You are on page 1of 42

Power quality and good housekeeping.

(part
1)
Feb 1, 1995 12:00 PM, Waggoner, Ray
0 Comments ShareThisNew
Power quality can be negatively affected by such mundane factors as loose power connections.
When we consider all of the sophisticated apparatus used to improve the operational integrity of
our sensitive equipment and systems, it seems absurd to talk about those common sense, "good
housekeeping" practices.
Nevertheless, they are a valuable part of the power quality equation and require our attention. In
fact, looking after our electrical environment can result in big dividends. Let's find out what's
involved.
What is good housekeeping?
Most solutions to our so-called power quality related events have come from attending to little
things around the shop. In almost every case involving a power quality investigation or site
analysis, the first items that show up in need of correction are those in the area of wiring and
grounding. These include "daisy-chaining" of neutrals and grounds, undersizing of ground wires
(which are expected to carry electrical noise away from sensitive equipment), and even, believe
or not, loose or missing connections.
The NEC and good housekeeping
We often hear a facility manager say, "One day, we're going to bring this facility up to the NEC."
While this is an admirable goal and one that should be achieved, it reflects a major
misconception: the NEC does not represent the peak of wiring and grounding methods and
procedures. Instead, it reflects the minimum requirements needed to protect persons and property
"from hazards arising from the use of electricity." In other words, the Code is a place to start in
the battle for good wiring and grounding of sensitive equipment. But keep in mind, there is no
substitute for, or compromise with, the provisions of the NEC.
In the context of power quality and the NEC, there are conflicting philosophies promoted by
people with differing technical backgrounds and commercial interests. One controversy stems
from the conflicting interpretations of the NEC's grounding requirements. While a safe operating
power system is universally endorsed, some feel that required wiring practices interfere with the
smooth operation of electronic systems. As such, they believe that you may have to install
sensitive electrical and electronic equipment by "skirting" the provisions of the NEC. In fact,
many equipment manufacturers, with warranty forfeiture threatened, direct the end user to do so.

This apparent conflict can only be settled by making safety the overriding concern, which is the
intent of the NEC. If revisions have to be made for the system to operate properly, the equipment
manufacturer must incorporate them in the equipment design, rather than asking for deviations
from the NEC. In fact, when you wire with NEC safety provisions in place, satisfactory
operation will not be sacrificed. The safest system will be the best operating one.
Good housekeeping is not just following the NEC, but includes all of the diligent, sensible
practices taken during installation and maintenance. If more attention is given at the start of
project to the important power quality implications, future problems will be headed off before
they result in shutdowns or damage.
Where to get good housekeeping practices
You can get a feel for these practices by going through the FIPS (Federal Information Processing
Standards) Pub 94, Guideline on Electrical Power for ADP Installations or the IEEE Standard
1100-1992, Recommended Practice for Powering and Grounding Sensitive Electronic Equipment
(the Emerald Book).
The latter publication is especially useful in that it includes recommended design, installation,
and maintenance practices for electrical power and grounding of sensitive electronic processing
equipment used in commercial and industrial applications. These practices include both powerand signal-related noise control techniques.
In many instances, we are confronted with conflicting information and/or confusion in the power
quality area, all stemming primarily from different viewpoints of the same problem. By referring
to this consensus of recommended practices, the confusion and conflicting information is greatly
diminished.
The Emerald Book addresses electronic equipment performance issues while maintaining a safe
installation. A brief description of the power quality problem is usually given, along with
possible solutions and the resources available for assistance in dealing with problems.
Fundamental concepts are reviewed and instrumentation and procedures for conducting site
power analysis are considered. Of special interest are the case histories provided to illustrate
typical problems.
You can also follow case study discussions in a variety of publications. The case histories we
discuss each month here are good examples.
Case history
Many times, we're involved with a consultant, contractor, or facility engineer who has not
thought through the "hows" or "what-ifs" pertaining to wiring, grounding, or protecting a piece
of sensitive equipment or its signal system. Thus, we are called in when this equipment doesn't
work, or when there is no correlation to the errant behavior, or when the damage has already
taken place. Thankfully, this is changing: more planning is being done using the new power
quality guidelines.

One case study example comes to mind as a reminder to check our procedures for installing and
maintaining our equipment, especially power conditioners used on computers or other sensitive
loads.
A large telecommunications facility that manufactured digital electronic devices was
experiencing a rapid increase in requests for UPS systems from operators in its software
engineering area. This was the direct result of one of the workers asking to have a small
electronic equipment system (4 to 6kVA) put on a self-contained battery supported UPS. (At the
time, there were no other UPS needs apparent.) When this request was honored and a 10kVA
UPS was provided for this application, the worker began to experience "clean power" - since this
affected work area was now separated from the wiring problems in the rest of the software
engineering area.
As the word spread among the other personnel in this department, more requests were made for
small UPS systems, so many requests, in fact, that the facility manager had to put a stop to the
separate ordering until it was determined why everyone needed a UPS! The answer from the
workers was, "You're giving us bad power!" It was humorous, in a way, since the manager was
being blamed for all theproblems, even though he had nothing to do with the power company
delivering the energy.
The investigation into this problem led to some very disturbing housekeeping conditions on the
UPS systems themselves.
Site investigation and solution
We were called in to do a site analysis at this facility and, in the course of the visit, we tested the
output of one of these small UPS systems to determine its load characteristics. We were
performing our examination with the load connected and in a nonintrusive fashion, or so we
hoped.
Working at the floor level, we were suddenly covered by a " shadow" - someone was standing
directly over us. It was a computer operator and she wanted to know what we had done to her
computer! It was now "dead in the water", and she was suppose to go home in 30 min. Thanks to
us, however, she now had a 6-hour recovery program to go through to retrieve the lost data!
Needless to say, our credibility was suspect.
We re-checked the output of the UPS. All indicating lights showed good output, the voltage was
correct, but no current was present. Then we went to the respective power outlet and found the
same condition. The obvious theory: something was wrong between the UPS, which was running
well, and the power outlet. The operator maintained, "It's you. Nothing else has happened except
when you tampered with the system!"
We returned to the UPS output to prove that we had not changed anything and that the power
was correct at the output. When we placed the voltmeter leads on the output to show her the
voltage, one of the output wires moved sideways. Looking underneath the connection into the
lug position, we saw that the outgoing wire had never been inserted under the lug. In fact, the lug

had never been opened! The wire had simply been "jammed" up to make contact, and as long as
you didn't touch the arrangement, it worked.
Now the worker was aware of the poor "housekeeping" that had caused the problem. The rest of
the units were checked for the same problem, and better planning for installation and
maintenance was instituted for the future.
In Part 2, we'll look at another example of a problem that began as an assumption about
harmonics but really stemmed from poor wiring.
SUGGESTED READING
EC&M Books:
Practical Guide to Power Distribution Systems for Computers. Practical Guide to Quality Power
for Sensitive Electronic Equipment. For ordering information, call 1-800-654-6776
Standards:
FIPS Pub 94, Guideline on Electrical Power for ADP Installations. For ordering information,
contact the National Technical Information Service, U.S. Dept of Commerce, Springfield, VA
22161.
IEEE Standard 1100-1992, Recommended Practice for Powering and Grounding Sensitive
Electronic Equipment (Emerald Book). For ordering information, call 1-800-678-IEEE.
EC&M Articles:
"New Power Quality Consensus Standard" February '93 issue. "There's More to Power Quality
Than Meets the Eye" October '94 issue. For copies, call 913-967-1801.
RELATED ARTICLE: GOOD HOUSEKEEPING MEANS CHECK EVERYTHING - ASSUME
NOTHING
The IEEE Emerald Book indicates that problems in industrial and commercial premises wiring
and grounding account for a large share of all reported power quality problems. The greatest
number of these is in the feeders and branch circuits serving the critical loads.
As such, the first activity in checking for power problems is to survey the soundness of the AC
distribution and grounding system supplying the equipment. Problems here include such items as
missing, improper, or poor quality connections in the power wiring and grounding from the
source of power to the load. These problems can be generally classified as mechanical in nature.
Through error or oversight, intentional or unintentional, the power distribution and grounding
system is not installed per national, state, or local electrical codes and other specifications. For
example, the NEC only permits a neutral-to-ground bond at the source of power; that is, at the

main panel or isolation transformer secondary. Yet, this improper connection is a common site in
the field and is the source of many of the common power quality problems.
In new installations, connections may be left off or not properly tightened. Reversal of
conductors also can occur. Connections may become loose because of equipment vibration. Even
the ON-OFF cycling of a connected load, with its creation of cycled heating and cooling, can
eventually result in high-impedance connections. Also, periodic additions or modifications to a
distribution system may result in missing, improper, or poor quality connections.

Choosing grounding options for electrical


power systems.
Feb 1, 1995 12:00 PM, Bridger, Baldwin, Jr.
0 Comments ShareThis2
Your choice of either solid, low-, or high-resistance grounding is based on the power system
application and degree of power interruption tolerated.
Solid, low-, or high-resistance grounding? That is the question asked by those involved in the
design or retrofit of power systems. The answer depends upon certain important factors. To make
the correct choice, the designer must have a complete understanding of system configurations,
favorable performance traits, and drawbacks. Also required is the relative importance of the
powered process or load.
The table shown on the facing page includes a comparison of the characteristics of these various
grounding methods. Let's look at them more closely.
Historical information
Most older industrial plants were powered by ungrounded, 3-phase, 3-wire, delta power systems.
Many of these systems are still in use today. This system choice was based on two factors. First,
it made the most efficient use of conductor copper. Second, no fault current flowed when the first
ground-fault occurred, which was, and still is, considered an advantage in some applications,
although a shock hazard is introduced.
However, multiple motor failures in numerous industrial plants were seen and were due to severe
overvoltages caused by arcing or resonant ground faults on the ungrounded systems. To prevent
these overvoltages, many power system neutrals were grounded, usually solidly. There were
many factors that contributed to the change to solidly grounded systems and these factors are still
important today.

