current.
General purpose current-limiting fuse: A fuse capable of interrupting all currents
from the rated interrupting current down to the current that causes melting of the
fusible element in no less than one hour.
Full-range current-limiting fuse: A fuse capable of interrupting all currents from its
rated interrupting current down to the minimum continuous current that causes
melting of the fusible elements.
Due to the limitations of backup and general-purpose current limiting fuses, current-
limiting power fuses have melting characteristics defined as E or R, defined as
follows:
E-Rating: The current-responsive element for ratings 100 A or below shall melt in 300
seconds at an RMS current within the range of 200% to 240% of the continuous-
current rating of the fuse unit, refill unit, or use link. The current-responsive element
for ratings above 100 A shall melt in 600 seconds at an RMS current within the range
of 220% to 264% of the continuous-current rating of the fuse unit, refill unit, or fuse
link.
R-Rating: The fuse shall melt in the range of 15 seconds to 35 seconds at a value of
current equal to 100 times the R number.
Similarly, distribution current-limiting fuses are defined by given characteristic ratings,
one of which is the C rating, defined as follows:
C-Rating: The current-responsive element shall melt at 100 seconds, at an RMS
current within the range of 170% to 240% of the continuous-current rating of the fuse
unit.
A typical time-current curve for an E-rated current-limiting power fuse is shown in
Figure 65. The fuse in Figure 65 is a 125E-rated fuse. The curve starts at
approximately 250 A for a minimum melting time of 1000 seconds. Take care with
backup and general-purpose current-limiting fuses so that the load current does not to
exceed the E- or R-rating of the fuse. Failure to do this can result in the development
of a hot-spot and subsequent loss of the fuse and its mounting. For fuses enclosed in
equipment, this can have disastrous consequences since loss of the fuse and/or its
mounting can lead to an arcing fault in the equipment. The boundary of the
characteristic, denoting the minimum-melting current, should be further derated to
consider pre-loading of the fuse (consult the fuse manufacturer for details). As with
low-voltage fuses, the current-limiting fuse characteristic does not extend below .01
seconds since the fuse is in its current-limiting range below this interrupting time.
A current-limiting power fuse consists of a fuse mounting (typically fuse clips) and the
fuse unit itself. These are frequently mounted in metal-enclosed switchgear. A
distribution current-limiting fuse may consist of a disconnecting-style holder or clips,
and the fuse unit. Distribution current-limiting fuses may also be provided with under-
oil mountings for use with distribution transformers. They are frequently used for
capacitor protection as well, with clips designed to mount to the capacitor.
Current-limiting power fuses are typically used for short circuit protection of instrument
transformers, power transformers, and capacitor banks. Maximum Ratings for
Current-limiting Power Fuses 2.75 – 38 kV, page 101 gives maximum ratings for
medium-voltage current-limiting power fuses from 2.75 through 38 kV.
Rated Maximum Voltage (kV) Continuous-current Ratings (A), Maximum Short-circuit Maximum Interrupting Ratings (kA
RMS symmetrical)
IEEE Recommended Practice for Protection and Coordination of Industrial Power Systems40
During interruption current-limiting fuses produce significant arc voltages. These must
be considered in selecting equipment. They are typically compared to the BIL level of
the equipment, including downstream equipment at the same voltage level. The
maximum permissible overvoltages for current-limiting power fuses are shown in
Maximum Permissible Overvoltages for Current-limiting Power Fuses, page 102 (see
IEEE Recommended Practice for Protection and Coordination of Industrial Power
Systems40).
In practice, the arc voltages for current-limiting fuses generally indicate the use of the
smallest available fuse voltage class for the given system voltage, for example, 5.5 kV
fuses instead of 8.3 kV fuses for a 4160 V system.
After a fault interruption, in a three-phase set of current-limiting fuses all three fuses
are replaced, even if only one fuse interrupted the fault. This is due to the possibility of
damage to the other two fuses due to the fault, which could change their time-current
characteristics and make them unsuitable to carry load current without loss.
Because medium-voltage current-limiting fuses interrupt short circuits without the
expulsion of gas or flame, they are widely utilized in a variety of applications.
Power expulsion fuses generally consist of an insulating mounting and a fuse holder
which accepts the fuse refills. The fuse holder may be of the disconnecting or non-
disconnecting type. Only the refill is replaced when a fuse interrupts an overcurrent,
and if only one phase of a three-phase set interrupted the fault, only that fuse needs
replacement. Power expulsion fuses are typically used in substations and enclosed
equipment.
Distribution expulsion fuses are generally distribution fuse cutouts, which are adapted
to pole or cross arm mounting. They consist of the fuse holder and refill unit. The fuse
holder is usually of the disconnecting type. These are typically used as pole-mounted
fuses on utility distribution systems.
Expulsion fuses use the liberation of de-ionizing gasses to interrupt overcurrents.
Boric acid is typically used as the interrupting medium for power expulsion fuses and
bone fiber is typically used for distribution fuse cutouts. When an expulsion fuse
interrupts an overcurrent the interrupting medium liberates de-ionizing gas,
interrupting the overcurrent. The exhaust gasses are then emitted from the fuse,
accompanied by noise. The exhaust gasses for a boric acid fuse may be condensed
by an exhaust control device (commonly called an exhaust filter, silencer, or snuffler).
Unlike current-limiting fuses, expulsion-type fuses interrupt high overcurrents at
natural current zeros. They are therefore non-current-limiting, and as a result typically
have lower interrupting ratings than current-limiting fuses. Maximum Continuous
Current and Short Circuit Interrupting Ratings for Refill Type Boric-acid Expulsion-type
Power Fuses, page 103 shows the maximum continuous current and short-circuit
interrupting ratings for refill-type boric-acid expulsion-type power fuses (see IEEE
Recommended Practice for Protection and Coordination of Industrial Power
Systems41). Because expulsion-type fuses are non-current-limiting, they do not
produce the significant arc voltages that current-limiting fuses produce, and thus it is
permissible to use a fuse with a larger voltage class than the system, for example, a
14.4 kV-rated fuse on a 4160 V system. This makes expulsion-type fuses particularly
useful on systems which may be upgraded in the future to a higher voltage. However,
the lower interrupting ratings of expulsion-type fuses are often an issue versus.
current-limiting fuses as the largest expulsion-type fuse interrupting ratings require
larger physical dimensions which cannot always be easily accommodated in enclosed
equipment. Further, in some cases the expulsion-type fuses prohibit some space-
saving mounting configurations in enclosed equipment that are available with current-
limiting fuses.
Table 16 - Maximum Continuous Current and Short Circuit Interrupting Ratings for Refill Type Boric-acid
Expulsion-type Power Fuses
Rated Maximum Voltage (kV) Continuous-current Ratings (A), Maximum Short-circuit Maximum Interrupting
Ratings (kA, RMS Symmetrical)
E-ratings are used for power expulsion fuses. A typical time-current characteristic for
a 125E boric-acid fuse is given in Typical Boric Acid Power Expulsion Fuse Time-
current Characteristic, page 104.
The characteristic extends to the available fault current (in this case, 29.4 kA), unlike
the current-limiting fuse. It is common practice to treat these as current-limiting fuses
so far as the E-rating is concerned, that is, the maximum load current is usually kept
below the E-rating. However, the boric-acid fuse is not subject to damage when
loaded above its E-rating, and they are often referred to in the industry as non-
damageable due to this fact.
When applying medium-voltage fuses, the voltage rating and the interrupting rating
are of importance. The maximum line-to-line voltage of the system should not exceed
the fuse voltage rating. The published interrupting rating for power fuses is typically for
a test X/R ratio of 15, and for distribution fuses the test X/R ratio is typically eight;
consult the fuse manufacturer for derating factors for X/R ratios above these values.
Also consult the manufacturer if the test X/R is in doubt.
Medium-voltage fuses provide economical short-circuit protection when applied within
their ratings, particularly for transformers, cables, and capacitors. For more
sophisticated protection at the medium-voltage level, employ other means.
Rated Dry Withstand Voltage: The RMS voltage that the circuit breaker in new
condition is capable of withstanding for one minute under specified conditions.
Rated Wet Withstand Voltage: The RMS voltage that an outdoor circuit breaker or
external components in new condition are capable of withstanding for ten seconds.
Rated Lightning Impulse Withstand Voltage: The peak value of a standard 1.2 x 50
µs wave, as defined in IEEE Std 4-2013, that a circuit breaker in new condition is
capable of withstanding.
Rated Continuous Current: The current in RMS symmetrical amperes that the circuit
breaker is designed to carry continuously.
Rated Interrupting Time: The maximum permissible interval between the energizing
of the trip circuit at rated control voltage and the interruption of the current in the main
circuit in all poles.
Rated Short Circuit Current (Required Symmetrical Interrupting Capability): The
value of the symmetrical component of the short-circuit current in RMS amperes at the
instant of arcing contact separation that the circuit breaker shall be required to
interrupt at a specified operating voltage, on the standard operating duty cycle, and
with a DC component of less than 20% of the current value of the symmetrical
component.
Required Asymmetrical Interrupting Capability: The value of the total RMS short-
circuit current at the instant of arcing contact separation that the circuit breaker shall
be required to interrupt at a specified operating voltage and on the standard operating
duty cycle. This is based upon a standard time constant of 45 milliseconds (X/R ratio
=17 for 60 Hz and 14 for 50 Hz systems) and an assumed relay operating time of one-
half cycle.
Rated closing and latching capability: The circuit breaker shall be capable of
closing and latching any power frequency making current whose maximum peak is
equal to or less than 2.6 (for 60 Hz power frequency; 2.5 for 50 Hz power frequency)
times the rated short-circuit current.
Rated Short-Time Current: The maximum short-circuit current that the circuit
breaker can carry without tripping for a specified period of time.
Maximum Permissible Tripping Delay: The maximum delay time for protective
relaying to trip the circuit breaker during short-circuit conditions, based upon the rated
short-time current and short-time current-carrying time period.
Rated Transient Recovery Voltage (TRV): At its rated maximum voltage, a circuit
breaker is capable of interrupting three-phase grounded and ungrounded terminal
faults at the rated short-circuit current in any circuit in which the TRV does not exceed
the rated TRV envelope. For a circuit breaker rated below 100 kV, the rated TRV is
represented by a one-cosine wave, with a magnitude and time-to-peak dependent
upon the rated maximum voltage of the circuit breaker.
Rated Voltage Range Factor K: Defined in earlier versions of IEEE Standard Rating
Structure for AC High-Voltage Circuit Breakers44 as the factor by which the rated
maximum voltage may be divided to determine the minimum voltage for which the
interrupting rating varies linearly with the interrupting rating at the rated maximum
voltage by the following formula:
Where:
• Ivmax is the rated short-circuit current at the maximum operating voltage.
• Vmax is the rated maximum operating voltage.
• Vop is the operating voltage where Vop V≥(Vmax/K()).
• Ivop is the short-circuit current interrupting capability where Ivop≤Ivmax.
For values of Vop below (Vmax ÷ K) the short-circuit interrupting capability was
considered to be equal to (Iv max x K). This model was more representative of older
technologies such as air-blast interruption. Because most modern circuit breakers
employ vacuum technology, the current version of IEEE Standard Rating Structure for
Rated Rated Rated Rated Short- Rated TRV Rated Rated Rated
Maximum Voltage Contin-uous Circuit and Interrupting Max. Closing
Voltage, Range Current (A Short-Time Rated Peak Rated Time Time (ms) Permissi- and
(kV) Factor K RMS) Current (kA Voltage E2, to Peak T2, ble Latching
RMS) (kV peak) (µs) Tripping Current,
Time (kA Peak)
Delay Y (s)
Although 83 milliseconds or five cycles is the “preferred” value per IEEE Standard
Rating Structure for AC High-Voltage Circuit Breakers47for the rated interrupting time,
three-cycle designs are common.
Other related preferred ratings, such as dielectric ratings and capacitance switching
ratings, are also given in AC High-Voltage Circuit Breakers Rated on a Symmetrical
Current Basis – Preferred Ratings and Related Required Capabilities46.
Table 18 - Preferred Ratings for Indoor Circuit Breakers with Voltage Range Factor k > 1.0
Rated Rated Voltage Rated Rated Short- Rated Rated Maximum Closing and
Maximum Range Factor K Continuous Circuit Current Interrupting Maximum Symmetri- Latching
Voltage, kV Current at 60 at Rated Time, Cycles Voltage cal Capability
Hz (A RMS) Maximum kV Divided by K, Interrupting 2.7 K Times
(kA RMS) kV RMS Capability Rated
and Rated Short-
Short-Time Circuit
Current (kA, Current (kA
RMS) Crest)
AC High-Voltage Circuit Breakers Rated on a Symmetrical Current Basis – Preferred Ratings and Related Required Capabilities48
closing coils to allow tripping and closing operations via protective relays, manual
control switches, PLC’s, SCADA systems. The circuit breaker internal control circuitry
is arranged per IEEE C37.11-1997. Circuit breakers are also equipped with a number
of auxiliary contacts to allow interlocking and external indication of circuit breaker
position.
For medium-voltage protection applications, circuit breakers offer flexibility that cannot
be obtained with fuses. Further, they do not require a separate switching device as
fuses do. These benefits are gained at a price: Circuit breaker applications are more
expensive than fuse applications, both due to the inherent cost of the circuit breakers
themselves and due to the protective relays required. For many applications,
however, circuit breakers are the only choice that offers the flexibility required. Large
medium-voltage services and distribution systems and most applications involving
medium-voltage generation employ circuit breakers.
Protective Relays
For medium-voltage circuit breaker applications, protective relays serve as the
“brains” that detect abnormal system conditions and direct the circuit breakers to
operate. They also serve to provide specialized protection in low-voltage power circuit
breaker applications for functions not available in the circuit breaker trip units.
Most modern protective relays are solid-state electronic or microprocessor-based
devices, although older electromechanical devices are still available. Solid-state
electronic or microprocessor-based relays offer more flexibility and functionality than
electromechanical relays, including the ability to interface with common
communications protocols such as MODBUS for integration into a SCADA
environment. However, they do require “reliable” control power to maintain operation
during abnormal system conditions. This control power is most often provided by a DC
battery system, although AC UPS-based systems are also encountered.
Electromechanical relays are typically single-phase devices. Solid-state electronic
relays are typically available in single-phase or three-phase versions. Microprocessor-
based relays are typically three-phase devices. While electromechanical and solid-
state electronic relays typically incorporate one relay function per device,
microprocessor-based relays usually encompass many functions in one device,
making a single microprocessor-based relay capable of performing the same
functions that would require several electromechanical or solid-state relays. This
functionality usually makes microprocessor-based relays a good choice for new
installations.
Protective relays are not rated for direct connection to the power system where they
are applied. For this reason, instrument transformers are used to reduce the currents
and voltages to the levels for which the relays are designed. Instrument transformers
generally fall into one of two broad categories: Current Transformers (CT’s) and
Voltage Transformers (VT’s). The loads on instrument transformers, such as relays
and meters, are known as burdens to distinguish them from power system loads.
A current transformer consists of a coil toroidally-wound around a ferromagnetic core.
The conductor for which the current is to be measured is passed through the center of
the toroid. The magnetic field generated by the current through the conductor causes
current to flow in the coil. In essence, a CT may be thought of as a conventional
transformer with one primary turn.
CT’s in the United States typically have 5 A-rated secondaries, with primary ratings
from 10 – 40,000 A and larger. For relaying applications in industrial facilities, CT
ratios are typically 50:5 – 4000:5. IEEE Std. C57.13-2016 designates certain ratios as
standard, as well as a classification system for relaying performance. The
classification system consists of a letter and a number. The letter may be C,
designating that the percent ratio correction may be calculated, or T, denoting that the
ratio correction has been determined by test. The number denotes the voltage that the
CT can deliver to a “standard burden” (as described in IEEE Std. C37.13-2016) at 20
times the rated secondary current without exceeding 10% ratio error. As a more
accurate alternative, manufacturer-published CT excitation curves may be used to
determine the accuracy. For a relaying application, the issue at hand is the
performance of the relay during worst-case short-circuit conditions, when the CT
secondary currents are the largest and may cause the secondary voltage to exceed
the CT’s rating due to the voltage developed across the relay input coil. This condition
causes the CT to saturate, significantly changing the ratio and thus the accuracy of
the measurement. For cases of severe CT saturation the relay may respond in an
unpredictable manner, such as not operating or producing “chatter” of its output
contacts.
CT's where the power conductor passes through the window formed by the toroidal
CT winding are known as window-type CT’s. CT’s designed with an integral bus bar
running through device are known as bus-bar type CT’s. Other designs, such as
wound primary CT’s for metering applications and non-saturating air-core CT’s, are
available. Additional information on CT application can be found in IEEE
Recommended Practice for Protection and Coordination of Industrial Power
Systems51.
Quasi-Physical Circuit
Arrangement Representation
Voltage transformers (VT’s) are used to step the power system voltage down to a level
that the relay can utilize. The operation of voltage transformers is essentially the same
as for conventional power transformers except that the design has been optimized for
accuracy. Like current transformers, voltage transformers are assigned accuracy
classes by IEEE Std. C57.13-2016. VT accuracy classes are designated W, X,M Y, Z,
and ZZ in order of increasing burden requirements. Refer to IEEE Recommended
Practice for Protection and Coordination of Industrial Power Systems51for more
information regarding the application of voltage transformers.
Protective relays are classified by function. To make circuit representations easier,
each function has been defined and assigned a number by IEEE Std. C37.2-2022.
The IEEE standard function numbers are given in Commonly Used Protective Relay
Device Function Numbers, page 112. Commonly Used Suffix Letters Applied to Relay
Function Numbers, page 112 gives the commonly-used suffix letters to further
designate protective functions IEEE Recommended Practice for Protection and
Coordination of Industrial Power Systems51.
These designations can be combined in various ways. For example, 87T denotes a
transformer differential relay, 51N denotes a residual ground time-overcurrent relay,
87B denotes a bus differential relay.
25 Synchronizing
27 Undervoltage
32 Directional Power
50 Instantaneous Overcurrent
51 Time-overcurrent
59 Overvoltage
67 Directional Overcurrent
86 Lockout
87 Differential
IEEE Recommended Practice for Protection and Coordination of Industrial Power Systems52
A Alarm only
B Bus protection
G Ground fault protection [relay current transformer (CT) in a system neutral circuit]
or generator protection]
L Line Protection
M Motor Protection
N Ground fault protection (relay coil connected in residual CT circuit)
T Transformer protection
V Voltage
IEEE Recommended Practice for Protection and Coordination of Industrial Power Systems52
relays have no inherent time delay and are used for fast short-circuit protection. 50
and 51 Overcurrent Relay Characteristics, page 113 shows the timing characteristics
of several typical 51 time-overcurrent relay curve types, along with the 50
instantaneous characteristic.
The pickup level is set by the tap setting, which may be primary, secondary, or per unit
current values. Each relay curve has a time dial/delay setting that allows the curve to
be shifted up or down on the time-current characteristic curve. In 50 and 51
Overcurrent Relay Characteristics, page 113, the time dial settings are different to
give enough space between the curves to show their differences.
The above are IEEE-standard curves; others are available, depending upon the relay
make and model. A solid-state electronic or microprocessor-based relay has all of
these curves available on one unit; electromechanical relays must be ordered with a
given characteristic that cannot be changed.
The 50 instantaneous function is only provided with a pickup setting. The 30
milliseconds delay shown in 50 and 51 Overcurrent Relay Characteristics, page 113
for the 50 function is typical and considers both the relay logic operation and the
output contact closing time. Most microprocessor-based units also have an adjustable
delay for the 50 function; when an intentional time delay is added the 50 is referred to
as a definite-time overcurrent function. On solid-state electronic and microprocessor-
based relays, the 50 function may be enabled or disabled. On electromechanical
Source
Load
Transformer-Neutral
For ungrounded systems, little ground current flows during a single phase-to-ground
fault. Low-voltage solidly-grounded systems are discussed below.
The typical application of phase and residual neutral ground overcurrent relays in one-
line diagram form is shown in Typical Application of Overcurrent Relays, page 115.
In Typical Application of Overcurrent Relays, page 115, the designation 52 is the IEEE
Std. C37.2-2022 designation for a circuit breaker. The phase relays are designated 51
and the residual ground overcurrent relay is designated 51N (both without
instantaneous function). IEEE Recommended Practice for Protection and
Coordination of Industrial Power Systems53 denotes that there are three phase
overcurrent relays and three CTs. The dotted line from the relays to the circuit breaker
denotes that the relays are wired to trip the circuit breaker on an overcurrent condition.
Another type of overcurrent relay is the voltage-restrained overcurrent relay 51V and
the voltage-controlled relay 51C. Both are used in generator applications to allow the
relay to be set below the generator full-load current due to the fact that the fault
contribution from a generator decays to a value less than the full-load current of the
generator. The 51C relay does not operate on overcurrent unless the voltage is below
a preset value. The 51V relay pickup current shifts as the voltage changes, allowing it
to only respond to overcurrents at reduced voltage. Both require voltage inputs, and
thus require voltage transformers for operation.
Directional Overcurrent Relays (Devices 67, 67N)
When fault currents can flow in more than one direction with respect to the load
current it is often desirable to determine which direction the fault current is flowing and
trip the appropriate devices accordingly. This is usually due to the need to de-energize
only those parts of the power system that must be de-energized to contain a given
fault.
Standard overcurrent relays cannot distinguish the direction of the current flow.
Directional relays (67, 67N) are required to perform this function.
An important concept in the application of directional overcurrent relays is
polarization. Polarization is the method used by the relay to determine the direction of
current flow. For phase directional overcurrent relays, this is accomplished by the use
of voltage transformers, which provide a voltage signal to the relay and allow it to
distinguish the current direction. The details of polarization methods are not discussed
here, but can be found in IEEE Recommended Practice for Protection and
Coordination of Industrial Power Systems53. Because the voltage on a faulted phase
can be unreliable, each phase is restrained via the voltage from a different phase.
Care must be used when defining CT polarities as each manufacturer typically defines
a preferred polarity to match their standard connection diagrams.
Polarization for a 67N relay is more difficult. They must be polarized with zero-
sequence current or zero-sequence voltage. Electromechanical 67N relays must be
polarized via either a CT in the source transformer neutral (zero-sequence current
polarization) or three VT’s connected with a wye-connected primaries and broken-
delta connected secondaries (refer to A Ground Detection Method for Ungrounded
Systems, page 59 for an example of the wye-broken delta connection with
ferroresonance-swamping resistor). Solid-state 67N relays usually must be polarized
the same way but do sometimes offer a choice of either method. Microprocessor-
based relays typically offer a choice of either method and, in some cases, can self-
polarize by calculating the zero-sequence voltage from the measured three-phase line
voltage.
As an example of the effectiveness of directional overcurrent relays, consider the
primary-selective system arrangement from Radial System, page 37. The primary
main and tie circuit breakers and an example of protective relaying for those circuit
breakers are shown in Example Protective Relaying Arrangement for Closed-
Transition Primary-selective System, page 117.
Utility Utility
Fault
Feed Feed Current
#1 #2 Flow
VT (3) VT (3)
Fault Fault
Current Fault
Flow Current
Flow
Aux Aux
(3) VT. VT. (3)
(3) (3)
CT 51 51 CT
600:5 600:5 Fault
(3) (3) Current
51N 51N Flow
52-M1 52-M2
CT 67N 67N CT
600:5 600:5
(3) 52-T (3)
67 N.C. 67
(3) (3)
(3)
51
51N
conjunction with 27 relays. Either application gives a voltage “window” within which
the system is allowed to operate. In this application 59 relays should be time-delayed
just as 27 relays are.
59 relays may also be used for ground-fault detection on high-resistance grounded or
ungrounded systems. Application for a high-resistance grounded system is shown in
Pulsing Ground Detection System, page 62. For an ungrounded system the 59 relay
may be used across the broken-delta secondary of a ground-detection VT circuit, such
as the circuit shown in A Ground Detection Method for Ungrounded Systems, page
59.
When an electromechanical or solid-state electronic relay includes both 27 and 59
functions it is referred to as a 27/59 relay. When an electromechanical or solid-state
electronic relay includes 27, 47, and 59 functions it is referred to as a 27/47/59 relay.
F. Lockout Relays (Device 86)
The lockout relay is used to trip a device and prevent its reclosure until the lockout
relay is reset. In most cases the lockout relay is essentially a switch, and in fact is
typically mounted in close proximity to circuit breaker control switches. The relay is
spring-loaded, and a trip coil, when energized, causes the lockout relay to trip the
connected devices and prevent them from reclosing. There is typically a conspicuous
target on the lockout relay to alert operating personnel that it has tripped. When the
lockout relay is reset, the opening springs are compressed and the relay is ready for
the next tripping operation.
86 relays are commonly used where one protective relay must trip several protective
devices, and where reclosure of the tripped devices needs to be controlled to avoid
closing onto a fault.
G. Differential Relays (Device 87)
Differential relays operate on the principle that if the current flowing into a device does
not equal the current flowing out, a fault must exist within the device.
Differential relays generally fall within one of two broad categories: Current-differential
or high-impedance differential.
Current-differential relays are typically used to shield large transformers, generators,
and motors. For these devices detection of low-level winding-to-ground faults is
essential to avoid equipment damage. Current differential relays typically are
equipped with restraint windings to which the CT inputs are to be connected. For
electromechanical 87 current differential relays, the current through the restraint
windings for each phase is summed and the sum is directed through an operating
winding. The current through the operating winding must be above a certain
percentage (typically 15%-50%) of the current through the restraint windings for the
relay to operate. For solid-state electronic or microprocessor-based 87 relays the
operating windings exist in logic only rather than as physical windings.
A typical application of current-differential relays for protection of a transformer is
shown in Typical Application of Current-differential Relays for Delta-wye Transformer
Protection, page 120. In Typical Application of Current-differential Relays for Delta-
wye Transformer Protection, page 120, the restraint windings are labeled as “R” and
the operating windings are labeled as “O”. Because the delta-wye transformer
connection produces a phase shift, the secondary CT’s are connected in delta to
counteract this phase shift for the connections to the relays. Under normal conditions
the operating windings carry no current. For a large external fault on the load side of
the transformer, differences in CT performance in the primary vs. the secondary (it is
impossible to match the primary and secondary CT’s due to different current levels)
are taken into account by the proper percentage differential setting. Because the CT
ratios in the primary versus secondary is not always able to match the current
magnitudes in the relay operating windings during normal conditions, the relays are
equipped with taps to internally adjust the current levels for comparison. The specific
connections in this example apply to a delta primary/wye secondary transformer or
transformer bank only. The connections for other winding arrangement vary, to
properly cancel the phase shift. For many solid-state electronic and microprocessor-
based relays, the phase shift is made internally in the relay and the CT’s may be
connected the same on the primary and secondary sides of the transformer
regardless of the transformer winding connections. The manufacturer’s literature for a
given relay make and model must be consulted when planning the CT connections.
Source
Source
Relay Current Inputs
A B C
A-Phase
R R
B-Phase
R R
C-Phase
R R
A B C
Load
Load
High-impedance differential relays are typically used for bus protection. Bus protection
is an application that demands many sets of CT’s be connected to the relays. It is also
an application that demands that that relay be able to operate with unequal CT
performance, since external fault magnitudes can be quite large. The high-impedance
differential relay meets both requirements.
High-impedance Differential Relay Concept, page 122 shows the application of bus
differential relays to a primary-selective system. In High-impedance Differential Relay
Concept, page 122 the zones of protection for Bus #1 and Bus #2 overlap. Here the
86 relay is extremely useful due to the large number of circuit breakers to be tripped.
All circuit breakers attached to the protected busses are equipped with differential
CT’s and are tripped by that busses’ respective 86 relay. The 87 relays are denoted
87B since they are protecting busses. The same applies for the 86B relays. The
protective zones overlap; this is typical practice so that that all parts of the bus work
are protected.
The high-impedance differential relay is typically set in terms of voltage across the
input. The voltage setting is typically set so that if one CT is fully saturated and the
others are not the relay do not operate. By its nature, the high-impedance differential
relay is less sensitive than the current-differential relay, but since it is typically applied
to shield bussing, where fault magnitudes are typically high, the additional sensitivity
is not required.
Figure 80 - Low-voltage Ground Fault Protection for Four-wire Radial System with Electronic-trip Circuit
Breaker
Source
A B C N
Main or System
Electronic Trip Bonding
Circuit Breaker Jumper
Trip Unit
with GFT
CB Current (LSG,
Sensors LSIG)
Neutral Sensor
or CT
Safety Ground
may or may not
exist on Neutral
Sensor CT Circuit
Ground
Fault
Load
In Low-voltage Ground Fault Protection for Four-wire Radial System with Electronic-
trip Circuit Breaker, page 124 the neutral sensor may be an air-core CT or a
conventional iron-core CT. The ground fault current is diverted around the neutral
sensor when it is placed on the load-side of the main or system bonding jumper (see
System Grounding, page 50 for the definition of main and system bonding jumpers
and related discussion). Under normal unbalanced-load conditions the neutral sensor
detects the neutral current and stops the circuit breaker from tripping. If the system is
a three-wire system without a system neutral the neutral CT is omitted.
If the circuit breaker is not equipped with an electronic trip system, an external ground
fault relay may be used with a zero-sequence sensor to trip the circuit breaker. The
circuit breaker must be equipped with a shunt trip attachment in this case. Low-
voltage Ground Fault Protection for Four-wire Radial System Without Electronic Trip
Circuit Breaker, page 125 shows an example of this arrangement. In Low-voltage
Ground Fault Protection for Four-wire Radial System Without Electronic Trip Circuit
Breaker, page 125 the external ground fault relay is noted as “GS”. In low-voltage
systems this is the typical notation rather than “51G”, although “51G” could also be
used. In a three-wire system the neutral is omitted, and the zero-sequence sensor
includes the phase conductors only.
Figure 81 - Low-voltage Ground Fault Protection for Four-wire Radial System Without Electronic Trip Circuit
Breaker
Source
A B C N
Main or
System
Bonding
Jumper
Circuit Breaker
without ST
Electronic Trip
Zero-Sequence GS
Sensor or CT
Safety Ground
may or may not
exist on
Ground Zero-Sequence
Fault Sensor/CT Circuit
Load
These methods provide sensitive ground fault protection for solidly-grounded radial
systems. However, if multiple sources are involved a more involved system is required
to obtain acceptable ground-fault protection.
Modified-differential Ground Fault Systems
Because four-pole circuit breakers are not in common use in the United States, the
issue of multiple ground current return paths has a large effect upon ground-fault
protection in four-wire systems. To illustrate this point, consider a secondary-selective
system as shown in Secondary-selective System with Radial Ground-fault Protection
of Figure 80 Applied, page 126.
A ground fault on one bus has two return paths: Through its source-transformer main/
system bonding jumper or the other source-transformer main/system bonding jumper
neutral. How much ground fault current flows in each path is dependent upon the
ground or zero-sequence impedances of the system, which is difficult to evaluate.
Therefore, assume a factor of A x the total ground-fault current flows through the
source transformer main/system bonding jumper neutral and B x the total ground-fault
current flows through the other transformer main/system bonding jumper, where A + B
= 1. As can be seen from Secondary-selective System with Radial Ground-fault
Protection of Figure 80 Applied, page 126, the ground-fault protection for the faulted
bus can be de-sensitize or, worse, the wrong circuit breaker(s) may trip.
The solution is the modified-differential ground fault system. A typical example of such
a system is shown in Secondary-selective System with Radial Ground-fault Protection
of Figure 80 Applied, page 126.
This characteristic is adjustable both for pickup and time delay. Discrete relays for use
with non-electronic circuit breakers are also available with similar characteristics.
Take care when coordinating ground-fault protection if multiple levels of ground-fault
protection do not exist downstream from the service or source of a separately-derived
system. The NEC Article 230.95 (A) service-entrance requirement The National
Electrical Code56for a maximum of 1200 A pickup and maximum one second delay at
3000 A ground-fault current can lead to a lack of coordination for downstream feeder
and branch-circuit ground faults. This is one of the reasons for the use of other than
solidly-grounded systems where maximum system reliability is to be achieved.
Surge Protection
Surge protection is protection of conductors and equipment against the effects of
voltage surges. These are usually due to lightning, although switching transients can
also cause damaging overvoltages. Unlike overvoltage relaying, surge protection is
directly connected to the power circuit, and for the maximum protection is usually
located as close as physically practical to the protected equipment.
56. NFPA 70, The National Fire Protection Association, Inc., 2020 Edition.
Duty-cycle Voltage (kV) MCOV (kV) Four-wire Effectively-grounded Three-wire Grounded and
Neutral System a Resistance-Grounded Wye Systemsb
4800
9 7.7 12000Y/ 6930 6900
12470Y /7200
10 8.4 13200Y/ 7620
13800Y/ 7970
12 10.2
15 12.7 20780Y/12000 12000
12470
18 15.3 22860Y/ 13200 13200
b Includes grounded-wye systems where the path to the upstream transformer neutral includes an earth path
Figure 86 - Example Protection for a 1000 kVA, 13.2 kV Delta:480Y/27 7V, 5.75%Z
Dry-type Transformer
Figure 87 - Fault-current Flow for Delta-Wye Transformer L-N Faults and Delta-Delta Transformer L-L Faults
0.577 1.0
X1
H1
1.0 L-N
1.0
0.577 0 X3
0
0
0.577 X2
0
H3 H2
L-N Fault - Current values are per unit of the 3-ph fault current
0.866
0.866
X1
H1
L-L
0.866
0.866 0.289
0.289 0.577
0.577
0.289
0.289
0
X3 X2
H3 H2
L-L Fault - Current values are per unit of the 3-ph fault current
Generator Protection
The subject of generator protection is a complex one, and due to this fact it is not
presented here. Refer to IEEE Recommended Practice for Protection and
Coordination of Industrial Power Systems58 for detailed descriptions of generator
protection methods, as well as descriptions of protective relay types that are not
discussed above that used for generator protection.