First, solid grounding very effectively limits the maximum phase-to-ground voltage. Second, it
allows phase-to-neutral loads to be served without encountering dangerous neutral-to-ground
voltages under ground-fault conditions. Third, simple and effective ground relaying systems can
be used to isolate the defective portion of the system under ground-fault conditions.
Solid grounding limitations
There are, however, some limitations to solid grounding. In medium voltage (MV) systems
(2400V through 35kV), even with good ground-fault relaying, the damage at the point of fault
can be excessive. In fact, this problem led to the common use of low-resistance grounding, which
allows the passing of anywhere from several hundred to several thousand amperes of groundfault current. This practice reduces fault damage to acceptable levels while maintaining enough
ground-fault current flow to effectively relay off the defective portion of the system.
In addition, solidly grounded, low voltage (LV) systems in the 480 to 600V range have two other
problems. The first problem stems from application concerns. Some users prefer to maintain
service, if possible, with a ground fault present on the system, or at least to arrange for an
orderly, controlled shutdown. This is especially true for such continuous process industries as
electric power generation, oil refining, chemical and steel manufacturing, and the paper industry.
Since many of these power systems are worked hot, electricians are exposed to a considerable
flash hazard from a possible line-to-ground fault caused by a misplaced tool.
Second, since most such systems rely on the phase overcurrent devices to protect against ground
faults, it's possible to have a destructive arc of several thousand amperes in magnitude for several
minutes duration without initiating an automatic trip.
To overcome the problems of unwanted shutdown, flash hazard, and burndown while still
maintaining the transient overvoltage protection of a grounded system, high-resistance grounding
was developed.
Pro high-resistance arguments
High-resistance grounding involves the grounding of the system neutral through a resistance that
limits ground-fault current flow to a value equal to or slightly greater than the capacitive
charging current of the system. This value is chosen because it is the lowest level of ground-fault
current flow at which system overvoltages can be effectively limited. Increasing the current flow
improves overvoltage control at the expense of increasing damage at the point of fault;
decreasing the current flow reduces point-of-fault damage at the expense of greater risk of
overvoltage.
High-resistance grounding is applicable to LV and MV power distribution systems serving 3phase, 3-wire loads or line-to-line, single-phase loads. It effectively controls transient
overvoltages during ground faults, minimizes arcing damage and flash hazard at point of fault,
and allows continued operation of the system with a ground fault present at voltages of 5kV and
below.

Components of high-resistance grounding system


A high-resistance grounding system consists of five basic parts: a system neutral, a grounding
resistance, a fault-detector and [TABULAR DATA OMITTED] alarm scheme, a fault-locating
scheme, and packaging for these components. Strictly speaking, only the first two items are
required; however, the grounding system's usefulness is severely limited without the other three
items.
System neutral. By far the easiest way to obtain a system neutral is to use the neutral of a wyeconnected power transformer or generator supplying the system. On any new system, it's
recommended that this method be used.
On existing delta-connected systems (or on new systems that must be delta-connected to allow
paralleling with existing systems), a neutral may be derived by using a bank of three small
transformers connected in wye on the primary and in delta on the secondary. The primary voltage
rating must be equal to the system line-to-line voltage since the transformers connected to the
ungrounded phases will see that voltage under conditions of solid ground fault on one phase. The
secondary should be rated 120V for convenience of fault detection. The kVA rating should be
chosen such that the rated primary current of the transformer equals or exceeds 1/3 of the
selected system ground current, since the ground current divides equally among the three
transformers.
For example, if you decided to ground a 2400V system so that 10A of ground current can flow,
the transformer size required is equal to the quantity 2400 times 10, divided by 3, or 8000VA.
Thus, three standard 10kVA transformers would be used.
Grounding resistance. The grounding resistance determines the value of ground-fault current that
will flow. Since the desired value is dependent on the system capacitive charging current, the
charging current must be determined before the resistor can be selected. The only accurate
method of determining this current for any given system is measurement.
Since measurement is not possible during the design stages of an installation, normal practice is
to estimate the capacitive charging current, provide a tapped resister that allows several settings
in the range of the estimated current, make the necessary measurements, and set the resister at
installation time.
Sufficient data have been accumulated from system measurements to allow fairly accurate
estimates of system capacitive charging currents for various systems.
Typical values of capacitive charging currents have been found to be as follows.
* 480V systems: usually less than 1A with maximum of about 5A.
* 2400V and 4160V systems: 2 to 7A.
* 13.8kV systems: 10 to 20A.

These values are for in-plant power systems, such as auxiliary systems for generating systems or
distribution systems for industrial plants. Utility distribution systems would exhibit higher values
because of the greater length of conductor involved.
(July 1994 issue, "What To Know About High-Resistance Grounding.")
Determining required resistance
After the system charging current has been estimated and a value of ground-fault current
selected, the value of the required resistance is determined. For 480V systems, a very practical
grounding resistor can be made from four 77 ohm resistors rated 750W, 240V each. These can be
connected in various series-parallel arrangements to produce the appropriate current flow.
LV delta-connected systems. It's more common to connect a grounding transformer bank to the
480V secondary and insert the resistance between the neutral of this bank and ground, as shown
in Fig. 1, with no load connected to the secondary. As shown, ground-fault current can be limited
to 1.2A by a resistance of 277 divided by 1.2, or 230.8 ohms. Thus, three 77 ohm resistors
connected in series will provide this value.
MV delta-connected systems. For MV systems having a power transformer with a deltaconnected secondary, a grounding transformer bank is connected to the power transformer
secondary, and the grounding resistance is connected in the secondary of this bank, as shown in
Fig. 2. This permits the fault-detecting and locating circuitry components to be operated at the
secondary voltage level. With this connection, the secondary current may be calculated by
multiplying the transformer primary current by the transformer ratio. This is the current through
the grounding resistor, and its value establishes the continuous current rating of the grounding
resistor. The voltage across the resistor under ground-fault conditions is 1.732 times the
secondary voltage of the grounding transformer bank, or 208V for a 120V rating. The grounding
resistance required can be determined from these values of current and voltage. For the example
shown in Fig. 2, the transformation ratio is 4160 divided by 120, or 34.67 to 1. For a fault current
of 5A, primary current will be 5 divided by 3, or 1.67A; secondary current will be 1.67 times
34.67, or 57.9A. The required grounding resistance will be 208 divided by 57.9, or 3.6 ohms.
This will be seen by the fault current as a high resistance when reflected in the primary.
MV wye-connected systems. For a power transformer with a wye-connected secondary, the
primary of a single-phase grounding transformer is connected between the neutral point and
ground, and the resistor is connected in the secondary circuit, as shown in Fig. 3. The primary
voltage rating of the transformer must be at least equal to the line-to-neutral system voltage and
may be equal to the line-to-line system voltage, if that is more convenient. The kVA rating must
be chosen so that the rated primary current of the transformer is not exceeded by the system
ground-fault current. The secondary voltage rating may be either 120 or 240V. The secondary
current under ground-fault conditions will be the system ground-fault current multiplied by the
transformer ratio. The secondary voltage under ground-fault conditions will be the system lineto-neutral voltage divided by the transformer ratio. Using these values, the resistance and
wattage of the ground resistor can be calculated. The values shown in Fig. 3 are the results for a

ground-fault current of 5A. Note that the ohmic value is different from that of Fig. 2, but the
wattage required is the same.
As in LV systems, it's common practice for manufacturers to supply a tapped resistor that covers
the range of expected values. Field measurements will determine the final setting.
High-resistance grounding do's and don'ts
The following guidelines should be followed when using a high-resistance grounding system.
* Use high-resistance grounding to limit transient overvoltages without shutting down grounded
equipment on occurrence of first ground fault (5kV and below).
* Use sensitive ground-fault relays to trip breakers feeding faulted system elements at voltages
above 5kV.
* Enforce maintenance procedures for locating and removing ground faults promptly upon
detection.
* Test all systems for actual system capacitive charging current upon installation and set
grounding resistor accordingly.
* Do not use high-resistance grounding where 3-phase, 4-wire loads must be served.
* Do not use high-resistance grounding as a substitute for proper system maintenance.
* Do not provide additional ground connections on other electrical equipment when using highresistance grounding equipment. Ground only at grounding resistor.
SUGGESTED READING
EC&M Articles:
"What To Know About High-Resistance Grounding" July '94 issue.
Baldwin Bridger, Jr., P.E. is Technical Director, Powell Electrical Manufacturing Co., Houston,
Tex. and past president of the IEEE Industry Applications Society.

What to know about protective relays.


Feb 1, 1995 12:00 PM
1 Comment ShareThis20

The successful operation of an MV distribution system depends on the proper selection and
setting of switchgear relays.
Protective relays are arguably the least understood component of medium voltage (MV) circuit
protection. In fact, somebelieve that MV circuit breakers operate by themselves, without direct
initiation by protective relays. Others think that the operation and coordination of protective
relays is much too complicated to understand. Let's get into the details and eliminate these
misbeliefs.
Background information
The IEEE Standard Dictionary defines a circuit breaker as follows.
A device designed to open and close a circuit by nonautomatic means, and to open the circuit
automatically on a predetermined overload of current without injury to itself when properly
applied within its rating.
By this definition, MV breakers are not true circuit breakers, since they do not open
automatically on overcurrent. They are electrically operated power-switching devices, not
operating until directed by some external device to open or close. This is true whether the unit is
an air, oil, vacuum, or [SF.sub.6] circuit breaker. Sensors and relays are used to detect the
overcurrent or other abnormal or unacceptable condition and to signal the switching mechanism
to operate. The MV circuit breakers are the brute-force switches while the sensors and relays are
the brains that direct their functioning.
The sensors can be current transformers (CTs), potential transformers (PTs), temperature or
pressure instruments, float switches, tachometers, or any device or combination of devices that
will respond to the condition or event being monitored. In switchgear application, the most
common sensors are CTs to measure current and PTs to measure voltage. The relays measure
sensor output and cause the breaker to operate to protect the system when preset limits are
exceeded, hence the name "protective relays." The availability of a variety of sensors, relays, and
circuit breakers permits the design of complete protection systems as simple or as complex as
necessary, desirable, and economically feasible.
Electromechanical relays
For many years, protective relays have been electromechanical devices, built like fine watches,
with great precision and often with jeweled bearings. They have earned a well-deserved
reputation for accuracy, dependability, and reliability. There are two basic types of operating
mechanisms: the electromagnetic-attraction relay and the electromagnetic-induction relay.
Magnetic attraction relays. Magnetic-attraction relays, as shown in Fig. 1, have either a solenoid
that pulls in a plunger, or one or more electromagnets that attract a hinged armature. When the
magnetic force is sufficient to overcome the restraining spring, the movable element begins to
travel, and continues until the contact(s) close or the magnetic force is removed. The pickup