Other Devices and Additional Information
For protection other devices, refer to IEEE Recommended Practice for Protection and
Coordination of Industrial Power Systems58 and/or the applicable standards for the
device in question. For additional information on the protection of cables and
transformers, refer to IEEE Recommended Practice for Protection and Coordination of
Industrial Power Systems58.
Protection Selectivity
The selectivity of protection refers to its ability to isolate an abnormal condition to the
smallest portion of the system possible. In most cases selectivity is a function of how
well-coordinated the overcurrent protective devices in the system are. As an example,
consider the system of Example System for Selectivity Discussion, page 135.
Example System for Selectivity Discussion, page 135 shows a small radial system
with a medium-voltage utility service, a service substation consisting of a primary
switch step-down transformer protected by a primary fuse, and a secondary
switchboard. One of the switchboard feeder circuit breakers is shown feeding al
lighting panel and other loads.
For optimum selectivity, a fault at point G should only cause its lighting panel feeder
circuit breaker to trip. The panel main circuit breaker and all devices upstream should
not be affected. If the lighting panel feeder circuit breaker time-current characteristic
does not coordinate with that of the lighting panel main, the main may trip, de-
energizing the entire panelboard.
Going upstream, a fault at point F should only cause the panelboard main circuit
breaker to trip and a fault at point E should only cause the switchboard main circuit
breaker to trip. A fault at point D may cause the switchboard main circuit breaker to
trip or the primary fuse to blow, but the effect on the system is the same since all of the
loads will be de-energized in either event. A fault at point C should only cause the
transformer primary fuse to blow.
Lack of selectivity causes more of the system to be de-energized for a fault in a given
location. The severity of the outage increases as the fault location is considered
farther and farther upstream. In this example, if the transformer primary fuses and the
upstream utility recloser, protective relays, or fuses are not coordinated the entire
utility distribution line, or a segment of the line, could be de-energized, affecting other
customers.
To analyze system selectivity, a time-current coordination study must be performed.
This study analyzes the time-current coordination characteristics of the protective
devices in the system and plots them on time current curves such as those illustrated
in this section. Coordination is achieved between two devices if their time-current
bands show sufficient clear space between them on the time-current curve or, in the
Utility Considerations
Abstract: Nearly all utility companies have direct control over the requirements for
connections to their system. Many requirements align with published codes and
standards. However, take care to ensure each utility company’s exact standards are
met. In addition, special requirements for microgrids and other forms of distributed
generation may be involved.
Introduction
Most industrial and commercial facilities are served from public utilities. However, the
utility interface is often the most neglected aspect of system design. This is especially
true at the medium-voltage level. Often, the service equipment manufacturer is
expected to resolve issues that severely impact the design of the system. This can
result in unexpected costs and project delays. Address these issues during the
system design stage, where the impacts to system reliability and cost can be
adequately managed; only by knowing the utility’s requirements is this possible.
service requirements among its member utilities, to publish existing utility service
requirements for electric service equipment, and to provide direction for development
of future metering technology. EUSERC publishes a manual EUSERC Manual59which
delineates requirements for electric service equipment through 34.5 kV. At the time of
publication, 80 utilities from 12 states are involved with EUSERC. While EUSERC
does not eliminate the need for individual utility requirements, it does help a great deal
in making electrical service equipment more standardized and less costly.
Introduction
The term power quality may take on any one of several definitions. The strict definition
of power quality is “the concept of powering and grounding electronic equipment in a
manner that is suitable to the operation of that equipment and compatible with the
premises wiring system and other connected equipment” (see IEEE Recommended
Practice for Powering and Grounding Electronic Equipment60). In practice, however,
the term power quality is often used to denote the proximity of the system voltage to
its sinusoidal form at the nominal voltage level. Deviation from this sinusoidal norm
therefore denotes a power quality issue. Strictly speaking, this deviation is a power
disturbance, defined as “any deviation from the nominal value (or from some selected
thresholds based upon tolerance) of the AC input power characteristics” IEEE
Recommended Practice for Powering and Grounding Electronic Equipment60. The
most common power disturbances are, as defined by IEEE Recommended Practice
for Powering and Grounding Electronic Equipment60:
Overvoltage: An RMS increase in the AC voltage, at the power frequency, for a
period of time greater than one minute. Typical values are 110% - 120% of nominal.
Undervoltage: An RMS decrease in the AC voltage, at the power frequency, for a
period of time greater than one minute. Typical values are 80 – 90% of nominal.
Swell: An increase in RMS voltage or current at the power frequency for durations
from .5 cycle – one minute. Typical values are 110% - 180% of nominal.
Sag: An RMS reduction in the AC voltage, at the power frequency, for durations from
½ cycle to a few seconds.
Interruption: The complete loss of voltage. A momentary Interruption is a voltage
loss (<10% of nominal) for a time period between .5 cycles and three seconds). A
temporary interruption is a voltage loss (<10% of nominal) for a time period between
three seconds and one minute. A sustained interruption is the complete loss of
voltage for a time period greater than one minute.
Notch: A switching (or other) disturbance of the normal power system voltage
waveform, lasting less than ½ cycle; which is initially of opposite polarity to the
waveform, and is thus subtractive from the normal waveform in terms of the peak
value of the disturbance voltage. This includes a complete loss of voltage for up to ½
cycle.
Transient: A subcycle disturbance in the AC waveform that is evidenced by a sharp
discontinuity of the waveform. It may be of either polarity and may be additive to, or
subtractive from, the nominal waveform.
Flicker: A variation in input voltage, either magnitude or frequency, sufficient in
duration to allow visual observation of a change in electric light source intensity.
Harmonic Distortion: The mathematical representation of distortion of the pure sine
waveform. This refers to the distortion of the voltage and/or current waveform, due to
the flow of non-sinusoidal currents.
Electrical Noise: Unwanted electrical signals that produce undesirable effects in the
circuits of the control systems in which they occur. Noise may be further categorized
as transverse-mode noise, which is measurable between phase conductors but not
phase-to-ground, and common-mode noise, which is measurable phase-to-ground
but not between phase conductors. This noise may be conducted or radiated. Also
referred to as RFI (radio-frequency interference) or EMI (electro-magnetic
interference).
The causes of the common power disturbances listed can vary greatly. Common
causes are listed in Common Power Disturbance Causes, page 142.
Notches and harmonic Power electronic converter equipment such as rectifiers, inverters,
distortion drives, which produce non-sinusoidal load current and commutation
notches
Interruptions Faults causing overcurrent protective device operation
Power disturbances can greatly affect utilization equipment. For example, sensitive
electronic medical equipment can malfunction, adjustable speed motor drives may trip
off-line. Interruptions can cause microprocessor-based equipment such as computers
to lose data. In extreme conditions, such as for voltage surges caused by direct
lightning strikes, both power equipment and utilization equipment may be subject to
stop functioning. With the high reliability requirements imposed upon power systems,
it is imperative that power system disturbances, or potential disturbances, be
mitigated to avoid down-time, equipment loss, and risk to human life.
(13–1)
Where:
• Vh is the RMS harmonic voltage (or current) value at a frequency of n times the
fundamental frequency.
• V1 is the RMS fundamental-frequency voltage or current.
Alternate forms for the distortion factor are given in IEEE Recommended Practices
and Requirements for Harmonic Control in Electrical Power Systems62 as
percentages of the nominal voltage or demand load current for the system under
consideration, for use in evaluation of the harmonic content of the system voltage or
current. These are referred to as Total Harmonic Distortion (THDVn) and Total Demand
Distortion (TDD), defined as follows:
Where:
• Vh is the RMS value of the nth harmonic component of the voltage.
• Vn is the RMS nominal fundamental voltage value.
• Ih is the RMS value of the nth.
• IL is the maximum demand load current, typically the average maximum monthly
demand over a 12-month period.
Crest Factor: The ratio of the peak value of a periodic function to the RMS value, that
is:
Where:
• Ypeak is the peak value of a periodic function.
• Yrms is the RMS value of the function.
61. IEEE Recommended Practice for Powering and Grounding Electronic Equipment, IEEE Std. 1100-2005, December 2005.
62. IEEE Std. 519-2014, June 2014.
Because power system voltages and currents are nominally sinusoidal, the nominal
crest factor for these would be √2 , which is 1.414 (see Electric Power Fundamentals,
page 10 for details).
Notch Area: A notch in the power system voltage (or current) is illustrated in Common
Power Disturbance Causes, page 142 IEEE Recommended Practices and
Requirements for Harmonic Control in Electrical Power Systems63:
The notch area for the notch as illustrated in Common Power Disturbance Causes,
page 142 is defined as:
an = t x d
(13–5)
Where:
• An is the notch area in volt-microseconds.
• t is the notch time duration in microseconds.
• d is the notch depth in Volts.
Recovery Time: This is the time needed for the output voltage or current to return to a
value within the regulation specification after a step load or line change.
Displacement Power Factor: The ratio of the active power of the fundamental wave,
in Watts, to the apparent power of the fundamental wave, in VA. This is the traditional
definition of power factor.
Total Power Factor: The ratio of the total input power, in watts, to the total VA input.
This includes the effects of harmonics.
K Factor: A measure of a transformer’s ability to serve non-sinusoidal loads. The K
factor is defined as:
(13–6)
Where:
• Ih is the harmonic component at h times the fundamental frequency.
• h is the harmonic order of Ih in multiples of the fundamental frequency.
• hmax is maximum harmonic order present.
Voltage Surges
The causes of voltage surges may be split into two major categories: Power system
switching and environmental (see IEEE Recommended Practice for Powering and
Grounding Electronic Equipment64). Both exhibit decaying oscillatory transients.
Capacitor switching close to the point under consideration is the most common cause
of switching surges, while lightning is the most common cause of environmentally-
induced voltage surges. Both can cause severe damage to unprotected power system
components, with the potential for lightning damage being the most severe; in the
worst case, lightning damage can be catastrophic.
Surge arrestors, as described in System Protection, page 69, are typically used to
minimize voltage surges. On low-voltage systems surge protective devices (SPD),
also described in System Protection, page 69 are also used. For motors, surge
capacitors are an option. In severe cases, custom-designed R-C snubber circuits may
be required as well.
Harmonic Distortion
Harmonic distortion is a subject of great interest in modern power systems. Harmonic
distortion results from non-sinusoidal load currents. These currents are the result of
non-linear loads, such as drives, which employ power electronic devices to rectify the
AC waveform. These devices draw non-sinusoidal currents which, in turn, cause non-
linear voltages to be developed in the system.
IEEE Standard 519-1992 IEEE Recommended Practices and Requirements for
Harmonic Control in Electrical Power Systems65 gives recommended limits for current
distortion due to consumer loads and voltage distortion in the utility supply voltage.
Both are referenced at the point on the utility system where multiple customers can be
served, referred to as the Point of Common Coupling (PCC). The requirements from
IEEE Recommended Practices and Requirements for Harmonic Control in Electrical
Power Systems66 for current distortion limits on general distribution systems 120 V –
69 kV are given in Harmonic Distortion, page 145. IEEE 519-1992 Harmonic Voltage
Distortion Limits, page 146 shows the corresponding utility voltage distortion limits.
The current limits are given both as limits on the individual harmonic levels and a limit
on the TDD, and that as the ratio Isc/IL increases the limits also increase. The reason
for this is that the current distortion limits are designed to limit the voltage distortion at
the PCC, and the voltage distortion for a given current distortion worsens with a larger
source impedance (V̄ = Ī • Z̄)
Table 23 - IEEE 519-1992 Harmonic Current Distortion Limits for General Distribution Systems 120 V Through
69 kV
Current distortions that result in a DC offset, half-wave converters, are not allowed.
*All power generation equipment is limited to these values of current distortion, regardless of actual L SC /IL.
Where:
IEEE Recommended Practices and Requirements for Harmonic Control in Electrical Power Systems66
NOTE: NOTE – High-voltage systems can have up to 2.0% THD where the cause is an HVDC
terminal that attenuates by the time it is tapped for a user.
IEEE Recommended Practices and Requirements for Harmonic Control in Electrical Power Systems,
IEEE Std67
for the circuit in question. Passive filters are also used for power factor correction.
However, there is a limit to their effectiveness and if higher-order harmonics must be
attenuated their use is generally not cost-effective. Take care in all cases to balance
the harmonic and power factor correction considerations.
Phase multiplication operates on the principle that if m six-pulse rectifiers are shifted
60/m degrees from each other, are controlled by the same delay angle, and are
loaded equally, the only harmonics present are:
h = kq ± 1 (13–7)
Where:
• h is a harmonic order present.
• q = 6m and is known as the pulse number of the circuit.
• k is any integer.
Thus, for standard six-pulse rectifiers the harmonic orders present are 5, 7, 11, 13,….
18-pulse rectifiers are the current state-of-the-art; for an 18-pulse rectifier (m = 3), the
harmonic orders present are 17, 19, 35, 37, …,. For an 18-pulse converter, the lower-
order harmonics are thus eliminated. For systems with large numbers of phase-
multiplied converters the harmonic current limits in NFPA Emergency Power System
Levels, page 186 are increased by the factor (q/6)1/2, where q is the pulse-number of
the predominate non-linear load on the system. In this case the limits for the harmonic
orders that do not fit equation (11-7) for the q of the predominate non-linear load are
multiplied by a factor of 0.25. Phase-shifting transformer connections are used to
achieve the 60/m degree phase shift between six-pulse rectifier units.
Active Harmonic Filters (AHF) can help greatly reduce harmonics and bring systems
in compliance with IEEE-519 or other harmonic limits imposed by the utility
companies. Current state-of-the-art designs measure the current, filter out the
fundamental frequency of the measured current, and inject current that is the negative
of the result into the system to cancel the harmonics up to a given harmonic order.
These systems are generally used in existing installations that have existing six-pulse
drives where replacing the drives is not a cost-effective solution, or where multiple
smaller six-pulse drives are utilized since phase multiplication for a drive below 100
HP is generally not cost-effective. State-of-the-art units can also dynamically correct
the power factor as well as phase imbalance and are advantageous vs. passive filters
both in their effectiveness and their flexibility in power factor correction.
Utility
Meter #1
A D
Meter #2
Power Consumer
In most cases, a power quality centric EPMS (Energy & Power Monitoring Software) is
required to analyze the waveforms, timestamps, sequence of events, etc. to realize a
cohesive view of the power quality issues in a facility. Refer to Electrical Energy
Management, page 220 for more details.
The inclusion of power monitoring equipment in the initial power system design makes
diagnosis of any subsequent power quality issues, if they arise, much easier and more
efficient. Reference IEEE Recommended Practice for Monitoring Electric Power
Quality68 contains much information on power quality monitoring and should be
consulted for further reference.
Safety Considerations
Arc Flash Hazard
Abstract: The precedence for eliminating arc flash hazards has evolved into a major
consideration for both the design and implementation of power distribution systems
and the operation and maintenance of the gear. Awareness of high incident energy
and overall high risks and hazard present in customer facilities has changed the
landscape for the better. To reduce hazards and risks, equipment design and layout,
and operational practices have improved to keep operators and bystanders safer.
Introduction
The method of which to analyze arc flash hazards has evolved through several
iterations of NFPA 70E, NPFA 70, and IEEE 1584 standards which modified the
calculation methods to provide more accurate data based on testing, as well as UL
2986. Analysis of hazards has also evolved to consider both the actual arc flash
hazard and the impeding risk or chance of an occurrence. Manufacturer design and
engineered controls have also taken into consideration these changes. Energy
reducing methods are also now required for certain applications.
Background
Electrical arcs form when a medium that is normally an insulator, such as air, is
subjected to an electric field strong enough to cause it to become ionized. This
ionization causes the medium to become a conductor which can carry current. The
phenomenon of electrical arcing is as old as the world itself. Lightning is a natural form
of electrical arcing. Man-made electrical arcs exist in devices such as arc furnaces.
However, utilization of electrical energy invariably requires equipment where
unintentional arcing between conductors becomes a possibility.
Electric arcs in equipment liberate large amounts of uncontrolled energy in the form of
intense heat and light. Unintentional arcing in power equipment can impose several
different types of hazards:
• Heat from arc can cause severe flash burns many feet away (temperatures can
reach 20,000 K, four times the temperature at the surface of the sun.).
• Byproducts from the arc, such as molten metal spatter, can cause severe injury.
• Pressure wave effects caused by the rapid expansion of air and vaporization of
metal can distort enclosures and cause doors and cover panels to be ejected with
severe force, injuring personnel.
• Sound levels can damage hearing.
Example of Arcing Damage to Equipment, page 151 gives an indication of the amount
of uncontrolled energy an arc can contain, as seen by the amount of damage to the
equipment shown.
Electrical safety has traditionally been concerned only with electric shock hazards.
The recognition of arc flash hazards began formally in 1981 with a paper “The Other
Electrical Hazard: Arc Blast Burns” 69by Ralph Lee, presented at the 1981 IEEE IAS
Annual Meeting. This paper established theoretical modeling for the heat energy
incident upon a surface a given distance from the arc. Subsequent developments
followed over the next 20 years, including testing to develop more accurate empirical
calculation methods and to evaluate protective clothing.
69. Lee, R., “The Other Electrical Hazard: Electrical Arc Blast Burns,”, vol. 1A-18, no. 3, May/June 1982.
At the time of publication, there are two basic standards which establish requirements
for arc flash hazards. The first is NFPA 70E, Standard for Electrical Safety in the
Workplace70, which defines the basic practices to be followed for electrical safety,
including protective clothing levels which must be worn for given levels of arc flash
incident energy and what steps must be taken prior to live work on electrical
equipment. The second is the IEEE Guide for Performing Arc-Flash Hazard
Calculations, IEEE 1584-201871 which gives the engineer the methods for calculating
the severity of arc flash incident energy levels. The NEC The National Electrical Code,
NFPA 70,72 requires only that certain equipment (switchboards, panelboards,
industrial control panels, meter socket enclosures, and motor control centers in other
than dwelling occupancies and likely to require examination, adjustment, servicing, or
maintenance while live) be field marked to warn qualified persons of potential electric
arc flash hazards.
70. NFPA 70E, The National Fire Protection Association, 2021 Edition.
71. IEEE 1584-2018, September 2018.
72. The National Fire Protection Association, Inc., 2020 Edition.
prohibited approach boundary, if the person is insulated from any other conductive
object, or if the live part is insulated from the person and from any other conductive
objects at a different potential. Unqualified persons must stay outside the limited
approach boundary unless they are escorted by a qualified person. Unqualified
persons cannot cross the restricted approach boundary.
An Arc Flash Risk Assessment must be performed to protect personnel from the
possibility of injury due to arc flash. This can be done by identifying the arc flash
hazard and using the table found in 130.5(C) to estimate the likelihood of occurrence
of an arc flash incident, and to determine if any additional protective measures are
required, including the use of PPE. This is called the “Arc Flash PPE Category
Method”. Using this method is only permissible if equipment is in a proper state as
recommended by the manufacture. The second method permissible for determine the
appropriate PPE is by performing an 130.5(G) Incident Energy Analysis Method.
Arc flash boundary is covered in 130.5(E) which can be determined two ways:
• The arc flash boundary shall be the distance at which the incident energy equals
1.2 cal/cm2 or
• using the Table 130.7(C)(15)(a) or Table 130.7(C)(15)(b).
Flash
Protection
Boundary
Limited
Approach
Boundary
Limited
Space
Any point on an
exposed, energized
electrical conductor
or circuit part
Restricted
Approach
Boundary
Restricted
Space
Prohibited
Approach
Boundary
Prohibited
Space
Approach Boundaries, page 154 is from Standard for Electrical Safety in the
Workplace75
Per 130.5(H) All electrical equipment, such as switchboards, panelboards, industrial
control panels, meter socket enclosures, and motor control centers that are in other
than dwelling units and that are likely to require examination, adjustment, servicing, or
maintenance while energized shall be marked with a label containing all of the
following information. (1) Nominal system voltage (2) Arc flash boundary (3) At least of
the of the following: Available incident energy, minimum arc rating of clothing or site-
specific level of PPE.
The classifications for personal protective equipment (PPE) for arc flash protection
are given in NFPA Table 130.7 (C)(15)(c), reproduced below as Table 22. PPE for arc
flash protection is given an Arc Rating in cal/cm2, which must be compared to the arc
flash incident energy for the location in question to select the proper clothing.
Employees working within the flash protection boundary must wear nonconductive
head protection wherever there is a danger of head injury from electric shock or burns
or from flying objects resulting from electrical explosion. Face, neck, chin and eye
protection must be worn wherever there is a danger of injury from electric arcs or
flashes or from flying objects resulting from electrical explosion. Body protection, in
the form of flame-retardant (FR) clothing as defined in Personal Protective Equipment
(PPE), page 155, must be worn where there is possible exposure to arc flash incident
energy levels of 1.2 cal/cm2; an exception allows Category 0 clothing to be worn for
exposures of 2 cal/cm2 or lower. An example of a full flash suit is shown in Example of
a Full Flash Suit, page 156.
Protective Equipment
Hard hat
Safety glasses or safety goggles (SR)
Hearing protection (ear canal inserts)cHeavy-duty leather gloves, arc-rated gloves, or rubber insulating gloves with leather
protectors (SR)d
Leather footwear e (AN)
Protective Equipment
Hard hat
Safety glasses or safety goggles (SR)
Hearing protection (ear canal inserts) c
Heavy-duty leather gloves, arc-rated gloves, or rubber insulating gloves with leather protectors (SR) d
Leather footwear e
3 Arc-rated Clothing Selected so that the System Arc Rating Meets the Required Minimum Arc Rating of 25 cal/
cm2 (104.7 J/cm2)a
Arc-rated long-sleeve shirt (AR)
Arc-rated pants (AR)
Arc-rated coverall (AR)
Arc-rated arc flash suit jacket (AR)
Arc-rated arc flash suit pants (AR)
Arc-rated arc flash suit hood
Arc-rated gloves or rubber insulating gloves with leather protectors (SR) d
Arc-rated jacket, parka, high-visibility apparel, rainwear, or hard hat liner (AN) f
Protective Equipment
Hard hat
Safety glasses or safety goggles (SR)
Hearing protection (ear canal inserts) c
Leather footwear e
75. NFPA 70E, The National Fire Protection Association, 2021 Edition.
Protective Equipment
Hard hat
Safety glasses or safety goggles (SR)
Hearing protection (ear canal inserts) c
Leather footwear e
AN: As needed (optional). AR: As required. SR: Selection required.
a Arc rating is defined in Article 100.
bFace shields are to have wrap-around guarding to protect not only the face but also the forehead, ears, and neck, or,
alternatively, an arc-rated arc flash suit hood is required to be worn.
cOther types of hearing protection are permitted to be used in lieu of or in addition to ear canal inserts provided they are
worn under an arc-rated arc flash suit hood.
d Rubber insulating gloves with leather protectors provide arc flash protection in addition to shock protection. Higher
class rubber insulating gloves with leather protectors, due to their increased material thickness, provide increased arc
flash protection.
e Footwear other than leather or dielectric shall be permitted to be used provided it has been tested to demonstrate no
ignition, melting or dripping at the minimum arc rating for the respective arc flash PPE category.
fThe arc rating of outer layers worn over arc-rated clothing as protection from the elements or for other safety purposes,
and that are not used as part of a layered system, shall not be required to be equal to or greater than the estimated
incident energy exposure.
76. NFPA 70E, The National Fire Protection Association, 2021 Edition.
IEEE 1584
IEEE 158477 is the guide for determining arc flash incident energy levels and
protection boundaries. It contains an empirical calculation method based upon
extensive test results using a Design-of-Experiments (DOE) method, resulting in a
95% confidence level. In situations where the empirical method does not apply, the
“Lee” method from Lee, R., “The Other Electrical Hazard: Electrical Arc Blast Burns,”78
is recommended, and is described in IEEE 1584. IEEE 1584 only considers the heat
of an arc, and not the secondary effects such as molten metal spatter and pressure-
wave effects.
IEEE 1584 Empirical Method
This method is valid for the following systems with the following characteristics:
• Voltages in the range of 208 V–15 kV, three-phase
• Frequencies of 50 Hz or 60 Hz
• Bolted fault current in the range:
◦ LV: 500–106 kA
◦ MV: 200–65 kA
• Grounding of all types, not a variable in 2018 calculations
• Standard box per voltage level; Max dimension “49”
• Gaps between conductors:
◦ LV: 6.35–76.2 mm
◦ MV: 19.05–254 mm
• Working distance >12 in.
• Electrode configurations:
◦ VCB: Vertical electrodes in a cubic box enclosure (equivalent to 2002 in box)
◦ VCBB: VCB with electrodes terminating in an insulating barrier
◦ HCB: Horizontal electrodes in a cubic box enclosure
◦ VOA: Vertical open-air (equivalent to 2002 open air)
◦ HOA: Horizontal open-air
• Enclosure Size Correction Factor: enclosure correction factor de-rates incident
energy for larger-than-standard box sizes.
Steps for performing calculations:
1. Determine electrode configuration.
2. Arcing current calculation.
• Ibf is the bolted fault current for three-phase faults (symmetrical rms) (kA)
• Iarc 600 is the average rms arcing current at Voc = 600 V (kA)
• Iarc 2700 is the average rms arcing current at Voc = 2700 V (kA)
• Iarc 14300 is the average rms arcing current at Voc = 14300 V (kA)
• G is the gap distance between the electrodes (mm)
• k1 to k 10 are the coefficients provided in Table 5 of IEEE1584–2018
• lg is log10
3. Determine clearing time based on the arcing current in Step 1. The use of time
current curves (TCC) may be used for this step. Ensure the clearing time
considers this such as the condition of equipment, alternate fault sources, or time
delays in the control circuit.
77. IEEE Guide for Performing Arc Flash Hazard Calculations, IEEE 1584-2018, September 2018.
78. IEEE Transactions on Industry Applications, vol. 1A-18, no. 3, May/June 1982.
AFB≤600 = {k1+ k2lgG + (k3I arc 600/k4Ibf7 + k5Ibf6 + k6Ibf5 + k7Ibf4 (8–2)
+ k8Ibf3 + k9Ibf2 + k10Ibf) + k11lgIbf + k13lgIarc + lg(1/CF)] - lg(20/
T)]/-k12
Arcing Current Variation: First pass in the calculations is done with 100% arcing
current, as used in Step 2. To ensure the worst case is used, the following formulas
are used to provide minimum arc rating. Use the worst case incident energy.
The incident energy is proportional to the arcing time, which is set by the overcurrent
protective device time-current characteristic and the arcing current level. Because
overcurrent protective device tripping times are lower for larger currents due to
inverse time-current characteristics, this is an important point. Larger bolted fault
currents lead to larger predicted arcing fault currents, which lead to generally lower
values of arc flash incident energy. Lower bolted fault currents lead smaller predicted
arcing fault currents, which lead to generally higher values of incident energy.
“Lee” Method
Where the IEEE 1584 empirical method cannot be used due to being outside the limits
of applicability as defined above, the theoretically-derived “Lee” method per A. C.
Parsons, “Arc Flash Application Guide Arc Flash Energy Calculations for Circuit
Breakers and Fuses”79 may be used. This is based upon maximum power transfer
and is very conservative above 15 kV. To calculate the incident energy with this
method, the following equations are used (see IEEE Guide for Performing Arc Flash
Hazard Calculations80):
Application Guidelines
Arc Flash Calculations
The following guidelines are helpful when performing arc flash calculations (see The
National Electrical Code, NFPA 70,81):
• When choosing a calculation method, be sure the system conditions fall into the
calculation method’s range of applicability.
• Use the newest methods given in IEEE 1584-2018. Older methods given in
previously published papers are superseded by this standard.
• If the manufacturer publishes device-specific equations, use them.
• Use realistic fault current values. The actual minimum available fault current,
rather than the worst-case values typically used for short-circuit analysis, give
more conservative (and realistic) results.
• Consider the effects of arc fault propagation to the line side of the main
overcurrent device when determining which device to use to calculate the arcing
time. For example, for the electrical panel in Example Electrical Panel, page 159,
device A would be used rather than device B for calculating the arcing time for a
fault on the panelboard bus, since the fault can propagate to the line side of
device B. Make similar considerations for switchboards, MCC’s.
• Quantify the variables. The working distance, bus gap, equipment configuration,
and system grounding are all dependent upon the particular installation and must
be accurately determined.
• Be aware of motor contribution. Motor contribution can both increase and
decrease the arc flash incident energy, depending upon where in the system the
arcing fault occurs.
• Use a computer for analysis. This is the most efficient way to accurately calculate
the incident energies and flash protection boundaries where multiple sources,
such as generation and motor contribution, must be taken into account. Several
commercial software packages are available for arc flash hazard analysis. Be
aware, though, what the user-configurable options for the software are and be
sure they are set correctly for accurate results.
System Design
Arc flash hazard analysis is typically performed after the system design process,
including the time-coordination study, is complete. This can result in the need for
“tweaking” of overcurrent protective device settings to obtain acceptable arc flash
results or, in the worst case, system re-design with additional equipment. The
following guidelines, if observed during the system design phase, can serve to
minimize the need for such activities:
Arc Flash Protection Protection Reduced Recovery Impact Impact Modify- Ca- OpEx
Mitigation During During Incident Energy Time on on ing pEx
Types Operation Maintenance / (cal / cm2) Foot- Commis- Existing
Abnormal print sioning Equip-
Operation ment
Prevention By Protection Protection Reduced Recov- Impact Impact Modify- CapEx OpEx
Design: Arc During During Incident Energy ery Time on on ing
Flash Operation Maintenance / (cal / cm2) Foot- Commis- Existing
Mitigation Abnormal print sioning Equip-
Types` Operation ment
Prevention By Protection Protection Reduced Recovery Impact Impact Modify- CapEx OpEx
Design: Arc During During Incident Energy Time on on ing
Flash Operation Maintenance (cal / cm2) Foot- Commis- Existing
Mitigation / Abnormal print/ sioning Equip-
Types` Operation Commis- ment
sioning
Energy Limited Limited Less than 8/12 Hours / None Low Possible $$ $
reducing daysa, /Low
page 162
maintenance
switch depend-
ing on
ERMS
switch
been turn-
ed on
Circuit breaker Limited Limited Less than 8/12 Hours / None Medium Limited $$ $
with days /Medium
instantaneous
or override
below arcing
level
84. A. C. Parsons, “Arc Flash Application Guide Arc Flash Energy Calculations for Circuit Breakers and Fuses”, Square D/Schneider Electric
Engineering Services, August 2004.
Current-limiting Limited Limited Less than 8/12 Hours / Medium Low Limited $$ $
circuit breakers days /Low
/fuses
Digital multi- Yes Yes Less than 8/12 Weeks / Low/High High Possible $ $
function relay months
Zone selective Yes Yes Less than 12 Hours / None Medium Possible $$ $
interlocking days a, /Medium
page
162de-
pending
on calorie
availabili-
ty
Differential Limited Limited Less than 8/12 Hours / Low/High High Possible $$$ $
protection days
Transfer trip Yes Yes Less than 8/12 Hours / Low Medium Possible $$ $
scheme (virtual days /Medium
main)
Arc flash Yes Yes Less than 8/12 Hours / Medium Medium Applica- $$ $
detection days /Medium tion
device (optical depend-
sensors) ent
High speed Yes Yes Less than 1.2 Hours / High/High High Possible $ $
shorting switch days
(quenchers)
Line side Yes Yes Less than 1.2 Hours / Low/Low Low Possible $ $
isolation with days
passive
reduction
a For additional information on arc flash mitigation methods, refer to Schneider Electric Arc Flash Mitigation Guide V11.
NFPA 70E article 205.32 states that a single-line diagram, where provided for the
electrical system, shall be maintained in a legible condition and shall be kept current.
If utilizing the incident energy analysis method of determine the arc flash hazard, the
analysis must be reevaluated when changes occur in the electrical system that could
affect the results of the analysis and reevaluated at intervals not to exceed five years,
per NFPA 70E article, 130.5(G). An effective method of accomplishing this is to have a
study performed and maintained to be kept current. The concept of a digital twin can
accomplish this.
The general approach of a digital twin model is a virtual representation of a distribution
system that can be as simple as a single-line diagram with relevant data to complex,
updated real-time data, via digital readings through networked metering and
communication. Digital twins can create a foundation for the customer and clients to
support reliability assessment, asset management, real-time interfaces for SCADA
systems, and system studies.