point is the current or voltage at which the plunger or armature begins to move and, in a
switchgear relay, the pickup value can be set very precisely.
These relays are usually instantaneous in action, with no intentional time delay, closing as soon
after pickup as the mechanical motion permits. Time delay can be added to this type of relay by
means of a bellows, dashpot, or a clockwork escapement mechanism. However, timing accuracy
is considerably less precise than that of induction-type relays, and these relays are seldom used
with time delay in switchgear applications.
Attraction-type relays can operate with either AC or DC on the coils; therefore, relays using this
principle are affected by the DC component of an asymmetrical fault and must be set to allow for
this.
Induction relays. Induction relays, as shown in Fig. 2, are available in many variations to provide
accurate pickup and time-current responses for a wide range of simple or complex system
conditions. Induction relays are basically induction motors. The moving element, or rotor, is
usually a metal disk, although it sometimes may be a metal cylinder or cup. The stator is one or
more electromagnets with current or potential coils that induce currents in the disk, causing it to
rotate. The disk motion is restrained by a spring until the rotational forces are sufficient to turn
the disk and bring its moving contact against the stationary contact, thus closing the circuit the
relay is controlling. The greater the fault being sensed, the greater the current in the coils, and the
faster the disk rotates.
A calibrated adjustment, called the time dial, sets the spacing between the moving and stationary
contacts to vary the operating time of the relay from fast (contacts only slightly open) to slow
(contacts nearly a full disk revolution apart). Reset action begins when the rotational force is
removed, either by closing the relay contact that trips a breaker or by otherwise removing the
malfunction that the relay is sensing. The restraining spring resets the disk to its original position.
The time required to reset depends on the type of relay and the time-dial setting (contact
spacing).
With multiple magnetic coils, several conditions of voltage and current can be sensed
simultaneously. Their signals can be additive or subtractive in actuating the disk. For example, a
current-differential relay has two current coils with opposing action. If the two currents are
equal, regardless of magnitude, the disk does not move. If the difference between the two
currents exceeds the pickup setting, the disk rotates slowly for a small difference and faster for a
greater difference. The relay contacts close when the difference continues for the length of time
determined by the relay characteristics and settings. Using multiple coils, directional relays can
sense direction of current or power flow, as well as magnitude. Since the movement of the disk is
created by induced magnetic fields from AC magnets, induction relays are almost completely
unresponsive to the DC component of an asymmetrical fault.
Most switchgear-type relays are enclosed in a semiflush-mounting drawout case. Relays usually
are installed on the door of the switchgear cubicle. Sensor and control wiring are brought to
connections on the case. The relay is inserted into the case and connected by means of small
switches or abridging plug, depending on the manufacturer. It can be disconnected and

withdrawn from the case without disturbing the wiring. When the relay is disconnected, the CT
connections in the case are automatically shorted to short circuit the CT secondary winding and
protect the CT from overvoltages and damage.
Many relays are equipped with a connection for a test cable. This permits using a test set to
check the relay calibration. The front cover of the relay is transparent, can be removed for access
to the mechanism, and has provisions for wire and lead seals to prevent tampering by
unauthorized personnel.
Solid-state relays
Recently, solid-state electronic relays have become more popular. These relays can perform all
the functions that can be performed by electromechanical relays and, because of the versatility of
electronic circuitry and microprocessors, can provide many functions not previously available. In
general, solid-state relays are smaller and more compact than their mechanical equivalents. For
example, a 3-phase solid-state overcurrent relay can be used in place of three single-phase
mechanical overcurrent relays, yet is smaller than one of them.
The precision of electronic relays is greater than that of mechanical relays, allowing closer
system coordination. In addition, because there is no mechanical motion and the electronic
circuitry is very stable, they retain their calibration accuracy for a long time. Reset times can be
extremely short if desired because there is no mechanical motion.
Electronic relays require less power to operate than their mechanical equivalents, producing a
smaller load burden on the CTs and PTs that supply them. Because solid-state relays have a
minimum of moving parts, they can be made very resistant to seismic forces and are therefore
especially well suited for areas susceptible to earthquake activity.
In their early versions, some solid-state relays were sensitive to the severe electrical environment
of industrial applications. They were prone to failure, especially from high transient voltages
caused by lightning or utility and on-site switching. However, today's relays have been designed
to withstand these transients and other rugged application conditions, and this type of failure has
essentially been eliminated. Solid-state relays have gained a strong and rapidly growing position
in the marketplace as experience proves their accuracy, dependability, versatility, and reliability.
The information that follows applies to electromechanical and solid-state relays, although one
functions mechanically and the other electronically. Significant differences will be pointed out.
Relay types
There are literally hundreds of different types of relays. The catalog of one manufacturer of
electromechanical relays lists 264 relays for switchgear and system protection and control
functions. For complex systems with many voltage levels and interconnections over great
distances, such as utility transmission and distribution, relaying is an art to which some engineers
devote their entire careers. For more simple industrial and commercial distribution, relay

protection can be less elaborate, although proper selection and application are still very
important.
The most commonly used relays and devices are listed here in the Table (see page 42) by their
American National Standards Institute (ANSI) device-function number and description. These
standard numbers are used in one-line and connection diagrams to designate the relays or other
devices, saving space and text.
Where a relay combines two functions, the function numbers for both are shown. The most
frequently used relay is the overcurrent relay, combining both instantaneous and inverse-time
tripping functions. This is designated device 50/51. As another example, device 27/59 would be a
combined undervoltage and overvoltage relay. The complete ANSI standard lists 99 device
numbers, a few of which are reserved for future use.
Relays can be classified by their operating-time characteristics. Instantaneous relays are those
with no intentional time delay. Some can operate in one-half cycle or less; others may take as
long as six cycles. Relays that operate in three cycles or less are called high-speed relays.
Time-delay relays can be definite-time or inverse-time types. Definite-time relays have a preset
time delay that is not dependent on the magnitude of the actuating signal (current, voltage, or
whatever else is being sensed) once the pickup value is exceeded. The actual preset time delay is
usually adjustable.
Inverse-time relays, such as overcurrent or differential relays, have operating times that do
depend on the value of actuating signal. The time delay is long for small signals and becomes
progressively shorter as the value of the signal increases. The operating time is inversely
proportional to the magnitude of the event being monitored.
Overcurrent relays
In switchgear application, an overcurrent relay usually is used on each phase of each circuit
breaker and often one additional overcurrent relay is used for ground-fault protection.
Conventional practice is to use one instantaneous short-circuit element and one inverse-time
overcurrent element (ANSI 50/51) for each phase.
In the standard electromechanical relay, both elements for one phase are combined in one relay
case. The instantaneous element is a clapper or solenoid type and the inverse-time element is an
induction-disk type.
In some solid-state relays, three instantaneous and three inverse-time elements can be combined
in a single relay case smaller than that of one induction-disk relay.
Overcurrent relays respond only to current magnitude, not to direction of current flow or to
voltage. Most relays are designed to operate from the output of a standard ratio-type CT, with 5A
secondary current at rated primary current. A solid-state relay needs no additional power supply,
obtaining the power for its electronic circuitry from the output of the CT supplying the relay.

On the instantaneous element, only the pickup point can be set, which is the value of current at
which the instantaneous element will act, with no intentional time delay, to close the trip circuit
of the circuit breaker. The actual time required will decrease slightly as the magnitude of the
current increases, from about 0.02 sec maximum to about 0.006 sec minimum, as seen from the
instantaneous curve in Fig. 3 (see page 47) [ILLUSTRATION OMITTED]. This time will vary
with relays of different ratings or manufacturers and also will vary between electromechanical
and solid-state relays.
Note that this curve is based on multiples of the pickup setting for the instantaneous element,
which is usually considerably higher than the pickup setting of the inverse-time element.
Time delays can be selected over a wide range for almost any conceivable requirement. Timedelay selection starts with the choice of relay. There are three time classifications: standard,
medium, and long time delay. Within each classification, there are three classes of inverse-time
curve slopes: inverse (least steep), very inverse (steeper), and extremely inverse (steepest). The
time classification and curve slopes are characteristic of the relay selected, although for some
solid-state relays these may be adjustable to some degree. For each set of curves determined by
the relay selection, the actual response is adjustable by means of the time dial.
On the inverse-time element, there are two settings. First the pickup point is set. This is the value
of current at which the timing process begins as the disk begins to rotate on an electromechanical
relay or the electronic circuit begins to time out on a solid-state relay.
Next the time-dial setting is selected. This adjusts the time-delay curve between minimum and
maximum curves for the particular relay. Typical inverse, very inverse, and extremely inverse
curves are shown in Fig. 3. A given relay will have only one set of curves, either inverse, very
inverse, or extremely inverse, adjustable through the full time-dial range. Note that the current is
given in multiples of pickup setting.
Each element, instantaneous or time delay, has a flag that indicates when that element has
operated. This flag must be reset manually after relay operation.
Setting the pickup point
The standard overcurrent relay is designed to operate from a ratio-type CT with a standard 5A
secondary output. The output of the standard CT is 5A at the rated nameplate primary current,
and the output is proportional to the primary current over a wide range. For example, a 100/5
ratio CT would have a 5A output when the primary current (the current being sensed and
measured) is 100A. This primary-to-secondary ratio of 20-to-1 is constant so that for a primary
current of 10A, the secondary current would 0.5A; for 20A primary, 1.0A secondary; for 50A
primary, 2.5A secondary; etc. For 1000A primary, the secondary current is 50A, and similarly for
all values of current up to the maximum that the CT will handle before it saturates and becomes
nonlinear.
The first step in setting the relay is selecting the CT so that the pickup can be set for the desired
primary current value. The primary current rating should be such that a primary current of 110 to