Introduction
Power Distribution Equipment is a term generally used to describe any apparatus
used for the generation, transmission, distribution, or control of electrical energy. This
section concentrates upon commonly used power distribution equipment:
Panelboards, Switchboards, Low-Voltage Motor Control Centers, Low-Voltage
Switchgear, Medium Voltage Power and Distribution Transformers, Medium-Voltage
Metal Enclosed Switchgear, Medium Voltage Motor Control Centers, and Medium-
Voltage Metal-Clad switchgear. Each has its own unique standards and application
guidelines, and one facet of good power system design is the knowledge of when to
apply each type of equipment and the limitations of each type of equipment. The
equipment described herein are typically custom-engineered on a per-order basis.
effects on the equipment due to the ingress of water (rain, sleet, snow, splashing
water, and hose directed water); that provides an additional level of protection against
corrosion; and that is undamaged by the external formation of ice on the enclosure.
Type 5: Enclosures constructed for indoor use to provide a degree of protection to
personnel against access to hazardous parts; to provide a degree of protection of the
equipment inside the enclosure against the ingress of solid foreign objects (falling dirt
and settling airborne dust, lint, fibers, or other items); and to provide a degree of
protection with respect of harmful effects on the equipment due to the ingress of water
(dripping and light splashing).
Type 12: Enclosures constructed (without knockouts) for indoor use to provide a
degree protection to personnel against access to hazardous parts; to provide a
degree of protection of the equipment inside the enclosure against ingress of solid
foreign objects (falling dirt and circulating dust, lint, fibers, or other items); and to
provide a degree of protection with respect to harmful effects on the equipment due to
the ingress of water (dripping and light splashing).
Major Industry Standards UL 50, UL 67, CSA C22.2 No. 29, CSA C22.2 No. 94, NEMA PB 1,
Federal Specification W-P-115C, NEC
86. NFPA 70, The National Fire Protection Association, Inc., 2003 Edition.
Figure 96 - Panelboards
Panelboards are used to group the overcurrent protection devices for several circuits
together into a single piece of equipment. In small installations they may serve as the
service equipment. The NEC The National Electrical Code87 divides panelboards into
two categories:
Lighting and Appliance Branch-Circuit Panelboard: A panelboard having more
than ten percent of its overcurrent devices protecting lighting and appliance branch
circuits.
Power Panelboard: A panelboard having ten percent or fewer of its overcurrent
devices protecting lighting and appliance branch circuits.
Separated Distribution Panelboard: A panelboard combining the above Lighting
and Appliance Branch-Circuit and Power Panelboards.
Lighting and appliance branch-circuit panelboards are limited to a maximum of 42
overcurrent devices, excluding mains. UL 67 UL Standard for Safety for Panelboards,
UL 67, Underwriters Laboratory, Inc.88 designates Class CTL Panelboard as the
marking for appliance and branch circuit panelboards; CTL stands for “circuit limiting”.
In some manufacturer’s literature lighting and appliance branch-circuit panelboards
for residential or light commercial use are referred to as load centers.
Panelboards are available with built-in main devices or as main lugs only (MLO). The
NEC The National Electrical Code87 requires appliance and branch circuit
panelboards to be individually protected on the supply side by not more than two main
circuit breakers or two sets of fuses having a combined rating no greater than the
rating of the panelboard. Lighting and appliance branch circuit panelboards are not
required to have individual protection if the feeder overcurrent device is no greater
than the rating of the panelboard.
87. NFPA 70, The National Fire Protection Association, Inc., 2003 Edition.
88. November 2003.
Switchboards
Table 30 - Switchboards: Quick Reference
89. NFPA 70, The National Fire Protection Association, Inc., 2003 Edition.
90. NFPA 70, The National Fire Protection Association, Inc., 2005 Edition.
Figure 97 - Switchboards
Switchboards are available with a main circuit breaker or fusible switch, or as main
lugs only. The available ampacities and multi-section availability makes them more
flexible than panelboards. They are generally available utilizing either copper or
aluminum bussing, and with a variety of bus plating options. Custom bussing for
retrofit applications is also possible.
Switchboard circuit breakers may be stationary-mounted (also referred to as fixed-
mounted), where they can be removed only by unbolting of electrical connections and
mounting supports, or drawout-mounted, where they can be without the necessity of
removing connections or mounting supports. It is possible to insert and remove
drawout devices with the main bus energized. The section that contains the main
circuit breaker(s) or service disconnect devices is referred to as a main section. A
section containing branch or feeder circuit breakers is referred to as a distribution
section.
Devices mounted in the switchboard may be either panel mounted (also referred to as
group mounted), where they are mounted on a common base or mounting surface, or
individually mounted, where they do not share a common base or mounting surface.
Individually mounted devices may or may not be in their own compartments. A device
which is segregated from other devices by metal or insulating barriers and which is not
readily accessible to personnel unless special means for access are used is referred
to an isolated device. Group and Individually-mounted Devices, page 168 shows
examples of sections with group and individually-mounted devices.
The main through-bus is often referred to as the horizontal bus. The bussing in a
section which connects to the through-bus is referred to as the section bus (also
known as vertical bus). The bussing that connects the section bus to the overcurrent
devices is referred to as the branch bus. Section and branch busses may be smaller
than the main through-bus; if this is the case UL 891 (see The National Electrical
Code91) gives the required section bus size as a function of the number of overcurrent
devices connected to it.
91. NFPA 70, The National Fire Protection Association, Inc., 2005 Edition.
Group-mounted Individually-mounted
MCCs are classified into two classes by UL Standard for Safety for Motor Control
Centers93and Motor Control Centers94:
Class I Motor Control Centers: Mechanical groupings of combination motor control
units, feeder tap units, other units, and electrical devices arranged in an assembly.
Class II Motor Control Centers: A Class I motor control center provided with
manufacturer-furnished electrical interlocking and wiring between units, as specifically
described in overall control system diagrams supplied by the user.
MCC wiring is classified by UL Standard for Safety for Motor Control Centers93and
Motor Control Centers95 into three types:
Type A Wiring: User (field) load and control wiring are connected directly to device
terminals internal to the unit; provided on Class I MCCs only.
Type B Wiring: User (field ) control wiring is connected to unit terminal blocks; the
field load wiring is connected either to power terminal blocks or directly to the device
terminals.
Type C Wiring: User (field control wiring is connected to master terminal blocks
mounted at the top or bottom of vertical sections which contain combination motor
control units or control assemblies; the field load wiring is connected to master power
terminal blocks mounted at the top or bottom of vertical sections or directly to the
device terminals.
MCCs generally consist of a common power bus and a vertical bus for each section to
which combination motor controllers are plugged on. The combination starter consists
of motor starter, fuses or circuit breakers, and power disconnect. MCCs may also
have push buttons, indicator lights, variable-frequency drives (VFDs), programmable
logical controllers (PLCs), and metering equipment. The individual plug-in units are
often referred to as buckets and may be inserted and removed with the main bus
energized so long as the disconnecting device for the individual unit is open. A vertical
wireway is supplied internal inter-unit connections and field connections within each
section.
MCCs offer the opportunity to group several motor starters together in one location
with a space-efficient footprint versus individual control cabinets, and like
switchboards are available with many options. Removable plug-on units allow quick
change-outs if spare units are kept on hand for the most common sizes of starters in
the facility. Low-voltage soft-starters and variable-speed drives may also be mounted
within MCCs.
Low-voltage Switchgear
Table 32 - Low-voltage Switchgear: Quick Reference
Major Industry Standards ANSI/IEEE C37.20.1, ANSI/IEEE C37.51, NEMA SG-5, CAN/CSA
C22.2 NO 31-M89, UL 1558
Low-voltage Transformer
“A transformer is a static electrical device that transfers energy between two or more
circuits through electromagnetic induction” (https://www.se.com/). The main
applications of a transformer include converting utility voltage to building distribution
voltage, and converting distribution voltage to application voltage requirements.
These units can be classified into multiple types: Dry-type, Control Power, and Mini
Power-zone.
Table 33 - Low-voltage Transformers: Quick Reference
Available Primary Voltage Ratings 208, 240, 480, 600 Vac Delta
Available Secondary Voltage Ratings 208Y/120, 240 Vac Delta 120V, 480Y/277, 204/120
Figure 101 - Low Voltage Dry-type Transformers. Left to right: Dry-type Ventilated Enclosure, Dry-type Non-
ventilated Enclosure, Open Core and Coil
Additionally, the impedance of the supply system can influence the amount of inrush
current the transformer can draw. To avoid tripping circuit breakers or blowing fuses
on the primary side of the transformer during energizing, it is essential to carefully
coordinate fuse sizes or breaker handle ratings, and magnetic trip settings. To provide
optimal coordination and minimize inrush nuisance tripping, adjust the primary
overcurrent protection based on the maximum inrush current. This results in the
primary overcurrent protection exceeding the 125% allowance in the NEC for primary-
only protection, and secondary protection is required (see 7400CT1901).
Due to concerns regarding the impact on the efficiency of the transformers and market
needs for improvements in energy consumption, low voltage distribution transformers
are regulated through the Energy Policy and Conservation Act (7400CT1901). The
Department of Energy (DOE) evaluates and sets minimum efficiency standards for
low voltage dry-type distribution transformers. Transformer efficiency can be defined
as the percentage of power out compared to the percentage of power in. A perfect
zero loss transformer would have the same power in as out and would be 100%
efficient (see 7400CT1901). The efficiency of the transformers shall be no less than
that which is required for their kVA rating, as shown in Energy Conservation
Standards for Low Voltage Dry-type Distribution Transformers, page 173.
Table 34 - Energy Conservation Standards for Low Voltage Dry-type Distribution Transformers
15 97.70 15 97.89
25 98.00 30 98.23
37.5 98.20 45 98.40
50 98.30 75 98.60
75 98.50 112.5 98.74
100 98.60 150 98.83
167 98.70 225 98.94
250 98.80 500 99.14
Table 34 - Energy Conservation Standards for Low Voltage Dry-type Distribution Transformers (Continued)
— — 750 99.23
— — 1000 99.28
The following low-voltage dry-type transformers must comply with the DOE efficiency
standards as shown in Energy Conservation Standards for Low Voltage Dry-type
Distribution Transformers, page 173:
• Three- and single-phase transformers
• Step-up and step-down transformers
• General purpose ventilated transformers
• Harmonic-mitigating transformers
• General purpose open core and coil transformers
The following low-voltage dry-type transformers are not required to comply with the
efficiency standards as shown in Energy Conservation Standards for Low Voltage
Dry-type Distribution Transformers, page 173:
• Auto-transformers
• Drive-isolation transformers
• Non-ventilated transformers
• Resin-encapsulated transformers
• Buck-boost transformers
• Control-power transformers
• Medical isolation panel transformers
Control Power: Industrial control power transformers are designed with low
impedance windings for voltage regulation and can accommodate the high inrush
current associated with contractors, starters, solenoids, and relays. Their function is to
meet the diverse needs of panel builders and machinery OEMs. They are typically 50/
60 Hz rated and are designed with various temperature classes as shown in
Temperature Rises for Low-voltage Transformers, page 174.
Mounting
Brackets
Lifting
Eye
Transformer
Durable,
permanently
legible nameplate
Hinged cover —
weather-protective
with horizontal
locking features
Ample knockouts
accessible through
bottom of enclosure
Medium-voltage power and distribution transformers are used for the transformation
of voltages for the distribution of electric power. They can be generally classified into
two different types:
Dry-Type: The windings of this type of transformer are cooled via the circulation of
ventilating air. The windings may be one of several types, including Vacuum Pressure
Impregnated (VPI), Vacuum Pressure Encapsulated (VPE), and cast-resin. The cast-
resin types generally are more durable and less likely to absorb moisture in the
windings than the VPI or VPE types. In some cases, the primary windings are cast-
resin and the secondary windings are VPI or VPE.
Liquid-Filled: The windings of this type of transformer are cooled via a liquid medium,
usually mineral oil, silicone, or paraffinic petroleum-based fluids.
Liquid-filled units have a generally low in first-cost, but the requirements in NEC The
National Electrical Code, NFPA 7098 Article 450 must be reviewed so that installation
requirements can be adequately met, and maintenance must be taken under
consideration since fluid levels should be monitored and the condition of the fluid
examined on a regular basis. They have an expected service life of approximately 20
years. VPE and VPI dry-type transformers also generally have low first-costs, have
longer lifetimes than liquid-filled units, and are much easier than liquid-filled types to
install indoors; however, give consideration to the absorption of moisture by the
windings if these are used outdoors. Installed indoors, these have expected service
lifetimes of around 30 years. Cast-resin dry-type transformers have generally high
first-costs compared to the other types but have the same installation requirements as
dry-type transformers and have the longest expected service life (around 40 years).
Enclosure styles may also be divided into two basic types: pad-mounted, which is a
totally-enclosed type generally mounted outdoors and with specific tamper-resistance
features to minimize inadvertent access by the general public, and unit substation
type, which is an industrial-type enclosure suitable for close-coupling into an
integrated unit substation lineup with primary and secondary equipment (unit
substation-style transformers may also be equipped with cable termination
compartments as well).
Medium Voltage Power and Distribution Transformers, page 177 shows typical
examples of medium-voltage power and distribution transformers.Top to bottom: cast-
coil dry type with unit substation-style enclosure, VPI dry-type with unit substation-
style enclosure, liquid-filled type with unit substation-style enclosure, and dry-type with
pad-mounted enclosure.
1.2 10 30
2.5 20 45
5.0 30 60
7.2 30 60
8.7 45 75
15.0 60 95
25.0 110 125
35.0 N/A a, page 177 150
a VPI/VPE dry-type transformers are typically not available above 25.0 kV Class.
Liquid-filled 55/65 or 65
Impedance levels vary; the manufacturer must be consulted for the design impedance
of a specific transformer. In general, units 1000–5000 kVA typically have 5.75%
impedance ± 7.5% tolerance.
Rated Maximum Voltage (kV) Power Frequency Withstand Impulse Withstand (kV)
(rms) (kV)
4.76 19 60
8.25 36 95
15.0 36 95
27.0 60 125
38.0 80 150
100. IEEE Standard for Metal-Enclosed Interrupter Switchgear, IEEE Std. C37.20.3-2001, August 2001.
Internal barriering requirements for medium-voltage areas within the switchgear are
minimal. All low-voltage components are required to be separated by grounded metal
barriers from all medium-voltage components. Interlocks are required to minimize
access to medium-voltage fuses while their respective switch is open and to minimize
closing their respective switch while they are accessible. In the rare case that this type
of switchgear contains drawout devices, shutters must be provided to minimize
accidental contact with live parts when the drawout element is withdrawn.
Available options for this type of switchgear include shunt trip devices for the
switches, motor operators for the switches, blown fuse indication. Relaying of any
type, including voltage relaying, must be carefully reviewed to avoid exceeding the
limits of the switches. The application of overcurrent relaying to this type of switchgear
is not recommended unless a short-circuit interrupting element is included, such as a
vacuum interrupter.
Medium-voltage motor controllers are used to control the starting and protection for
medium-voltage motors. They generally utilize vacuum contactors rated up to 400 A
continuous, in series with a non-load-break isolation switch and R-rated fuses, fed
from a common power bus. The motor starting methods in Arc Flash Hazard, page
150 are all generally supported, including soft-start capabilities. Class E2 units per
Industrial Control and Systems: Medium Voltage Controllers Rated 2001 to 7200 Volts
AC101, which employ fuses for short-circuit protection, are generally the most
common. Medium-voltage MCC, page 181 shows a typical example of a medium-
voltage MCC.
Medium-voltage MCC’s are generally available with several options depending upon
the manufacturer, including customized control and multi-function microprocessor-
based motor protection relays The contactors are generally of roll-out design to allow
quick replacement.
Above 7200, metal-clad switchgear is generally used for motor starting.
102. IEEE Standard for Metal-Clad Switchgear, IEEE Std. C37.20.2-1999, July 2000.
• The door through which the circuit interrupting device is inserted into the housing
may serve as an instrument or relay panel and may also provide access to a
secondary or control compartment within the housing.
Medium-voltage metal-clad switchgear is generally used as the high-level distribution
switchgear for medium- to large-sized facilities. It is also the preferred choice for
service entrance equipment for these types of facilities.Metal-clad Switchgear, page
182 shows an example of metal-clad switchgear.
Retrofit Solutions
Retrofit solutions are pre-engineered solutions that are designed with a new breaker
element truck and carriage used to be installed in the existing switchgear. The
interfaces with the existing structure maintain all safety interlocks inherent with the
original design. The interior of the existing cell is not modified. These types of
solutions work well with any brand of legacy switchgear.
103. IEEE Standard for Metal-Clad Switchgear, IEEE Std. C37.20.2-1999, July 2000.
Table 44 - NFPA 110 Emergency Power Supply System Types NFPA 110 Table 4.1
(b)
10 10 seconds
Table 44 - NFPA 110 Emergency Power Supply System Types NFPA 110 Table 4.1
(b) (Continued)
60 60 seconds
120 120 seconds
M Manual stationary or nonautomatic – no time limit
The Class of an EPSS refers to the minimum time, in hours, for which the system is
designed to operate at its rated load without being refueled or recharged. The classes
for emergency power systems are shown in NFPA 110 Emergency Power System
Classes NFPA 110 Table 4.1(A), page 186.
Table 45 - NFPA 110 Emergency Power System Classes NFPA 110 Table 4.1(A)
2 2 hour
6 6 hour
48 48 hour
X Other time, in hours, as required by the application, code, or user.
The Level of an EPSS refers to the level of equipment installation, performance, and
maintenance requirements. The levels for emergency power systems are shown in
NFPA Emergency Power System Levels, page 186.
2 When loss of the equipment to perform is less critical to human life and safety and
where the authority having jurisdiction shall permit a higher degree of flexibility than
that provided by a Level 1 system (A4.4.2)
NFPA 101
NFPA 101 [4], Life Safety Code, addresses those construction, protection, and
occupancy features necessary to minimize danger to life from fire, including smoke,
fumes, or panic. It defines the requirements for what systems the Emergency Power
System supplies.
NFPA 99
NFPA 99 defines establishes criteria to minimize the hazards of fire, explosion, and
electricity in health care facilities. It defines several specific features of electric power
systems for these facilities.
Power Sources
Generators are the most prevalent source of power for emergency and standby power
systems. For most commercial and industrial power systems these are engine-
generator sets, with the prime-mover and the generator built into a single unit. For
reciprocating engines, diesel engines are the most popular choice of prime-mover for
generators, due to the cost of the diesel engines as compared to other forms of power
and the relative ease of application. Engine generator sets can also run on natural
gas, however natural gas engines typically have slower response times, than diesel
units.
Another option available is the turbine generator, typically powered by natural gas.
Gas-turbine generator sets are generally lighter in weight than diesel engine-
generator sets, run more quietly, and generally require less cooling and combustion
air, leading to lower installation costs. Turbine generators typically are utilized in large
capacity applications, when lengthy continuous operation is required and in combined
heat and power (CHP) applications. Gas-turbine generator sets are more expensive
and typically less efficient than diesel engine-generator sets. They have more
complex controls and require longer starting times (normally around 30 seconds
compared to the 10-15 seconds or less for diesels). The long starting-time
requirement, cost, and lack of available small sizes (< 500 kW) makes the gas-turbine
generators infeasible in most emergency and standby power applications.
Considerations for generator installations include the combustion and cooling air
required by the generator and prime mover, provisions for the removal of exhaust
gases, noise abatement, expected run time hours and emissions. These
considerations can increase the installation costs, especially for diesel engines. Fuel
supply must also be considered; building code and insurance considerations may
force the fuel storage tank to be well removed from the generator(s), usually forcing
the addition of a fuel transfer tank near the generator(s).
Engine-generator sets must be sized properly for the application. Several ratings exist
for the output capability of an engine generator set. The continuous rating is typically
the output rating of the engine-generator set on a continuous basis with a non-varying
load. The prime power rating is typically the continuous output rating with varying
load. The standby rating is typically the output rating for a limited period of time with
varying load. The manufacturer must be consulted to define the capabilities of a given
unit.
Automatic-transfer Switches
A means must be provided to switch the critical loads from the normal utility source to
the standby power source. Several types of devices are available for this.
An Automatic Transfer Switch (ATS) is defined as “self-acting equipment for
transferring one or more load conductor connections from one power source to
another” [1]. Automatic-transfer switches are the most common means of transferring
critical loads in Emergency Power Supply Systems (EPSS). An automatic-transfer
switch consists of a switching means and a control system that senses both normal
and emergency supply voltage and frequency. The major functions of an automatic-
transfer switch include the following:
• Carry current continuously
• Detect power failures
• Initiate the alternate source (Send a start signal to an engine generator)
• Transfer load
• Sense restoration of the normal source
• Retransfer load to the normal source
• Withstand and close on fault currents – ATS Withstand Current Rating (WCR)
An automatic-transfer switch determines when an outage occurs and after an
adjustable time delay (Typically one to six seconds in the event of a momentary
outage) sends a start signal to the emergency generator. Upon sensing the generator
has achieved acceptable voltage and frequency the automatic-transfer switch
transfers the load to the alternate source. When the normal source returns and after
an adjustable time delay the automatic-transfer switch controls sense the normal
source voltage and frequency and transfer the load back to the normal source when it
has achieved acceptable voltage and frequency levels. automatic-transfer switches
are available in ratings from 30 - 4000 A, and up to 600 V [1]. Low Voltage Transfer
Switches are listed to UL1008. There are options for medium-voltage transfer
switches up to 15 kV that are listed to UL 1008A. Because automatic-transfer
switches are designed to continuously carry the loads they serve, even under normal
conditions, take care in sizing these so that the potential for loss is minimized.
Similarly, pay attention to available short circuit current at each transfer switch so that
proper Withstand Current Rating (WCR) capability is provided. Automatic-test
switches with adjustable pickup and dropout setpoints and integral testing capability
should be included.
In Bypass/isolation Switch Application, page 190 the bypass switch contacts bypass
the automatic-transfer switch, and isolation contacts serve to isolate the automatic-
transfer switch. The Bypass/isolation switch is typically manually operated. Bypass/
isolation switches are available with a "make before break” feature allowing bypass of
the automatic-transfer switch to be completed without disconnecting the load.
Automatic-transfer/Bypass-isolation switches are available in Open-transition, Closed-
transition and Delayed-transition configurations as described above.
• Line Interactive: This product is similar to standby, with the additional feature of
an autotransformer that adjusts to ongoing sag and surge conditions to provide
conditioned power. These are a lower cost options to compared to double
conversion and are typical implemented in small back office computer rooms, to
keep small servers and network switches energized during brief outages.
Surge
Suppressor Filter
AC
Source
Transfer
Switch
Output
Load
Battery
Charger Battery Inverter
Output
Load
Battery Converter
• Double Conversion: The majority of large scale three-phase UPS in the world
utilize double conversion systems. As the name suggests, takes input utility,
converts it to DC via a rectifier and charges the Direct Current (DC) source while
simultaneously converting back to AC off the DC bus via an inverter. This system
provides highly conditioned and controlled AC power to more susceptible critical
loads. A bypass static switch allows for continuous power to the load in the case
of power train loss in the double conversion system. These UPS modules
typically start at five - ten KVA single-phase and can be as large 1.5 MW.
Additional system capacity can be achieved by paralleling multiple units together.
Total system capacity is limited to paralleling bus and breaker sizes as well as
limiting control capabilities. Rarely are system capacities see greater than 3.2
MW.
Static Bypass
Switch
Rectifier Inverter
AC Output
Source Load
Battery
Use Examples
UPS systems are used in a wide variety of locations. In the home, personal desktop
computers may be backed up by a small offline/standby system. In small business
operations, the back-office server, phone switch, network equipment, and other
business critical components may be protected by a line-interactive UPS. Typically,
units are placed in the bottom of the two-post rack.
In larger applications, double-conversion UPS are used for a variety of purposes.
Applications include lighting inverter support for emergency lighting, CT/MRI/PET
scanning systems, data centers to back up the critical computing facility, and
manufacturing processes to prevent production losses. Many times, a UPS is used to
back up the controls portion of whatever process is occurring. This allows for the PLC,
computer, or whatever is controlling, for example, the manufacturing process to stay
energized through a temporary outage, and thus knows where the process left off until
power is restored for the process to continue.
Codes and Standards
Uninterruptible Power Supplies are UL listed conforming to UL1778. Special
application standards include UL924 for emergency lighting. Recent DC storage
adoptions have seen the rise of UL1973 for lithium batteries, and large-scale fire
testing to UL9540A test standards. DC Energy sources are directly coupled with the
UPS. They are governed by International Building Code (IBC), International Fire Code
(IFC), and National Fire Protection Agency (NFPA) standards, addressed later in this
document.
Design Considerations
The use of UPS topologies for different applications also requires specific design
considerations. This is specifically true for larger double-conversion UPS. The
following sub-sections outline the most common design considerations.
• Transformer versus Transformer-less: Double-conversion UPS have seen a
fundamental transformation over the last ten years. Prior generations of UPS
included output transformers (a transformer in series with the inverter) for
waveform shaping and electrical isolation. Some also included an input isolation
transformer, either for galvanic isolation, or to allow for special rectifier
configurations. With inverter technology improvements, transformers on both
input and output have been all but eliminated. The only common remaining
transformer is in UL924 lighting inverter systems at 480 V. Since lighting is many
times at 277 V off a 480/277 V system, the UPS takes 480 V 3 W+G input and
provides an output isolation transformer, 480-480/277 V 4 W+G, developing the
neutral to support lighting loads.
• Bypass: A second major design improvement has been in the static bypass
switch technology. Higher efficiency requirements have driven more efficient
modes, such as ECO or E-Conversion, whereby power is transferred to the load
via the static bypass. In the event of a power outage, the UPS transfers back to
double conversion mode. This requires the static bypass to be a 100%
continuous duty rated component. Previous components were typically
momentary duty rated with wraparound contractors.
• Power Factor: Power factor and power capability of UPS have been transformed
in the last decade as well. In the 1990s, UPS were typically .8 power factor (pf).
For example, a 500 kVA UPS could only provide 400 kW of usable power. In the
early 2000s, these were improved to .9 pf (500 kVA/450 kW). Power electronics
improvements have now made unity power factor the de facto standard, where
kVA equals kW (kVA/kW).
• Input Voltages: Input versus output voltage of data center UPS remains a
common confusion point. Incoming utility transformation is typically to 480/277 V
main switchboard use. From here, lighting panels, mechanical loads, and the
UPS are fed. While the other equipment (lighting and mechanical loads) might
use the neutral and require it to be pulled, the UPS typically does not require, nor
desire, a neutral to be run. Downstream of the UPS, in the data center, or point of
use, most IT loads utilizes 208/120 V, or something similar. Therefore, a Power
Distribution Unit (PDU) incorporates a transformer to develop the 208/120 V
required. Thus, a UPS can be 480 V 3 W+G input and output, feeding the 480 V
input of the PDU to provide appropriate transformation. This can be a significant
cost saver in not requiring a neutral run, four-pole breakers for bypass operation,
complexity of operation, and other issues.
208 V is a common voltage for double conversion in the range between 10 kVA
and 150 kVA, and frequently used in small to medium size data centers, as well
as Medium Distribution Facilities (MDF) and larger edge computing deployments.
specific amp-hour), therefore they have very specific run times available, from as little
as two minutes to commonly five-seven minutes, depending on UPS KW. Ni-Zn will
have similar considerations to LiB.
These LiB installations are typically referred to as using “Power” batteries. As
described, they can discharge the entire stored energy rapidly (min. versus hours).
Energy batteries, also LiB, on the other hand, allow to discharge the stored energy
over a much longer period (hours vesus minutes). Just recently are these types of
implementation considered to support additional use cases, such as Demand
Reduction and Energy Arbitrage to reduce overall electrical utility charges.
Conclusion
UPS implementations are used to protect against power disruptions in many diverse
applications. The two main reasons are to protect against financial losses or life-safety
concerns. This includes data centers and hospitals or control systems and entire
manufacturing processes. While those fundamentals of the different UPS
technologies have not changed dramatically over the last decades, what has
changed, due to advancements in power electronic components, are improvements in
the overall efficiency, energy storage systems, and the reduction of physical size of
the UPS. We have witnessed the use of mainly transformerless UPS in system
designs, the use of Li-Ion batteries, and the shift from monolithic to modular UPS
architectures. The most used UPS technologies are offline/standby, line interactive
and double conversion. They range from a few hundreds of watts to multiple
megawatts.
Normal Emergency
Source Source
Automatic
Transfer
Switch
Normal
Loads
Emergency/Standby
Loads
For Emergency Power Systems with a single alternate power source, NEC Article
700.3(F) requires a means of connecting temporary or portable power, an example is
shown in Provisions for Connection of a Temporary or Portable Power Source, page
199.
Figure 120 - Low Voltage Multiple Generator Paralleling System with Multiple
Automatic Transfer Switches
Figure 121 - Prioritized Block Load Control and Prioritized Step Load Control
Figure 122 - Low Voltage Generator Paralleling System with Closed Transition
Automatic Transfer Switches
Multiple Generator Paralleling System with Segmented Bus and Tie Circuit
Breaker
Tie circuit breakers can be employed in Emergency and Standby Systems to achieve
connection of loads in ten seconds for Emergency Systems per NEC Article 700. In
this configuration the highest priority loads are split between the two bus segments. In
Segmented Bus Configuration, page 202 the tie circuit breaker 52T is normally open.
When the transfer switches sense a normal power outage all engine generators are
signaled to start. Individual generators are concurrently connected to isolated
segments of the bus without the need for synchronizing first, bringing multiple engine
generators online simultaneously and connecting the highest priority loads within the
ten second requirement. As additional engines become available, they are
synchronized with the connected generators, connected to the bus and the lower
priority loads are transferred to the emergency system. After all the generators are
connected and all the transfer switches have transferred their loads to the emergency
system the tie breaker can be signaled to close by the emergency system Master
Controls. Because the tie circuit breaker 52T is connecting sources together a
synchronizing device is required across the tie circuit breaker to synchronize the bus
segments prior to it closing. Configurations of this type are common in healthcare
applications.
When normal power is restored, several options are available to retransfer load back
to the normal source. Retransfer sequences can be open transition, closed transition
or closed transition/soft load. In the example shown in More Complex Medium Voltage
Main-Tie-Tie-Main system with Multiple Generators Configuration, page 203 all
retransfer sequences occur between the respective utility and main breakers as
follow:
Open Transition Retransfer: When normal power is restored main circuit breakers
52GM1 and 52GM2 are opened and utility circuit breakers 52U1 and 52U2 are then
closed in a “Break before Make” operation, assuming facility load. The engine
generator circuit breakers are opened, the engine generators enter a cooldown period
and then shutdown. Provide electrical Interlocks so that the respective main and utility
circuit breaker pairs from both cannot be closed at the same time.
Closed Transition Retransfer: When normal power is restored main circuit breakers
52GM1 and 52GM2 remain closed. Both utilities are synchronized with the live bus via
additional synchronizing devices provided for each utility. Once in synchronism circuit
breakers 52U1 and 52U2 are then closed in a “Make before Break” operation. Closed
transition overlap time is typically 100 milliseconds or less. Both main circuit breakers
52GM1 and 52GM2 are opened and each utility assumes facility load on its respective
bus segment. The engine generator circuit breakers are opened, the engine
generators enter a cooldown period and then shutdown. Additional synchronizing
controls are required at each utility for closed transition operation. Benefits of closed
transition operation include no interruption to loads when transferring between two
acceptable sources or during system test modes.
Soft Load/Closed Transition Retransfer: Soft load sequence is similar to the closed
transition sequence however the overlap time is extended, allowing load to be ramped
off of the generators on onto the utility. Interconnection can be approximately 30
seconds to several minutes. Benefits are similar to closed transition but also include
less wear and tear on circuit breakers and UPS systems. Utility company approval is
required for soft load operation and may require specific utility approved protective
relaying at each utility circuit breaker. Synchronizing devices are required at each
utility circuit breaker.
Essential Electrical Systems for hospitals are separated into three distinct branches
as described in NEC Article 517.32-34 as follows:
• Life Safety Branch: The Life Safety Branch of the EES provides power for
lighting, receptacles to those functions or warning systems that are required to
allow building occupants to safely leave the building in an emergency. Transfer
switches feeding the Life Safety Branch must be automatic and must be non-
delayed. Emergency power must be supplied to the life safety branch within ten
seconds of a normal source power outage. Typically, these loads are served by
automatic-transfer/bypass-Isolation switches. Often these switches are provided
with Closed Transition features. Wiring for the Life Safety Branch is kept
independent of all other wiring.
• Critical Branch: The Critical Branch serves loads that either have immediate
impact on patient well-being or are essential to the clinical functionality of the
health care facility. Transfer switches feeding the critical branch must be
automatic. Emergency power to the critical branch must be supplied within 10
seconds of a normal source power outage. Typically, these loads are served by
automatic-transfer/bypass-Isolation switches. Often these switches are provided
with Closed Transition features. Wiring for the Critical Branch is kept independent
of all other wiring.
• Equipment Branch: The Equipment Branch serves loads for major electrical
equipment required for patient care. The equipment branch of the EES consists
of large electrical equipment loads can include chillers, compressed air systems,
exhaust systems and sump pumps needed for patient care and basic facility
operation facility. Transfer switches feeding equipment loads are configured for
delayed connection to the emergency system.