125% of the expected maximum load will produce the rated 5A secondary current. The
maximum available primary fault current should not produce more than 100A secondary current
to avoid saturation and excess heating. It may not be possible to fulfill these requirements
exactly, but they are useful guidelines. As a result, some compromise may be necessary.
On the 50/51 overcurrent relay, the time-overcurrent-element (device 51) setting is made by
means of a plug or screw inserted into the proper hole in a receptacle with a number of holes
marked in CT secondary amperes, by an adjustable calibrated lever or by some similar method.
This selects one secondary current tap (the total number of taps depends on the relay) on the
pickup coil. The primary current range of the settings is determined by the ratio of the CT
selected.
For example, assume that the CT has a ratio of 50/5A. Typical taps will be 4, 5, 6, 7, 8, 10, 12,
and 16A. The pickup settings would range from a primary current of 40A (the 4A tap) to 160A
(the 16A tap). If a 60A pickup is desired, the 6A tap is selected. If a pickup of more than 160A or
less than 40A is required, it would be necessary to select a CT with a different ratio or, in some
cases, a different relay with higher or lower tap settings.
Various types of relays are available with pickup coils rated as low as 1.5A and as high as 40A.
Common coil ranges are 0.5 to 2A, for low-current pickup such as ground-fault sensing; 1.5 to
6A medium range; or 4 to 16A, the range usually chosen for overcurrent protection. CTs are
available having a wide range of primary ratings, with standard 5A secondaries or with other
secondary ratings, tapped secondaries, or multiple secondaries.
A usable combination of CT ratio and pickup coil can be found for almost any desired primary
pickup current and relay setting.
The instantaneous trip (device 50) setting is also adjustable. The setting is in pickup amperes,
completely independent of the pickup setting of the inverse-time element or, on some solid-state
relays, in multiples of the inverse-time pickup point. For example, one electromechanical relay is
adjustable from 2 to 48A pickup; a solid-state relay is adjustable from 2 to 12 times the setting of
the inverse-time pickup tap. On most electromechanical relays, the adjusting means is a tap plug
similar to that for the inverse-time element. With the tap plug, it is possible to select a gross
current range. An uncalibrated screw adjustment provides final pickup setting. This requires
using a test set to inject calibration current into the coil if the setting is to be precise. On solidstate relays, the adjustment may be a calibrated switch that can be set with a screwdriver.
Setting the time dial
For any given tap or pickup setting, the relay has a whole family of time-current curves. The
desired curve is selected by rotating a dial or moving a lever. The time dial or lever is calibrated
in arbitrary numbers, between minimum and maximum values, as shown on curves published by
the relay manufacturer. A typical set of time-dial curves for an inverse-time relay is shown in Fig.
4 (see page 48) [ILLUSTRATION OMITTED]. At a time-dial setting of zero, the relay contacts
are closed. As the time dial setting is increased, the contact opening becomes greater, increasing

relay operating time. Settings may be made between calibration points, if desired, and the
applicable curve can be interpolated between the printed curves.
The pickup points and time-dial settings are selected so that the relay can perform its desired
protective function. For an overcurrent relay, the goal is that when a fault occurs on the system,
the relay nearest the fault should operate. The time settings on upstream relays should delay their
operation until the proper overcurrent device has cleared the fault. A selectivity study, plotting
the time-current characteristics of every device in that part of the system being examined, is
required. With the wide selection of relays available and the flexibility of settings for each relay,
selective coordination is possible for most systems.
Selecting and setting other than overcurrent relays are done in similar fashion. Details will vary,
depending on the type of relay, its function in the system, and the relay manufacturer.
Relay operation
An electromechanical relay will pick up and start to close its contacts when the current reaches
the pickup value. At the inverse-time pickup current, the operating forces are very low and
timing accuracy is poor. The relay timing is accurate at about 1.5 times pickup or more, and this
is where the time-current curves start (Fig. 4) [ILLUSTRATION OMITTED]. This fact must be
considered when selecting and setting the relay.
When the relay contacts close, they can bounce, opening slightly and creating an arc that will
burn and erode the contact surfaces. To prevent this, overcurrent relays have an integral auxiliary
relay with a seal-in contact in parallel with the timing relay contacts that closes immediately
when the relay contacts touch. This prevents arcing if the relay contacts bounce. This auxiliary
relay also activates the mechanical flag that indicates that the relay has operated.
When the circuit breaker being controlled by the relay opens, the relay coil is deenergized by an
auxiliary contact on the breaker. This protects the relay contacts, which are rated to make
currents up to 30A but should not break the inductive current of the breaker tripping circuit, to
prevent arcing wear. The disk is then returned to its initial position by the spring. The relay is
reset. Reset time is the time required to return the contacts fully to their original position.
Contacts part about 0.1 sec (six cycles) after the coil is deenergized. The total reset time varies
with the relay type and the time-dial setting. For a maximum time-dial setting (contacts fully
open), typical reset times might be 6 sec for an inverse-time relay and up to 60 sec for a very
inverse or extremely inverse relay. At lower time-dial settings, contact opening distance is less,
therefore reset time is lower.
A solid-state relay is not dependent on mechanical forces or moving contacts for its operation but
performs its functions electronically. Therefore, the timing can be very accurate even for currents
as low as the pickup value. There is no mechanical contact bounce or arcing, and reset times can
be extremely short.
CT and PT selection

In selecting instrument transformers for relaying and metering, a number of factors must be
considered; transformer ratio, burden, accuracy class, and ability to withstand available fault
currents.
CT ratio. CT guidelines mentioned earlier are to have rated secondary output at 110 to 125% of
expected load and no more than 100A secondary current at maximum primary fault current.
Where more than one CT ratio may be required, CTs with tapped secondary windings or multiwinding secondaries are available.
CT burden. CT burden is the maximum secondary load permitted, expressed in voltamperes (VA)
or ohms impedance, to ensure accuracy. ANSI standards list burdens of 2.5 to 45VA at 90%
power factor (PF) for metering CTs, and 25 to 200VA at 50% PF for relaying CTs.
CT accuracy class. ANSI accuracy class standards are [+ or -] 0.3, 0.6, or 1.2%. Ratio errors
occur because of [I.sup.2]R heating losses. Phase-angle errors occur because of magnetizing core
losses.
CTs are marked with a dot or other polarity identification on primary and secondary windings so
that at the instant current is entering the marked primary terminal it is leaving the marked
secondary terminal. Polarity is not required for overcurrent sensing but is important for
differential relaying and many other relaying functions.
PT ratio. PT ratio selection is relatively simple. The PT should have a ratio so that, at the rated
primary voltage, the secondary output is 120V. At voltages more than 10% above the rated
primary voltage, the PT will be subject to core saturation, producing voltage errors and excess
heating.
PT burden. PTs are available for burdens from 12.5VA at 10% PF to as high as 400VA at 85%
PF.
PT accuracy. Accuracy classes are ANSI standard [+ or -] 0.3, 0.6, or 1.2%. PT primary circuits,
and where feasible PT secondary circuits as well, should be fused.
CTs and PTs should have adequate capacity for the burden to be served and sufficient accuracy
for the functions they are to perform. However, more burden or accuracy than necessary will
merely increase the cost of the metering transformers. Solid-state relays usually impose lower
burdens than electromechanical relays.
SUGGESTED READING
Books:
ANSI/IEEE Standard 242-1986, IEEE Recommended Practice for Protection of Industrial and
Commercial Power Systems (Buff Book).
RELATED ARTICLE: Standard device-function numbers (partial list) per ANSI C3 7.2.

2. Time delay starting or closing relay. 3. Checking or interlock relay 6. Starting circuit breaker
8. Control power disconnecting device 12. Overspeed device 14. Underspeed device 15. Speedor frequency-matching device 18. Accelerating or decelerating device 19. Starting-to-running
transition contactor 21. Distance relay 23. Temperature-control device 25. Synchronizing or
synchronism-check device 27. Undervoltage relay 30. Annunciator relay 32. Directional power
relay 36. Polarity device 37. Undercurrent or underpower relay 40. Loss of Excitation (Field)
relay 41. Field circuit breaker 42. Running circuit breaker 43. Manual transfer or selector device
46. Reverse-phase or phase-balance current relay 47. Phase-sequence voltage relay 48.
Incomplete sequence relay 49. Machine or transformer thermal relay 50. Instantaneous
overcurrent or rate-of-rise relay 51. AC time overcurrent relay 52. AC circuit breaker 55. Powerfactor relay 56. Field-application relay 59. Overvoltage relay 60. Voltage- or current-balance
relay 62. Time-delay stopping or opening relay 64. Ground-protective relay 67. AC directional
overcurrent relay 68. Blocking relay 69. Permissive control device 72. DC circuit breaker 74.
Alarm relay 76. DC overcurrent relay 78. Phase-angle measuring or out-of-step protective relay
79. AC reclosing relay 81. Frequency relay 82. DC reclosing relay 85. Carrieror pilot-wire
receiver relay 86. Locking-outrelay 87. Differential protective relay 91. Voltage directional relay
92. Voltage and power directional relay 94. Tripping or trip-free relay

Ground-fault coordination should include


MV cable shielding.
Feb 1, 1999 12:00 PM, DeDad, John A.
1 Comment ShareThis4
Find more articles on Ground Fault
Verification of shield performance under ground-fault conditions will prevent systemwide cable
damage.
As we all know, the metallic shielding in a medium voltage (MV) cable provides the necessary
uniform electric field within the cable's insulation. What is not as well known is its other equally
important function: that of carrying a portion of return ground-fault current. As such, your
ground-fault protection coordination study should address two specific shielding questions.
* What is the magnitude of ground-fault current that will be carried by the metallic shields of the
specific feeder, in respect to its raceway/cable configuration?
* Is the circular mil area of the metallic shield sufficient to carry the calculated ground-fault
currents within the operational time frame of the specific ground-fault relay and protective
device?
Functions of metallic shield

In MV cable constructions, the metallic shield functions as an electrical stress reducer. That is, it
keeps the electrical field within the insulation uniform and evenly distributed, without any high
stress points. The stress lines in an electrical field are always attracted to any grounded surface.
Since the metallic shield surrounds the MV insulation and is grounded, all stress lines within the
insulation radiate out uniformly from the copper or aluminum conductor, at the cable's center,
through the MV insulation, to the metallic shield.
Shielding also protects the MV cable from induced voltages, such as from adjacent power
conductors, and minimizes interference with communications circuits.
Of equal importance but not commonly known, the metallic shield also functions as a return path
for ground-fault currents. Its physical construction affects its current-carrying capabilities much
like a copper conductor: larger circular mil area decreases resistance and impedance, thereby
increasing current-carrying capacity. The Insulated Cable Engineers Association (ICEA)
Publication No. P-45-482 details the specific shield performance parameters required for various
cable constructions.
The potential problem
Many industrial MV power distribution systems employ low-resistance grounding, which allows
from 200A to 2000A to flow during solid line-to-ground faults. In most cases, the metallic
shielding on those MV cables beyond the fault carry part of the fault return current, which may
return along the shields of other conductors or other equipment ground paths. When this
happens, any MV cable shielding with inadequate current-carrying capacity will overheat and
cause cable damage.
Determining fault-current return paths and magnitudes
The typical MV feeder consists of a conduit or raceway, three shielded phase conductors, and a
ground conductor. If one of the phases sustains a line-to-ground fault, the return fault current will
divide itself among the metallic shields, ground conductor, and raceway (if metallic). The
magnitude of current flow in each of these paths will vary inversely as to the impedance of the
specific path. In other words, the higher the impedance of the specific ground path, the less
amount of fault current will flow in that ground path. The common belief is that most of the
ground-fault current returns on the faulted phase shield. This is true if the ground fault is from a
conductor to its own metallic shield. However, testing has shown that only 3 to 14% of the
available ground-fault current will flow through each cable metallic shield. The percentages,
shown in the Table on page 60, vary with the cable/conduit/ground wire configuration of the
tested feeder. The majority of fault return current will flow in the metallic conduit and ground
wire since their impedances are much lower than that of typical shielding constructions. For
ground-fault coordination study purposes, you can safely assume that 15% of the available fault
current will flow in each metallic shield.
Verifying shield short-circuit performance