An example of an Emergency System for a Hospital Application is shown below in
Hospital Essential Electrical System with Life Safety, Critical and Equipment
Branches, page 206.
Figure 127 - Hospital Essential Electrical System with Life Safety, Critical and Equipment Branches
Essential Electrical Systems are required to have two independent sources of power,
the normal utility source and a backup generator or multiple paralleled generators.
When normal source power is lost the generator(s) must be started and Life Safety
and Critical Branch Automatic Transfer Switches must transfer to the emergency
system and provide power to their loads within ten seconds of the normal source
power outage. Life Safety Branch transfer switches must transfer to the emergency
source immediately upon sensing the availability of emergency power without a delay.
Life Safety automatic transfer switches are always considered Priority 1 loads. Critical
Branch automatic transfer switches may have a delay to allow the Life Safety
automatic transfer switches to connect to emergency power first but must be
connected to emergency power within ten seconds of a normal power source outage.
Equipment Branch automatic transfer switches can be delayed and are always
considered a lower priority than Life Safety or Critical Branch ATS. Automatic transfer
switches are required for equipment loads that serve suction.
Hospital Essential Electrical Systems have specific testing requirements as follow:
• NFPA 110 Chapter 8 specifies that an Essential Power Supply System (EPSS)
including its transfer switches "shall be exercised under load at least monthly”.
• System testing is required 12 times a year, at intervals not less than 20, or more
than 40 days – All ATS must be tested monthly.
• The essential electrical system must be maintained to supply emergency power
within 10 seconds of loss of normal power. If the ten second criteria is not met
during regular testing, the organization must have a process to confirm on an
annual basis that the ten second criteria can be met. Joint Commission
requirement based on NFPA 99.
• Tests must be run in accordance with NFPA 110 Requirements.
Industry Standards
Industry standards related to microgrids (both ANSI and IEC) are evolving rapidly and
can be classified roughly as operating at three different levels.
• Individual DER, or component level: These include microgrid related language
within component-level switchboard or panelboard standards, solar PV inverter
standards.
• System level: These include explicit microgrid system standards relating to
energy management or controls that involve multiple DERs operating as a
synchronized unit. For example, the IEEE 2030.7 standard includes control
functions for microgrid as a system that can manage itself, operate autonomously
or grid connected, and seamlessly connect to and disconnect from the main
distribution grid.
• Interconnection level: These focus specifically on the interconnection and
interoperability between local microgrids or DERs with utility electric power
systems (EPSs). These standards provide requirements relevant to the
performance, operation, testing, safety considerations, and maintenance of the
interconnection. For example, California Rule 21 has specific requirements on the
types of data to be shared between microgrids or local smart DERs and the utility
energy management system. Architectural details about the interconnection such
as protocols of data interchange (IEEE2030.5), frequency of data updates, are
also specified in California Rule 21.
Microgrid Standards, page 209 classifies the various standards pertinent to the
deployment of microgrids but is not intended to be comprehensive. Many local, state,
and regional jurisdictions may also be relevant.
IEEE 1547.1 IEEE Standard Conformance Test Procedures for The type, production, commissioning, periodic tests, and
Equipment Interconnecting Distributed Energy evaluations that shall be performed to confirm that the
Resources with Electric Power Systems and interconnection and interoperation functions of equipment and
Associated Interfaces systems interconnecting distributed energy resources with the
electric power system confirm to IEEE 1547 are specified here.
IEEE 1547.2 IEEE Application Guide for IEEE Std 1547, IEEE Provides tips, techniques, and common practices to address
Standard for Interconnecting Distributed issues related to DER project implementation.
Resources with Electric Power Systems
IEEE 1547.3 IEEE Guide for Cybersecurity of Distributed Provides guidelines for Cybersecurity of DER’s interconnected with
Energy Resources Interconnected with Electric Electric Power Systems.
Power Systems
IEEE 1547.4 Guide for Design, Operation, Integration, and Provides approaches and good practices for the design, planning,
Interoperability of Intentional Electric Power maintenance, and operation of Intentional Island Systems and their
Systems Islands integration and interoperability with other EPSs.
IEEE 1547.6 IEEE Recommended Practice for Interconnecting Provides an overview of distribution secondary network systems
Distributed Resources with Electric Power design, components, and operation. Describes considerations for
Systems Distribution Secondary Networks interconnecting DR with networks and provides potential solutions
for the interconnection of DR on network distribution systems.
IEEE 1547.7 IEEE Guide for Conducting Distribution Impact This guide provides alternative approaches and good practices for
Studies for Distributed Resource Interconnection engineering studies of the potential impacts of a DR or aggregate
DR interconnected to the electric power distribution system. This
guide describes criteria, scope, and extent for those engineering
studies.
IEEE 1547.9 IEEE Guide for Using 1547 for Interconnection of Addresses interconnection of energy storage distributed energy
Energy Storage Distributed Energy Resources resources to electric power systems. Provides examples of such
with Electric Power Systems interconnection, guidance on prudent and technically sound
approaches to these interconnections.
UL1741, UL1741-SB Standard for Inverters, Converters, Controllers, Describes manufacturing (including software) and product testing
and Interconnection System Equipment for Use requirements to specify inverters more capable of riding through
with Distributed Energy Resources grid excursions and actively managing grid reliability functions.
UL891, UL1558 Standards for Switchboards and Switchgear Supplements ANSI switchgear standards C37.20.1 and C37.51.
Used in conjunction with NFPA70/ NEC.
NFPA99 Healthcare Facilities Code Covers aspects of emergency power systems and associated
testing in healthcare facilities.
UL3001* (evolving) Standard for safety and performance of distributed Covers DER system design, integration, and operation.
energy systems
IEEE 2030.5 IEEE Standard for Smart Energy Profile Defines the application layer with TCP/IP providing functions in the
Application Protocol transport and Internet layers to enable utility management of the
end user energy environment, including demand response, load
control, time of day pricing, management of distributed generation,
and electric vehicles.
IEEE 2030.7 IEEE Standard for the Specification of Microgrid Address functions at the microgrid system level (above the
Controllers component control level) to enable control functions to manage
themselves, operate autonomously or grid-connected and
seamlessly connect/disconnect from the grid.
IEEE 2030.8 IEEE Standard for the Testing of Microgrid Testing procedures to enable verification, performance
Controllers quantification and comparison of different functions of microgrid
controllers.
IEEE 2030.9 IEEE Recommended Practice for the Planning and Best practices for the planning and design, including system
Design of the Microgrid configuration, electrical system design, safety, power quality
monitoring and control, electric energy measurement and scheme
evaluation.
California Rule 21 Tariff document describing the interconnection, Rules for the performance, function, metering, and
operating, and metering requirements for communications of generation and energy storage systems.
generation facilities to be connected to a utility’s
distribution system.
IEEE P2030.11 DER Management Systems Functional Guides the development of functional specifications for DER
(Project started) Specification management systems. It includes guiding principles for the
application and deployment of DER management systems.
IEEE P2030.12 Draft Guide for the Design of Microgrid Protection Design and selection of protective devices and the coordination for
(Project started) Systems various modes of microgrid operation, including grid-connected
and islanded modes and related transitions between modes.
Microgrid Functions
Typical microgrid functions can be classified into two main categories – microgrid
operation, and microgrid optimization - as described below.
Microgrid Operation
• Monitoring: Many microgrids require monitoring from remote network operation
centers (NOCs). Dependable monitoring of microgrids require the measurement
and display of energy, power, and other metrics for individual DERs and loads. If
microgrid sites have on-site operators, factor into the design a local HMI to
display microgrid loads, generation and status information.
• Alarming and notification: Many microgrids require monitoring from remote
network operation centers (NOCs). Dependable monitoring of microgrids require
the measurement and display of energy, power, and other metrics for individual
DERs and loads. If microgrid sites have on-site operators, factor into the design a
local HMI to display microgrid loads, generation and status information.
Alerting on abnormal operating conditions or malfunctions of microgrid
components is an essential component of microgrid design.
• Export management: In some microgrid deployments, utilities may prohibit or
limit export of active power to the grid. In these cases, use excess PV production
to charge BESS or curtailed to minimize exceeding export thresholds. In addition
to the basic control-limiting functions, export management can be extended to
include optimization functions. For example, BESS may be preemptively
discharged in preparation to absorb the expected excess PV based on weather
forecasts.
• Grid connection management: When the microgrid is islanded (off-grid mode),
many islanding sequences of operation must be safely managed. Balancing
various generation sources optimally while islanded is another control function.
For example, PV production and BESS charging or discharging may need to be
orchestrated precisely to avoid imbalanced conditions. Conversely, when the
utility grid is restored, the restoration sequences may need to be tuned to safely
manage state transitions. These grid connection and disconnection sequences
must be designed into the microgrid control algorithms.
• Load management: Both when islanded or when connected to utility grids, loads
must be monitored and either shed or reconnected automatically based on the
generation-consumption balance. For example, lower priority non-critical loads
may be disconnected when islanded and only restored last after utility restoration.
Even when connected to the utility grid, some local loads may need to be
disconnected for certain operating conditions. Prioritization and control
capabilities for loads is usually an important design criterion for microgrid
operation.
Microgrid Organization
• Forecasting: Forecasting of both generation and loading is a fast-evolving R&D
area with many applications. For example, weather forecasts can be used
effectively to optimize microgrids. When stormy conditions or natural disasters
like wildfires are forecasted, microgrids can be directed towards charging BESS
in preparation, to improve system resilience during outages.
• Federal, state, and local incentive programs: Federal incentives like the
Investment Tax Credit (ITC) are key enablers of microgrids and apply to all U.S.
geographies, reducing solar PV and BESS capital costs. Additional local incentive
programs at the state, local or municipal level may further subsidize the upfront
expense of onsite generation. For example, state-wide incentives like the Self
Generating Incentive Program (SGIP) in California and the DEEP Microgrid Grant
Program in Connecticut have proven to be highly effective mechanisms for
deploying more distributed, clean, and resilient energy systems. Successful
program references and an uptick of large policy measures such as the 2021
Federal Infrastructure Bill have prompted more state and local emphasis and
funding for microgrids. The commercial and technical requirements that are
applicable to take advantage of these incentives must be evaluated and applied
for on a project-by-project basis.
• Market participation: Additional programs applicable to specific regions may
also be available, enabling microgrids to offer unique benefits to the local Utility or
Regional Transmission Organizations (RTOs) via participation in market
programs like Demand Response or Frequency Regulation. This type of market
participation is highly location-specific and is typically oriented around specific
outcomes that DER technology or microgrid operational mode can drive.
All the above geography-specific factors impact the design, cost, and ultimately the
functionality of a microgrid system. These factors are evaluated carefully during initial
project development. With increased emphasis on dependable, efficient, and
sustainable energy systems, the local market landscape for microgrids is a fast-
evolving one. Staying informed on geographic-specific conditions remains an
important factor in implementing successful microgrid solutions.
Space and footprint considerations
Many physical space and footprint considerations impact the feasibility and design of
microgrids. As previously mentioned, site location is a critical factor that helps to
identify the size of DERs that can be installed and the potential locations for these
resources. Once site location has been selected, by analyzing the electrical drawings
and infrastructure maps, the available physical space can be determined for the
required electrical distribution, microgrid controls infrastructure, and the desired DERs
(such as solar PV, BESS, and generator sets). It is important to discuss these physical
space considerations with the site decision makers early in the design phase.
The space footprint of the electrical distribution equipment and controls cabinets is a
key consideration in microgrid design. Some of the design parameters impact the
space footprint are the main bus amperage rating, the number of sections needed to
fit all the required breakers and devices, and whether the electrical distribution
equipment is going to be located indoors or outdoors. Physical space is also a key
factor when determining potential options for the size, type and mix of DERs (solar
PV, BESS, generator sets) to be installed on site.
For solar PV systems, the type of available space (rooftop, land, or parking lot space)
is an important criterion. This dictates whether rooftop PV, ground mount PV, carport
PV, or some combination of the three can be designed into the system. Once PV type
is determined, the amount of space dictates the size of the solar system that can be
installed. Typically, physical space is a restricting factor with PV - meaning the system
capacity that can fit within the available space is often insufficient to meet site
demand. The design challenge is that of maximizing the size of the PV system based
on the available space. A rough estimate of space for one MW of ground-mount PV is
about four acres. Non-ground mount systems have additional considerations. Rooftop
PV systems must factor in existing rooftop equipment such as HVAC and sprinklers
and incorporate row spacing and rooftop setbacks for maintenance and safety.
Carport systems must factor in the parking lot set up and incorporate fire lane widths.
Additionally, local zoning and safety codes also impact the placement of PV systems.
For a BESS, the physical space requirements and performance varies widely
depending on the type of system and battery chemistry selected. A common battery
chemistry integrated into microgrid systems is Lithium-Ion Iron Phosphate (LFP).
Depending on the specific BESS, the power blocks can vary in size which impacts the
overall power and energy density of a system. Therefore, the physical space for a
BESS is dependent upon the power and energy capacity needed. For example, a
typical 300 kW/745 kWh BESS has an energy density of 12.4 kWh/sq. ft. and a
system power density of 5 kW/sq. ft. A typical footprint for this BESS is 90 inches H x
52 inches D x 155 inches L. In addition to required system size of a BESS, the depth
increases if the system needs to be outdoor rated. For other battery chemistries,
footprint and space design constraints vary widely.
Physical space requirements for generator sets vary depending on the fuel and
particular system selected. Two common generator fuel types are natural gas and
diesel. For example, a two MW Natural Gas Generator set has dimensions of 291
inches L x 84 inches W x 95 inches H, or a power density of 11.8 kW/sq ft.
Comparatively, an example two MW Diesel Generator Set maximum dimensions are
404 inches L x 99.62 inches W x 156.6 inches” H, or a power density of 10.6 kW/sq ft.
On-site fuel storage must be factored into the design. A typical fuel oil storage tank
with a capacity of 1000 gallons has a diameter of 48 inches and a length of 130
inches. There may be also local safety requirements for on-site fuel storage.
Additionally, certain mission-critical sites (like hospitals) may have minimum on-site
fuel storage capacities specified in corporate guidelines and safety codes.
In addition to PV, Lithium-Ion BESS and generators, there are many other
technologies that can be implemented for energy generation and storage in a
microgrid system. A few of these other technologies include fuel cell, flow battery,
Uninterruptable Power Supply (UPS), and wind turbines. Each of these technologies
vary widely in footprint and density, and physical space availability need to be
considered when designing the microgrid system.
Building Information Modeling (BIM) and other automated layout software tools are
typically used during the early design phase of microgrids to optimize the asset
locations and positions appropriately.
Economic analysis and project return-on-investment (ROI) tools
One of the main influences driving the adoption of Behind-The-Meter (BTM) microgrid
systems in Commercial and Industrial (C&I) facilities is lowering the net cost of energy
they purchase from utilities. Through the appropriate use of on-site Distributed Energy
Resources (DERs) to at least partially offset consumption and optimizing the
remaining utility usage, microgrid can significantly lower facility bills. Several key
components of data are needed during the early project assessment phase to
complete a complete economic analysis to estimate the potential return-on-
investment (ROI) of microgrid projects.
First, the site geographical location is very critical, and helps identify the size of DERs
that can be installed, their types, and potential locations. Estimates of the solar
irradiance (power per unit area of energy from the Sun) at a given location helps
determine the potential for solar PV deployment. As labor rates vary regionally,
location also helps determine the expected installation and operations and
maintenance (O&M) costs, which impacts ROI calculations.
Next, a comprehensive analysis of past (or projected, in the case of greenfield
projects) utility bills are needed to understand site energy usage and current tariffs. A
minimum of twelve months of utility bills is typically needed to establish a site load
baseline. Additional years of data helps minimize projection errors due to any year-to-
year variability. Using utility bill kWh rates, the impact of kWh’s delivered from on-site
DERs can be financially quantified. Time-of-use tariffs with on-peak, mid-peak and off-
peak rates can also be factored into the analysis. Other energy management
strategies such as demand-shifting, demand compensation, power factor penalty
compensation, can also be analyzed within the economic analysis.
Fifteen-minute interval data helps to size dispatchable DERs by modeling the daily
load profile of the site, rather than monthly averages. For example, a Battery Energy
Storage System (BESS) can be sized to avoid costly demand peaks and reduce
energy usage during expensive on-peak periods. If a Combined Heat and Power
(CHP) system is also within scope, usage data from the gas utility can help with
proper sizing and operation relative to thermal load. Generally, CHP tends to be
attractive when there are high and coincident peaks in electrical and thermal demand,
and when the price of electricity ($/kWh) is roughly three times or more than the unit
price of natural gas ($/therm). This unit price disparity is commonly referred to as
“Spark Spread.”
Fortunately, there are many open-source and commercial software tools available to
optimize microgrid deployment and provide recommendations for DER types and
sizing. These tools typically require the input data discussed above (such as site
location, utility tariff structures, load profile, installation costs). In addition to DER
sizing, these software tools may also produce a Discounted Cash Flow (DCF) model
to simulate incentive programs like the Investment Tax Credit (ITC), Modified
Accelerated Cost Recovery System (MACRS) for accelerated depreciation, and any
local or regional incentives. This DCF model outputs important metrics like the Net
Present Value (NPV) and payback period, which are critical values in determining if
the project is a worthy investment. Running these ROI calculations is typically an
iterative process – making best-guess estimates, examining the outputs, adjusting the
inputs, and repeating the exercise. By leveraging these analysis tools, iterative
changes to design and cost inputs can be run quickly and reliably, accelerating the
critical timeline between project development, and securing financing.
Protection considerations
Requirements for the system topology are designed to increase both the reliability of
the overall utility system and with the reliability of service to the installation in question.
These requirements typically take the following forms:
• Restrictions on the size of services.
• Restrictions on, or requirements for, normal and alternate services and transfer
equipment between the two.
• Restrictions or requirements for the configuration of emergency and standby
power systems.
• Restrictions on the types of service disconnecting devices allowed.
• Restrictions on the types of service overcurrent protection allowed.
• Requirements for service cable compartments in service equipment.
• Requirements or restrictions on the number and types of protective relaying.
• Overall requirements for the service switchgear.
The requirement that is applied to virtually every utility installation, is that the service
overcurrent device must coordinate with the upstream utility overcurrent device,
typically a recloser or utility substation circuit breaker. If there is standby power on the
premises, the utility typically requires that paralleling the alternate power source with
the utility source not be possible unless stipulated in the rate agreement for the
service in question.
Requirements for restricted access to service cable termination and service
disconnect compartments in the service switchgear are another common. In some
cases, these must be in a dedicated switchgear or switchboard section, increasing the
service equipment footprint. In many cases grounding means must be provided with
the equipment to allow the utility’s preferred safety grounding equipment to be
installed. In some cases, requirements may be imposed on the entire service
switchgear, such as electrical racking for circuit breakers or barriers that are not
standard for the equipment type used.
In some cases, the control power for the service switchgear, such as a battery, must
be designed to the utility’s specifications. Additional protective relaying may be
required to minimize abnormal conditions which, although not harmful to the system
being served, affect the reliability of the utility system. In some cases, the makes and
models of protective relays for the service overcurrent protection are restricted to
those the utility has approved.
Metering considerations
Measurement of load and system performance are critical to the functionality of the
microgrid, both while grid connected and while islanded from the grid. This section
does not address specific meter hardware but does address the information that
meters may need to collect, and their interaction with other devices, including edge
control systems and local or remote energy management systems.
Metering considerations must cover both functional and financial operation of
microgrids. Grid-connected metering and grid-independent metering requirements
must reflect the nature of the overall system, including possibly divided responsibilities
between owner and operator. Some microgrid systems are components within larger
power distribution systems, while others are self-contained, so the architecture of the
system influences the metering choices and options.
The following measurements are typical for most microgrid systems:
• Metering variables like energy, voltage, frequency, active power, reactive power.
at the Point of Common Coupling (POCC) that is, interconnection point with utility.
• State/status for each DER.
• Instantaneous and historical load profiles for the full system as well as secondary
and tertiary loads.
Functional considerations for grid-connected operations typically includes metering
and monitoring the grid reference, and individual DERs. The financial opportunities
and constraints of the system dictate the location and type of metering. For example,
revenue grade metering may be required for PV and BESS systems to comply with
Investment Tax Credit verification, utility billing/credit for renewable energy delivered
to the customer and the grid, and for ancillary services performance, such as
participation in demand response or frequency response programs. Measurement of
loads typically does not require revenue grade metering unless there is a specific
need for sub-billing or tenant metering functions.
Metering sample rates and the volume of historical data collected from metering
(through various devices such as dedicated meters, LV trip units, relays, or revenue
grade meters) influence the performance of the edge layer and analytics layer
software. There is an implicit tradeoff between volume of data captured and the
performance of these software packages. Optimization software processes multiple
streams of data such as BESS state of charge, metered PV production, CHP
utilization, metered grid data, to output control signals that meets the system and
customer goals. Production versus consumption decisions utilize meter data that is
optimal for the algorithm being processed. Typically, cloud optimization software
samples averaged metering values to send DER optimization signals to the site
microgrid.
To meet resilience goals, accurately monitoring the state of the grid reference is
crucial. Edge layer software monitors the POCC, establish and stabilize local grid
forming resources, and execute load management (shed/add) decisions based on the
collected metering data. The methodology and decision making related to sequence
of operations related to grid transitions relies on appropriate metering of loads and
sources.
Many sites and customers are likely not be mindful of the importance of power quality
in microgrids; however, this is an important design criterion. Initial microgrid
assessments and power system studies dictate the necessity to include power quality
metering or even power quality correction equipment. Appropriate power quality
metering help maintain system uptime (resilience) through awareness of changes of
DER or other system component performance prior to system malfunction.
Load management
A traditional model of backup power involves transferring full or partial site load to a
fixed diesel or natural gas generator, using an automatic transfer switch (ATS). The
ATS both senses the loss in utility voltage and signals the generator to engage.
Standard Loads
ATS
UPS
Essential Loads
Critical Loads
In the configuration above, there is typically no way to control individual loads. The
source simply supports loads on its own circuit and loading that exceeds source
capacity simply trips the source offline.
In contrast, in a behind-the-meter microgrid, multiple sources (Solar PV, Battery
Storage, Combined Heat and Power, Fuel Cells, Generators to name common assets)
may be combined in parallel, where the microgrid controller manages loading and
power stability between sources. In addition to active source management, microgrids
may also manage site loads directly and indirectly. The ability to manage both sources
and loads introduces a range of economic and functional project design options.
UPS
Essential Loads Standard Loads
Critical Loads
into the critical loads tier to help with adequate UPS coverage. Microgrid control code
must be designed to provide all the sequences of operations associated with
islanding, restoration, load and source management.
Microgrids must be carefully designed with load characteristics, as well as generation
capacity in mind to provide a solution that best accomplishes the site’s specific
response time needs and requirements.
104. IEEE Standard for Interconnection and Interoperability of Distributed Energy Resources with Associated Electric Power Systems Interfaces.
Revision of IEEE Std 1547-2003.
105. March 2018 (Version 2.1).
106. IEEE Standard for Smart Energy Profile Application Protocol.
Capacity Management
Many facilities are in a constant state of flux. Areas are being renovated, equipment is
being moved, new production lines are brought online, old equipment is being
upgraded.
Capacity of the electrical distribution infrastructure must evolve per these changing
environments while not exceeding the rating of electrical distribution equipment.
This is a problem for circuit breakers, UPSs, generators, ATSs, transformers,
capacitor banks, bus bars, conductors, fuses. Often, exceeding the rated capacity
means nuisance trips, but it can also result in overheating or fires.
The Facility Manager needs to:
• Understand the capacity needs of the electrical distribution infrastructure to plan
for expansions or modifications of the facility environment.
• Upgrade the facility while not exceeding the rated capacity of equipment and
mitigating potential risks to the electrical infrastructure (For example, Nuisance
trips, overheating, fires).
This application provides the following for the Facility Manager:
Live Data Display
• Electrical Health diagram
Trends
• Real-time and historical data can be viewed on a trend viewer.
Reports
• Branch Circuit Capacity report
• Generator Power report
• UPS Power report
• Equipment Capacity report
• Generator Capacity report
• Power Losses report
Also, the Joint Commission (also known as JCAHO) requires healthcare facility to test
their emergency power supply system monthly with specific guidelines on the test
procedure as well as data collection and reporting the results.
The Facility Manager needs to:
• Ensure the reliability and availability of backup power supply systems in the event
of unexpected power outages.
• Save time, improve productivity, and ensure accuracy of testing process and
documentation per standards or manufacturer recommendations.
Purpose of Backup Power Testing application is to provide the following:
• Centralized remote operator control and testing:
• Remotely test Automatic Transfer Switch transfer, re-transfer and bypass
• Start, run, stop, and cool down emergency generators
• Remotely control Load Banks for engine-generator and UPS loading
• Conduct system test of emergency generator paralleling
• Remotely test Fire Pump Controls emergency power
• Monitor, automatically record and report backup power tests
• Automatic transfer switches
• Back-up generators
• UPS
• Emergency generator paralleling system
• Load Banks
• Fire Pump Control Systems
• Record key legislated parameters for compliance reports including:
• Transfer time for Automatic Transfer Systems and generators.
• Generator run time, engine loading, exhaust and engine temperature, fuel levels
and battery health.
• UPS’s ability to sustain critical loads during power outage, and UPS battery
health.
• Load bank loading on emergency generator paralleling, engine-generators and
UPS systems.
This application provides the following outputs:
Live Data Display
• Animated device diagrams with status and analog values of ATS, generators,
power control systems, load banks, fire pump controls system and UPS.
Reports
• Generator Test (EPSS) report
• Generator Battery Health report
• Generator Load Summary report
• UPS Auto-test report
• UPS Battery Health report
• Automatic Transfer Switch performance report
• Utility power outage report
“Demand” is the average instantaneous power consumption over a set time interval,
usually 15, 30, or 60 minutes.
Energy
The other major component of an electric bill is energy. The same metering equipment
that measures power demand also records customer energy consumption. Energy
consumption is reported in kilowatt-hours or megawatt-hours. Unlike power demand
with its capacity relationship, customer energy consumption is sometimes related to
fuel requirements in electric utility generating stations. The cost per kilowatt-hour in
each electric utility rate structure, therefore, is often influenced by the mix of
generating plant types in the utility system. Coal, fuel oil, natural gas, hydroelectric,
and nuclear are typical fuel sources on which power generation is based.
Load Factor – Demand/Energy Relationship
One useful parameter to calculate each month is the ratio of the average demand to
the peak demand. This unit-less number is a useful parameter that tracks the
effectiveness of demand management techniques. A load factor of 100% means that
the facility operated at the same demand the entire month, a so-called “flat” profile.
This type of usage results in the lowest unit cost of electricity.
Few facilities operate at a load factor of 100%, and that is not likely to represent an
economical goal for most facilities. But a facility can calculate its historical load factor
and seek to improve it by reducing usage at peak times, moving batch processes to
times of lower demand, and so forth. Load factor can be calculated from values
reported on practically every electric bill:
LF = kWh / (kW * days * 24);
Where LF is Load Factor, kWh is the total energy consumption for the billing period,
kW is the peak demand set during the billing period, and days is the number of billing
days in the month (typically 28-32). “24”, of course is the number of hours in a day.
Time-of-Use customers may prefer to track load factor only during on-peak time
periods. In that case, the kWh, kW, days, and hours/day in the formula are changed to
reflect the parameters established only during the on-peak periods.
Typical load factor for an industrial facility depends to a great degree on the number of
shifts the plant operates. One shift, five-day operations typical record a load factor of
20-30%, while two-shifts yield 40-50%, and three shift, 24/7 facilities may reach load
factors of 70-90%.
Equal Energy
Unequal Demand
Demand, kW
Demand, kW
Load Factors, page 228 gives a graphical comparison of facilities with dramatically
different load factors. The three-shift facility produces an average demand that is
nearly equal to its peak demand, while the average and peak demand for the one shift
facility is much less than one.
Power Factor
The relationship of real, reactive, and total power has been introduced previously, and
described as the “power triangle”. For effective electricity cost reduction, it is important
to understand how the customer’s electric utility recoups its costs associated with
reactive power requirements of its system. Many utilities include power factor billing
provisions in rate schedules, either directly in the form of penalties, or indirectly in the
form of real-power billing demand that is higher than the actual measured peak.
Even if a utility does not charge directly for poor power factor, there are at least three
other reasons that a customer may find it economical to install equipment to improve
power factor within its facility, thereby reducing the reactive power requirements of the
utility:
• Reduce power factor penalties.
• Release capacity of an existing circuit.
• Reduce heating losses associated with power distribution (often called I2R
losses).
• Improve voltage regulation.
Fixed capacitor banks are best suited for use on electrical systems with no voltage or
current harmonics. In the presence of harmonics, Active Harmonic Filters (AHF) are a
better solution.
Production
Lighting Systems Equipment
8% 8%
Compressed Air
8% Packaging Lines
8%
Utility Systems
3%
HVAC
Miscellaneous 10%
6%
Cooling Tower
Fans
3%
Chilled Water
Pumps
9%
Condenser Water
Chillers Pumps
33% 3%
The FEP is best developed using actual power measurements from existing facility-
wide monitoring systems. Some types of loads, lighting, for instance, may comprise
part of the usage of every major circuit in the facility. This fact would suggest that the
meter measuring the power consumption of a feeder serving the building’s centrifugal
water chillers.
Actual power monitoring data from existing Circuit Monitors measuring the power
consumption of individual feeders is the best basis for establishing the Facility Energy
Profile.
profile appears; and what times of day these peaks are occurring. Armed with this
information, the energy auditor can better evaluate the potential for a variety of
demand reduction techniques.
The Demand Sort Table, page 231 is produced by rearranging individual integrated
demand readings for a given billing period. Meters record demand readings
chronologically, 3000 or so readings for a 30-day billing period at 15-minute demand
intervals; the demand sort utilizes a software tool to distribute the readings from
highest to lowest, so that times and values of peak usage are easily analyzed.
The Demand Sort Table, page 231 facilitates demand analysis by depicting the
number of intervals (or hours) during which the plant’s peak electrical demand
exceeded certain levels.
Table 49 - Demand Sort Table
2400 2400
2350 2350
2300 1 1 2300
2250 2 5 2 6 2250
2200 7 31 15 25 6 2200
2150 15 73 53 92 21 2150
Using the Demand Reduction Table, page 232, the engineer can determine that a
reduction in peak demand to 2200 kW at this example facility would have required a
demand reduction of 122 kW for 25 fifteen-minute intervals, or 6.25 hours, in August
of the sample year.
12,400.00
Peak-day Load Profiles, page 232 from actual power monitoring data can show
consistency, or, as in this case, a single-day aberration in peak demand that set the
demand minimum billing level (ratchet) for the remainder of the year.
Demand Control
Demand controls systems are available that perform these basic functions:
• Measure power consumption (demand) in real time.
• Predict demand level based on rate of instantaneous usage.
• Compare predicted value to target setpoint.
• Transmit signals to pre-determined equipment to turn off or curtail power usage if
demand is predicted to exceed target kW.
These demand controls systems are intended to reduce peak demand for a facility to
some predetermined level.
The design engineer’s foremost demand control system challenge is to identify loads
in the facility that can be controlled effectively. Ideal load candidates include those
machines or processes that are (1) currently contributing to the facility’s load at peak
times, and (2) whose function can be delayed or curtailed at times of peak.
Most facilities lack equipment or processes that fit this ideal description, despite the
numerous machines and processes that may be operating at peak times. In fact,
successful demand control is usually the exception rather than the rule.
One common candidate for the demand control system is the air conditioning system.
Buildings equipped with multiple packaged direct-expansion air conditioning systems
are typical targets of demand control sales efforts. Unfortunately, demand control of
air conditioning compressors usually leads to loss of temperature or humidity control
within the conditioned space, or lack of demand savings.
The reason for this paradox is twofold. One, natural diversity among multiple air
conditioning compressors maximizes the chance that all compressors are not
operating at full load at the same time. Strangely, this fact is often highlighted in the
demand control system sales pitch: “Not all compressors are running at the same
time, so you should turn some off for short periods of time”.
Secondly, basic thermodynamic principles of moist air and vapor-compression
refrigeration systems require compressor power consumption to reduce air
temperature and condense moisture. This process is controlled by thermostats and
humidistats within the facility. When cooling or dehumidification is removed or reduced
at times when these devices are “calling for” them, temperature and humidity rises in
the conditioned space.
So, if not air conditioning equipment, what loads have been successful demand
control candidates? An electrolysis process providing chemicals for a paper mill was
able to reduce peak demand and flatten the demand profile for the overall facility. A
battery-charging system for forklift vehicles in an automotive facility could produce
real demand savings during peak times. Finally, a large induction furnace melting
scrap metal proved to be an effective candidate for the rolling mill at a steel plant.
Chilled Water Supply and Return Temperatures, page 234 increase over the course of
a day due to demand control of inlet guide vanes on a centrifugal water chiller. Space
conditions could not be maintained because of the demand control.
Electricity generation and peak shaving can also be accomplished with steam
cogeneration systems typical of paper mills, refineries, and other large industrial
processes.
Compressed Air
• Provide additional small air compressor for loads.
• Provide outside air intake.
• Eliminate air leaks.