ICEA P-45-482 provides an equation you can use to determine the required cross sectional area
of the metallic shield.
[M.sup.2] = [[I.sub.o].sup.2] t/[A.sup.2] (eq. 1) where
[I.sub.o] = fault current (amperes)
t = time of fault (seconds)
A = effective cross sectional area of shield (circular mils),
M = constant.
Depending on the MV cable voltage rating, configuration, and construction, the constant M will
vary. For 90 [degrees] C-rated cable with copper shielding and thermoplastic jacket (PVC or
thermosplastic CPE), or impregnated paper insulation, the constant M will be as follows.
* 0.063 for 5kV through 15kV and 25kV rated cables.
* 0.065 for 35kV to 46kV rated cables.
* 0.066 for 69kV rated cables.
For 90 [degrees] C-rated cable with copper shielding and thermoset jacket [neoprene, hypalon, or
thermoset polyethylene (CPE), the constant M will be as follows.
* 0.089 for 5 kV through 15kV and 25kV rated cables.
* 0.090 for 35kV to 46kV rated cables.
* 0.091 for 69kV rated cables.
To find out various performance characteristics of your proposed shield construction, you insert
the proper value for the constant M into Equation 1 and then solve this equation for the
characteristic in question. For example, if you want to know the maximum amount of time (t)
that a given fault current can flow in a given shield, you can use the following equation:
t = [(MA/[I.sub.o]).sup.2] (eq. 2)
Or, you can determine the maximum fault current ([I.sub.o]) that can flow in a given shield for a
given amount of time:
[I.sub.o] = MA/[square root of t] (eq. 3)
Or, you can determine the shield effective cross-sectional area (A) required to withstand a given
fault current for a given time:

A = [I.sub.o] [square root of t] / M (eq. 4)


As you can see, by inserting any known two of the three variables, you can verify that the copper
shield in each of your MV feeders is capable of carrying its portion of fault current within the
time requirements dictated by the respective ground-fault relay and overcurrent protection
device.
Adapting shield construction to required performance
You may find that standard shield constructions may not have sufficient cross-sectional area to
handle the fault current for the cable in question. In this situation, you can then consider the
following construction options that may satisfy the more demanding requirement.
* No. 14 AWG concentric wire strands in lieu of the standard No. 18 AWG sizing.
* Helically-applied 5-mil tape with 17% overlap.
* One layer of overlapped, helically-applied, 5-mil tape plus a second layer of reverse lay
overlapped, helically-applied, 5-mil tape.
* Corrugated longitudinal 8-mil tape.
You can get effective cross-sectional area values for each of the above constructions, per cable
size, from the cable manufacturer. By plugging this information into Equations 2, 3, or 4, you
will obtain the best option for the specific feeder in question.
Sample problem
Fig. 1 shows a single-line diagram of a typical MV distribution system, where a distribution
substation is fed by cable from a main switchgear lineup. Ground-fault protective relays are also
shown. Let's analyze this system closer.
What would happen if Relay 50G (for Feeder No. 1) fails? Well, Relay 251G would have to clear
the fault before the shield in Feeder No. 1 is damaged.
Looking upstream of Relay 50G, we see that the operating time of Relay 251G at maximum
ground fault is 0.4 sec. With Relay 50G's operating time of 0.1 sec, Feeder No 1's shield must be
capable of withstanding its portion of the total ground-fault current of 1000A for 0.5 sec without
sustaining damage. This portion is 15% of 1000A, or 150A.
Suppose Feeder No. 1 consists of three No. 4/0 AWG, 15kV, shielded conductors in a rigid steel
conduit without a ground wire. Its shield must be able to withstand 150A of ground-fault current
for 0.5 sec. Similarly, Feeder No. 2 shield must withstand 150A of ground-fault current for 1.3
sec (0.5 sec for Feeder No. 1 plus 0.8 sec for operation of Relay 351G).

By inserting these ground-fault current and time duration values into equation 4, you can
determine the required shield circular mil area. By comparing this value with the actual shield
circular mil area of the cable in question, you can confirm the adequacy of the shield
construction.
Shield withstand-limit curves
In situations where numerous calculations are required, especially in coordination studies
involving many MV feeders, you can use shield withstand-limit curves that you prepare yourself.
What's involved?
First, you obtain actual circular mil area values of metallic shields per cable voltage and shield
and cable construction.
Second, you plug in these values, along with either incremental time values or ground-fault
current values, into equations 2 or 3. The solutions are then plotted on log-log graph paper and
joined together to form a specific shield withstand-limit curve, as shown in Fig. 2.
By verifying that the required shield withstand falls below and to the left of the limit curve, you
confirm the adequacy of the proposed shield construction.

Electrical design with EMF in mind.


Feb 1, 1995 12:00 PM, Gaskell, John D.
0 Comments ShareThis2
What are electromagnetic fields and what efforts should be made to minimize them in designing
electrical systems today?
While it may be years before a clear consensus is established on the health risks of
electromagnetic fields (EMF) exposure, the topic remains highly controversial. In the meantime,
electrical systems can be designed and installed to minimize the extent of the magnetic field
component of EMF, often at little or no added cost. Before discussing the steps that can be taken
during the design stages of electrical distribution systems, let's find out what EMF is.
What is EMF?
EMF at 60Hz is really made up of two separate entities: electric fields and magnetic fields. An
electric field, which exists when voltage is present and which is easily blocked by metal, can
cause currents to flow on the surface of the human body. Electric fields are not generally
considered to be a biological hazard.

A magnetic field, which exists when current flows and which is not appreciably blocked by
common materials, can cause current to flow through the human body. Magnetic fields at 60 Hz
are considered a possible biological hazard.
A magnetic field decreases as the distance from the source increases. However, the configuration
of the source actually determines how quickly a field diminishes. Multiple conductors with
current flowing in opposite directions or 3-phase circuits have magnetic fields that are inversely
proportional to the distance squared. Appliances and transformers are point sources and their
fields drops off inversely proportional to the distance cubed.
Exposure factors
Without consensus in the scientific community that a hazard actually exists, economics play a
big factor in setting limits. We don't know for sure whether long-term, low-level exposure is
worse than short-term, high-level exposure.
Some say that switched fields are more dangerous than steady-state fields. Individual
characteristics of the person being exposed may need to be considered; age, health, and even
fertility and pregnancy may be factors. Obviously, more research is needed. But in the meantime,
it may be practical to make some educated guesses, set some interim exposure limits, and
consider some changes in design.
Practical design suggestions
Services. For an overhead MV service lateral, consider selecting spacer cable in lieu of cross-arm
construction. Utilities are becoming more sensitive to the magnetic field issue and may welcome
your suggestions. Another choice for a service lateral, of course, is an underground service.
Preferably, pass a service lateral under a storage room rather than an office that is occupied for
long time periods. If this cannot be done, then the service lateral should be enclosed in a metallic
raceway. The magnetic field will induce a counter EMF in the metallic raceway, which will help
reduce the field.
Locate pad-mounted transformers at least 20 ft from buildings. Another consideration is to
encircle a pad-mounted transformer with a fence, located at a 4- to 6-ft distance from the edge of
the pad.
Switchboards and panels. Locate distribution equipment on exterior walls or walls adjacent to
storage areas or corridors. Where possible, locate free-standing switchboards to allow 3 ft of
working space on all sides. Use walls abutting occupied spaces for telephone or fire alarm
equipment. Avoid locating an electrical room directly below or above an occupied space.
Indoor transformers. Although a typical dry-type, step-down power transformer creates a large
magnetic field, this field's strength dissipates very quickly. However, it would still be good
practice to locate such a unit in an electrical room remote from occupied spaces.

Bus duct. For vertical distribution of power in a tall building, follow a procedure similar to an
indoor transformer siting. Where possible, run a vertical bus duct on a wall common to the
elevator shaft, janitor closet, or corridor. In a factory, a bus duct run should be located away from
operator positions.
Underfloor ducts. Individual conductors are usually installed in an underfloor duct system, and it
is possible for the phase and neutral conductors of a 2-wire branch circuit to be from 6 to 23 in.
apart (depending on the cross section of the cell). This configuration would create significant
magnetic fields. Recommended practice would be to twist the individual conductors of a branch
circuit together in pairs. For a large new construction project, twisted cable assemblies can be
purchased from specialty cable companies with only a slight increase in cost. At an existing
installation, all the wires in the duct could be tie-wrapped at each outlet in the duct system.
What to watch out for
Wiring errors. When conductors are installed in such a way that circuit conductors are not in the
same cable or raceway, substantial magnetic fields are created by the separated supply and return
currents. The most common error is violating Sec. 250-23(a) by making a connection between a
neutral and an equipment grounding conductor on the load side of the service disconnect; this
usually happens at sub-panels. In this instance, neutral currents will flow on both the neutral and
the equipment grounding conductor.
Other errors include incorrect wiring of 3-way switching circuits and the connection of neutrals
from two different branch circuits at some point other than panel neutral busses, such as at
junction boxes, switches, receptacles, etc.
Water pipe problems. Three specific areas can be addressed here.
* Neutral grounds to a water pipe on the load side of the service can cause problems similar to
the case mentioned above, where Sec. 250-23(a) is violated by making a connection between a
neutral and an equipment grounding conductor, on the load side of the service disconnect. If any
electrical equipment, such as a hot water heater, is also connected to a metal water pipe, the
water pipe will probably become a parallel path for current that should be flowing over the
neutral.
* Physical damage or corrosion can cause the neutral conductor (often uninsulated) on an
overhead service drop to open up, or sever. With this condition and with water supply laterals
connected to conductive water mains, the water system can function as a parallel neutral. When
this happens, unbalanced neutral currents seeking to return to the utility source will pass out of
the building over the water lateral. They will return to the source after passing through the water
main, a neighboring water lateral, and then through the neighboring service disconnect and out
over the neighboring neutral.
* Currents can originate outside a building from the power company distribution system
(typically, a looped primary circuit) or from neighboring buildings via the water service. An

insulating coupling, or fining, can be installed in the water service. However, it must be located
outside of the premises, at least 10 ft from the building wall, per NEC Sec. 250-81(a).
Note that installing an insulating fitting in a water line to prevent the entry of currents has the
disadvantage of decreasing ground conductivity, which under certain fault conditions may cause
excess voltage to appear on plumbing fixtures and appliance enclosures. Thus, an insulating
fitting can cause both a shock hazard and damage to an appliance that would not have occurred if
there had been a connection to the water system.
However, a device called an automatic ground connector, which reconnects the ground in the
event of a severe fault, is available. This device is similar in principle to a conventional surge
arrester, but operates on a different voltage range. Because this is a new device that has not yet
been tested by a recognized testing authority, approval from the local authority having
jurisdiction should be obtained.
When an insulating fitting is installed, a ground wire should be clamped beyond the fitting and
extended back into the building. This allows the option of installing an automatic ground
connector or reestablishing the ground in the future.
The importance of testing
Conducting a magnetic field survey upon completion of an installation in new construction is
recommended. Such a survey can be done also at an existing location to detect long standing
wiring errors or other abnormal conditions.
Two different types of meters are used. A single coil, or single axis, meter (about $235)
determines the direction of the field and is useful for finding the field source. This type of meter
can provide an approximate rms reading, but a calculation is required.
A 3-coil or 3-axis meter (about $1400) gives a true rms reading of fields from all directions and
is not dependent on the orientation of the meter. Capable of making readings every 4 sec over a
24-hr period, this type of meter can be left at a site, or a person can carry it along a
predetermined path. Accumulated data can be down-loaded into a computer.
John D. Gaskell, P.E. is President of Gaskell Associates, Ltd., a consulting engineering firm in
Warwick, R.I., and a member of the National Electromagnetic Field Testing Association.

What's the new TIA EIA 568-A standard?


Feb 1, 1995 12:00 PM, McElroy, Mark W.
0 Comments ShareThis4
Important revisions to the original TIA/EIA 568 standard may include new and previously issued
TSBs, and maybe even a protocol for testing UTP horizontal subsystems.

Revisions to one of the major voice/data telecommunications standards are in the process of
being made and it's important that you know about them. For one thing, the original TIA/EIA
568 standard is being updated to reflect certain performance enhancements to the use of
unshielded twisted pair (UTP) wiring for high-speed data networks. Also, the new standard,
TIA/EIA 568-A, may include previously issued Technical System Bulletins (TSBs) as well as
other important addenda.
The new standard, originally appearing in the form of the SP-2840 draft, is scheduled for another
(hopefully final) round of balloting this month. Nevertheless, it's important that you're aware of
these proposed revisions now. The following discussion provides a complete review of them.
What's included in the revised standard
The addenda for UTP wiring and connecting hardware issued after the original 568 standard was
released is to be included within the new standard. This material will no longer exist as addenda.
Also to be included within the 568-A standard are the following.
* TSB 36, Additional Cable Specifications for Unshielded Twisted Pair Cables (November
1991).
* TSB 40, Additional Transmission Specifications for Unshielded Twisted Pair Connecting
Hardware (August 1992).
* A separate addendum for shielded twisted pair (STP) wiring.
* A change in fiberoptic cabling connecting hardware from ST-type to SC-type connectors.
* The beginnings of a standard for "UTP link performance."
The big news
In the course of discussions with the 568 committee chairman, questions were raised about the
status of a procedure for testing UTP, which appears in the annex of the new 568-A standard (i.e.,
in the SP-2840). Referred to as Annex E, Unshielded Twisted Pair (UTP) Link Performance, this
annex attempts to establish a standard procedure for field testing the performance of UTP wiring
in horizontal subsystems (i.e., UTP wiring between wiring closets and work spaces).
UTP link performance refers to the performance of a complete end-to-end UTP. Performance, in
this context, refers to a cabling scheme's ability to reliably support a range of transmissions at
different frequencies. The ability to do so successfully or not is of great concern to end users and
installers alike when it comes to determining whether or not a cable plant is properly installed.
The problem has been that, in spite of all of the industry standards for telecommunications and
cabling, there are no standards to go by when it comes to how UTP cabling in the horizontal
subsystem should be tested.

The operative phrase on the subject is "end-to-end." This is meant to include testing of not just
the cabling, but the connecting hardware at both ends of a circuit as well. After all, cabling
systems are always installed with terminations at both ends of a circuit (i.e., an outlet at one end
and connecting hardware at the other). The use of cabling, therefore, always involves the
transmission of signals through the entire link, including outlets and connecting hardware. Thus,
it would seem appropriate to test not just the individual components but the entire link on an endto-end basis.
Unfortunately, all of the industry standards to date (especially the 568 standard) have focused on
performance parameters for individual components, but not on their performance in concert with
one another (i.e., in the form of a complete link).
Annex E of the new 568-A standard addresses this issue, but only in the form of a nonbinding,
"informative" manner. While this is helpful, it only begs the question of exactly how the industry
should handle this important matter.
And so the really big news on the horizon is that, while the 568-A standard will go no further
than the informative nature of its Annex E did in the SP-2840 draft, there is a new standard in the
works that will address the issue of the UTP testing in a definitive way. The new standard will be
referred to as follows: TIA/EIA Technical Systems Bulletin 67 (TSB-67), Link Performance
Specifications for Field Testing of Unshielded Twisted-Pair Cabling Systems. There is a draft
version of the TSB now in circulation, dated December 16, 1994.
The fact that we now have an addendum in circulation for a standard which, itself, has not yet
been finalized speaks for itself. It's no wonder that keeping up with the standards process can
sometimes feel like a wild goose chase. Nevertheless, it's pleasing to see that we finally have
some definitive movement on the subject of testing. Without consistent guidelines for testing,
there has really never been a truly meaningful way of determining whether or not complete endto-end systems have been properly installed, or in fact, meet the intended applications. Let's take
a closer look at the content of the new TSB-67 draft.
The proposed TSB-67
First of all, TSB-67 concerns itself with only 100-ohm UTP cabling and connecting hardware in
the horizontal subsystems of structured cabling systems. The "link" of interest, therefore,
generally includes outlets in faceplates at workstations, the UTP wiring itself, and terminating or
connecting hardware inside wiring closets. In addition, TSB-67 allows for a transition connector
close to the workstation, which might take the form of a junction box or some other crossconnection device used to extend horizontal UTP wiring to the workstation. (Note: this practice
was sometimes used in the past for nonpermanent wiring, but is strongly discouraged for new
installations of structured cabling systems.)
In terms of specific configurations defined by the new standard, TSB-67 actually defines two
possibilities. The first type is referred to as a channel and consists of the outlet, the transition
connector, the UTP wiring itself, connecting hardware in the wiring closet, and dedicated line or
patch cords on each end. See Fig. 1 for a view of this first configuration.

Table
Frequency
(MHz)
1-31.25
31.26-100
The second configuration is referred to by the new standard as a basic link, as shown in Fig. 2,
and is much more consistent with contemporary systems design. It basically assumes that there
are no intermediate cross-connections or splices in the horizontal UTP run, and includes the
outlet device at the workstation end and the connecting hardware at the closet end. The basic link
also allows for a line or patch cord at each end, but recognizes that these ancillary cords are
independent of the permanent horizontal subsystem and will vary depending on the types of
systems in use at any given time.
Per the parameters of the 568 standard, the new TSB-67 standard also assumes that in any
configuration, the length of the horizontal UTP wiring does not exceed 90 m, and the combined
length of the line and patch cords at either end of the link/channel does not exceed 10 m.
The new standard defines five specific tests to be performed on the UTP link.
* Wire map
* Length
* Attenuation
* Near end crosstalk (NEXT)
* Propagation delay
The attenuation and NEXT values are to be taken using swept/stepped frequency voltage
measurements, or equivalent methods, using field test equipment. The step size for NEXT tests
are as shown in Table 1. The step size for attenuation tests should not exceed 1 MHz.
Wire map test. The purpose of the wire map test is to determine whether or not the continuity and
polarity of pairs and conductors have been properly installed. TSB-67 further states that for each
of the 8 conductors in a 4-pair drop, the wire map test examines the following.
* Proper pin termination at each end.
* Continuity to the remote end.

* Shorts between any two or more conductors.


* Crossed, reversed, split pairs, or any other miswires.
Length test. The length test is just as it sounds: a measure of the length of the link to determine
compliance with 568 specifications. The test, however, is conducted in two slightly different
ways, depending on which of the two test configurations is in use. Since the channel type
configurations will, by definition, include terminated line and/or patch cords at both ends, the
length of the link could be up to the full 100 m allowed by the 568 standard.
In the case of the basic link configuration, however, the cable plant is assumed to be of the
structured, permanent type, and should, therefore, be limited to 90 m. Added to the 90 m then are
line cords on each end to serve as leads to the testing devices; such lead should be no more than
2 m. This results in a total length of 94 m in the case of the basic link length test.
Attenuation rest. The attenuation test focuses on signal loss throughout the link or channel. It
takes into account the cumulative attenuation produced by all components in a link, including
each plug/jack combination, patch and/or line cords, and of course the UTP cable itself.
Attenuation values when measured in accordance with the proposed TSB-67 standard for basic
links are summarized in Table 2. (Note: values for channel type links are omitted since most
contemporary cable plants are of the basic link type.)
NEXT test. NEXT loss is a measure of signal coupling from one pair to another. As in the case of
the attenuation test, the measures are derived from swept/stepped frequency voltage
measurements. The test calls for measurements to be taken on all pair combinations. When
performed in accordance with the proposed TSB-67 standard for basic links, the results should
conform to the values summarized in Table 3. (Note: values for channel type links are omitted
since most contemporary cable plants are of the basic link.)
Propagation delay test. The TSB-67's treatment of this test is quite brief and reads as follows.
The propagation delay can be determined from phase angle measurements of the output signal
relative to the input signal. Alternately, the propagation delay can be determined (approximately)
from TDR measurements.
The test calls for the measurement of propagation delay on all pairs, with the worst case (longest
delay) to be reported.
Helpful annexes
Like most of the TIA/EIA standards, TSB-67 includes a series of annexes that provide supporting
documentation on procedures (Annex A), accuracy guidelines for test instruments (Annex B),
interpretive explanations for tests that fail (Annex C), and electrical length measurement
methods (Annex D). Annexes A and B are considered to be part of the formal standard, while
Annexes C and D are provided for informative purposes only.

When to expect TSB-67 ratification


As far as when the proposed TSB-67 will be ratified into formal existence, it's impossible to say.
If experience is any indicator, however, it should be no sooner than next summer. But even with
TSB-67 in place, you should still recognize that there is much more to a structured cabling
system than solely the horizontal subsystem. And there are certainly more cable types to consider
than simply UTP. Until such time as the TIA/EIA expands the scope of its testing standards to
include these other subsystems and cable types, testing methods and, therefore, metrics in the
industry will still be inconsistent and, therefore, somewhat arbitrary.
SUGGESTED READING
EC&M articles:
"The EIA/TIA 568 Cabling Standard" October '93 issue. "The Do's and Don'ts of UTP Wiring"
June '94 issue. "Testing Methods for UTP and Fibre" August '94 issue. For copies, call 913-9671801.
Mark W. McElroy is Senior Manager with Peat Marwick's Management Consulting Practice
Malvern, PA. He is professionally certified by BICSI as a registered communications distribution
designer (RCDD).

Mobile generators power up Newark Airport.


Feb 1, 1995 12:00 PM, Lawrie, Robert J.
0 Comments ShareThis1
Power outage stopped all operations at busy airport. Here's how teams of electrical firms
supplied auxiliary power to critical facilities.
A little before 9:00 am on Monday, Jan. 9, a pile driver slammed through 4 ft of underground
reinforced concrete into primary power feeders, cutting central power to Newark Airport, one of
the busiest airports in the country. At the three main terminals, lighting, sewage treatment
systems, pumps, baggage conveyers, certain security systems, reservation computers, terminal
information monitors, and HVAC systems all lost power. While some standby power for lighting
and critical systems was available, the airport essentially was down.
The control tower remained operational because it has its own auxiliary generator power.
Navigational aid systems and aviation marker lighting were on a separate utility power source.
Also, some utility power was available to certain parts of the airport. Nevertheless, the
movement of hundreds of flights and thousands of people through the airport was slowed
significantly and made extremely difficult.

As darkness set in about 5:00 pm, the airport was declared officially closed to passenger traffic
for public safety reasons. Many passengers decided to be bussed to other area airports, sit it out,
or just go home and try the next day.
Accurate details concerning the accident are not yet available. What is known so far is that a
huge pile driver was preparing support for a planned rental car parking ramp. Although it appears
that markers were in place, and preparations done, somehow the driver hit and severed the main
26kV primary lines feeding the airport complex, which includes three main terminals, a large
hotel, restaurants, and similar retail and service facilities.
Mobile generators supply power
Electrical maintenance people at the Airport, under the direction of John Hallenbeck,
immediately sprang into action. Having already planned for the need for auxiliary power, they
were able to quickly put into motion local electrical engineering and service firms, electrical
contractors, generator specialists, and other needed service companies.
All airport maintenance electricians were called to check existing facilities and to start hooking
up auxiliary power. In addition, calls were sent out to other Port Authorities (PA) of N.Y. and
N.J. (The PA operates Newark and many other major transportation facilities in the area.)
Immediately, electrical people from LaGuardia Airport and Port Newark as well as electricians
from the PA central maintenance group were dispatched to the Airport. Most PA electricians
worked continuously, well into the night, and some for a few days more, to be certain that
reliable power had been established.
At one point in time, nearly 9000kW of auxiliary power was in operation. This included small
units from about 50kW to large 1750kW units mounted in tractor-trailer vans.
Service firm/maintenance crew teamwork
At one terminal building operated by Continental airlines, an engineering and service company,
Longo Industries of Morris Plains, N.J., was called by the airline. John Ziomek, Longo electrical
engineer, brought in a crew that worked through the day and into the night hooking up mobile
generator power.
Another firm, EES/SPB Inc. of Rahway, N.J., a design-and-build auxiliary power company,
brought in nearly 6MW of mobile generator power. Its' crews worked through the day and
evening, not only connecting its own auxiliary equipment but also aiding others in supplying
cables, materials, and making connections.
Also involved were electrical people from Continental Airlines and the PA. Key people from
each of the four groups coordinated all activity with skill and dedication, with the only objective
being to get the airport up and running quickly.

At the Continental Airlines terminal building, one 1750kW, one 1000kW, and two 600kW
generators were connected respectively to three 4000A, 480/277V, 3-phase, 4-wire substations to
provide auxiliary power. This was in addition to two permanently installed 1000kW on-site
generators within the terminal. The mobile generators were positioned near each of three 4000A
services.
Longo technicians then selected appropriate routing and pulled temporary power cables from the
mobile generators to each of the three substations. While cables were being pulled, Continental
electricians were shutting off all non-essential circuits so that when power renamed, a power
surge would be avoided. Of particular importance was "not forgetting" to shut off appropriate
mechanically held contactors because they automatically would be latched into the ON position
even during a power outage.
Other areas in need
At the other terminals, most work was done by PA electrical crews helped by independent firms.
Various sized mobile generators supplying auxiliary power were provided, depending on the
facilities at the particular terminal.
Steve McPartland, field engineer for ESS (Emergency Essential Services) points out that, in
addition to extensive work at the Continental terminal, requests came from a Marriott Hotel at
the airport that needed power to house airline passengers. Thus, a 1000kW unit was hooked up to
supply 480/277V to the hotel.
Requests for power for other loads came fast and furious. A sewage pumping station needed a
400kW unit to resume operation. Pumps for transporting fuel to gate-located underground tanks
also required power. This was handled by East-West Electric, using generators supplied by
Aggreco, a generator rental company. The control tower already had its own standby generator,
but the PA wanted a mobile backup generator to assure maximum reliability. Thus, a 750kW unit
was hooked up here.
Connection techniques
All mobile generating units were van enclosed and completely self sufficient, with generator
controls, fuel supply, a main circuit breaker (CB) and 400A output CBs, protective relays, and
output cables with high-amperage, quick disconnect cable connectors. These connectors were of
the pin-and-sleeve type, with insulating boots over the connection point.
At some locations, a quick disconnect receptacle was permanently installed on the load side of
the main circuit breaker or main fused disconnect. These were provided through the panel door
for fast connection of future emergency power cables in event they were ever needed.
To carry 480V temporary power from the generators to main CBs at each substation or service,
electricians installed 200-ft lengths of No. 4/0 AWG welding cable, which is rated to carry 500A.
This cable provides needed flexibility and the 4/0 size is standardized to simplify installation.

For example, where up to 500A was needed, one cable per phase was used; up to 1000A, two
cables per phase were used, etc.
When making connections, electricians observed all safety rules. For example, CBs were opened
and locked out; series fuses were removed and the panel double-checked for presence of any
voltage. A line-side (utility) voltage indicator was also connected. This is simply a pilotlamp
connected to the line side in the main panel. Or, one of the signal lamps on the panel itself was
used if it related to the CB ON/OFF status. If voltage returned or somehow became present in the
panel, the pilotlamp signal would light. Also, an audible alarm was connected for double
protection. Then, the temporary cables were connected to the CB load terminals.
The Longo crew kept themselves informed as to progress by the local utility. When word came
that the main lines had been reconnected to provide temporary utility power, Longo personnel
checked the utility power sources to make certain that proper phase sequence had been
established. This was essential to assure the proper direction of motor rotation. Voltage levels
were checked and those branch circuits designated to be OFF to avoid a surge were doubled
checked. When all circuits had been checked and cleared, auxiliary power was removed and
utility power re-connected.

Sensible transformer maintenance. (part 2)


Feb 1, 1999 12:00 PM, Raymond, Charles T.
1 Comment ShareThis12
Find more articles on: Transformers
Following specific checking and maintenance guidelines as well as conducting routine
inspections will help ensure the prolonged life and increased reliability of a dry-type transformer.
Because dry-type transformers are used so extensively in industrial, commercial, and
institutional power distribution systems, their maintenance should be a top priority. As such, you
should have a thorough knowledge of their maintenance requirements, which are similar to
liquid-filled units in many ways but differ enough to warrant separate coverage. The following
detailed discussion will help you attain the required knowledge.
Dry-type transformer classifications
Dry-type transformers are classified as ventilated, nonventilated, and sealed units, with each type
detailed in the ANSI/IEEE C57.12.01-1989 standard, General Requirements for Dry-Type
Distribution and Power Transformers. Because there are significant differences among these
three groups and because some of these differences have an impact on maintenance procedures,
it's important that you know the key aspects of each transformer type.

A ventilated dry-type transformer is constructed so that ambient air can circulate through vents in
the surrounding enclosure and cool the transformer core and coil assembly.
A nonventilated transformer operates with air at atmospheric pressure in an enclosure that does
not allow ambient air to circulate freely in and out.
A sealed transformer is self-cooled, with the enclosure sealed to prevent any entrance of ambient
air. These transformers are filled with an inert gas and operate at a positive pressure.
While construction varies per transformer type, inspection and maintenance guidelines are
somewhat similar.
Maintenance guidelines
As with liquid-filled transformers, a maintenance program for dry-type units should include
routine inspections and periodic checks. Acceptance tests should be performed when new units
are delivered as well as when the need is indicated by review of maintenance data and operating
history. The frequency for these inspections and checks will depend on the transformer
classification as well as the operating environment, load conditions, and requirements for safety
and reliability.
A valuable reference source for maintenance procedures is the ANSI/IEEE C57.94-1982
Standard, Recommended Practice for Installation, Application, Operation, and Maintenance of
Dry-Type General Purpose Distribution and Power Transformers, which covers many of the
maintenance aspects that should be considered.
The frequency of periodic checks will depend on the degree of atmospheric contamination and
the type of load applied to the transformer.
This is especially true for nonsealed transformers since ambient air and any contaminant dust or
vapors it carries can contaminate the internal, electrically-stressed components. As routine
inspections are made, the rate of accumulation of dust and moisture on the visible surfaces
should serve as a guide for scheduling periodic maintenance. Thus, ventilated transformers will
require more frequent periodic checks than nonventilated units. Sealed transformers will require
less frequent periodic checks than either type, because of their construction.
Routine checks and resultant maintenance
Neither nonventilated nor ventilated dry-type transformers have indicating gauges, as are needed
on liquid-filled transformers, to monitor temperature, pressure, and liquid level. Thus, routine
checks are more subjective and consist mainly of visual and audible observations.
Sealed dry-type transformers do have pressure gauges and these should be routinely checked.
Also, a complete checklist should be developed for each transformer and should include essential
observations, with data recorded and preserved. A program recommending various type checks
and their frequency is shown in the Table on page 86.

Dust accumulation. Visual inspections should cover louvers, screens, and any visible portions of
internal coil cooling ducts for accumulated dust. Do not remove any panel or cover unless the
transformer is deenergized. If dust accumulation is excessive, you should deenergize the
transformer in accordance with established safety procedures, remove its side panels, and
vacuum away as much of the dust as possible. Then, clean with lint free rags or soft bristled
brushes. Do not use any solvents or detergents as these may react with the varnishes or insulating
materials and lead to accelerated deterioration. They may also leave residues that will enhance
future accumulation of dust and various contaminates.
If dust accumulation remains in inaccessible areas after vacuuming, you can blow dry air into the
unit to clear ducts. You should use air or nitrogen that has a dew point of -50 [degrees] F or less
and regulate the pressure at or below 25 psi.
Checks during deenergization. The following items should be done while the transformer is
deenergized.
* When access panels are removed for cleaning, all insulation surfaces should be inspected for
signs of discoloration, heat damage, or tree-like patterns etched into the surface that are
characteristic of corona damage. The core laminations should be inspected for signs of arcing or
over-heating.
* All accessible hardware should be checked for tightness.
* Isolation dampeners between the base of the transformer and the floor should be checked for
deterioration.
* Cooling fans or auxiliary devices should be inspected and cleaned.
If the transformer is deenergized long enough so that it can cool to ambient temperature, make
sure that the unit is kept dry. If the ambient air is very humid, you may have to heat the
transformer with electrical strip heaters to avoid condensation of moisture on the winding
insulation. This is very important because a large percentage of dry-type transformer failures
occur after extended shutdowns, when the insulation is allowed to cool and moisture in the
ambient air condenses on the insulation.
Checks with transformer energized. The following items should be done with the transformer
energized.
* Pressure readings should be checked and recorded for transformers with sealed [TABULAR
DATA OMITTED] tank construction. The ambient temperature, time of day, and loading
conditions should be recorded along with the pressure.
* Audible sound should be monitored, concentrating on the sound's characteristics as well as its
level. Any noticeable change in the sound level or characteristics should be recorded. Significant
changes could be indicative of loose clamping hardware, defective vibration isolators, over
excitation, or possibly damage to the primary winding insulation.

* Proper ventilation should be verified. Although few dry-type transformers are equipped with
temperature gauges, the effectiveness of ventilation can be verified by measuring the air
temperature at the inlet (which should be near the floor) to an enclosed room and then measuring
either the ambient temperature of the air in the enclosed space or the temperature of the air at the
exhaust (which should be in the upper part of the room). The average temperature of the room
should not increase more than 40 [degrees] F over the incoming air and the exhaust should not
increase more than 60 [degrees] F. Additional details on ventilation requirements will be found in
ANSI/IEEE C57.94.
Periodic tests
You should conduct periodic testing as often as needed. The frequency is usually dependent on
the transformer's operating environment. If routine inspections indicate that cleaning is required,
periodic tests should be made at the shutdown for the cleaning operation, after the transformer is
thoroughly cleaned. The nominal period between scheduled tests is one year but this may be
longer or shorter, depending on the observed accumulation of contamination on the cooling
vents.
Sealed units should be opened only when the need is indicated by loss of pressure, operating
abnormalities, or at intervals as recommended in the manufacturer's instructions. With these
units, periodic tests should be confined to external inspections of the bushings and the
enclosures. Also, readings at external terminals should be taken of insulation resistance (IR),
power factor (PF), and turns ratio.
Section 6 (Testing) of ANSI/IEEE C57.94 states that induced or applied voltage tests also may be
done periodically. These are overvoltage tests; as such, they are not necessarily nondestructive,
even when skilled operators use the proper equipment. Damage may occur to the insulation,
which would otherwise be serviceable at normal operating voltage levels. Therefore, these tests
should not be considered as appropriate maintenance tests, which are by definition
nondestructive.
IR testing. The IR of each winding should be measured using a megohmmeter in accordance
with Sections 10.9 through 10.9.4 of the ANSI/IEEE C57.12.91-1979 Standard, Test Code for
Dry-Type Distribution and Power Transformers. The transformer should be deenergized and
electrically isolated with all terminals of each winding shorted together. The windings not being
tested should be grounded. The megohmmeter should be applied between each winding and
ground (high voltage to ground and low voltage to ground) and between each set of windings
(high voltage to low voltage). The megohm values along with the description of the instrument,
voltage level, humidity, and temperature should be recorded for future reference.
The minimum megohm value for a winding should be 200 times the rated voltage of the winding
divided by 1000. For example, a winding rated at 13.2kV would have a minimum acceptable
value of 2640 megohms ([13,200V x 200] / 1000). If previously recorded readings taken under
similar conditions are more than 50% higher, you should have the transformer thoroughly
inspected, with acceptance tests performed before reenergizing.

Turns ratio testing. The transformer turn ratio is the number of turns in the high voltage winding
divided by the number of turns in the low voltage winding. This ratio is also equal to the rated
phase voltage of the high voltage winding being measured divided by the rated phase voltage of
the low voltage winding being measured.
Transformer turns ratio measurements are best made with specialized instruments that include
detailed connection and operating instructions. ANSI/IEEE Standard C57.12.91 describes the
performance and evaluation of these tests. The measured turns ratio should be within 0.5% of the
calculated turns ratio. Ratios outside this limit may be the result of winding damage, which has
shorted or opened some winding turns.
Insulation PF testing. Insulation PF is the ratio of the power dissipated in the resistive component
of the insulation system, when tested under an applied AC voltage, divided by the total AC
power dissipated. A perfect insulation would have no resistive current and the PF would be zero.
As insulation PF increases, the concern for the integrity of the insulation does also. The PF of
insulation systems of different vintages and manufacturers of transformers varies over a wide
range (from under 1% to as high as 20%). As such, it's important that you establish a historic
record for each transformer and use good judgment in analyzing the data for significant
variations. ANSI/IEEE Standard C57.12.91 describes the performance and evaluation of
insulation PF testing.
Acceptance testing
Acceptance tests (defined in Part 1, June 1994 issue, which concentrated on liquid-filled
transformers) are those tests made at the time of installation of the unit or following a service
interruption to demonstrate the serviceability of the transformer. This testing also applies to drytype units. The acceptance tests should include IR testing, insulation PF measurement, and turns
ratio testing, all as described under periodic tests. In addition, winding resistance measurements
should be made and excitation current testing done. If you have a particular cause for concern,
say a significant fault in the secondary circuit or a severe overload, you should make an
impedance measurement and possibly an applied voltage test.
Winding resistance measurement. Accurate measurement of the resistance between winding
terminals can give you an indication of winding damage, which can cause changes to some or all
of the winding conductors. Such damage might result from a transient winding fault that cleared;
localized overheating that opened some of the strands of a multistrand winding conductor; or
short circuiting of some of the winding conductors.
Sometimes, conductor strands will burn open like a fuse, decreasing the conductor cross section
and resulting in an increase in resistance. Occasionally, there may be turn-to-turn shorts causing
a current bypass in part of the winding; this usually results in a decrease of resistance.
To conduct this test, you should de-energize the transformer and disconnect it from all external
circuit connections. A sensitive bridge or micro-ohmmeter capable of measuring in the microohm range (for the secondary winding) and up to 20 ohms (for the primary winding) must be
used. These values may be compared with original test data corrected for temperature variations

between the factory values and the field measurement or they may be compared with prior
maintenance measurements. On any single test, the measured values for each phase on a 3-phase
transformer should be within 5% of the other phases.
Excitation on current measurement. The excitation current is the amperage drawn by each
primary coil, with a voltage applied to the input terminals of the primary and the secondary or
output terminals open-circuited. For this test, you should disconnect the transformer from all
external circuit connections. With most transformers, the reduced voltage applied to the primary
winding coils may be from a single-phase 120V supply. The voltage should be applied to each
phase in succession, with the applied voltage and current measured and recorded.
If there is a defect in the winding, or in the magnetic circuit that is circulating a fault current,
there will be a noticeable increase in the excitation current. There is normally a difference
between the excitation current in the primary coil on the center leg compared to the that in the
primary coils on the other legs; thus, it's preferable to have established benchmark readings for
comparison.
Variation in current versus prior readings should not exceed 5%. On any single test, the current
and voltage readings of the primary windings for each of the phases should be within 15% of
each other.
Applied voltage testing. The applied voltage test is more commonly referred to as the "hi-pot
test." This test is performed by connecting all terminals of each individual winding together and
applying a voltage between windings as well as from each winding to ground, in separate tests.
Untested windings are grounded during each application of voltage.
Although ANSI/IEEE C57.94 lists the applied voltage test as an optional pre-service or periodic
test, this test should be used with caution as it can cause insulation failure. It should be regarded
as a proof test to be conducted when there has been an event or pattern in the transformer's
operating history that makes its insulation integrity suspect.
ANSI/IEEE C57.94 states that either AC or DC voltage tests are acceptable for applied potential
testing but that the DC applied voltage should not exceed the rms value of the standard test level.
AC voltage rms values are limited by C57.94 to 75% of the original test levels (these levels
range from 2 to 4 times the operating voltage) for initial installation tests and 65% of the original
test levels for routine maintenance tests. The original or factory test levels are specified in
ANSI/IEEE C57.12.01 and the tests are described in ANSI/IEEE C57.12.91. You should review
these standards carefully before conducting any applied potential tests. If the original factory test
reports are available, you should consult them to determine the original factory test levels.
DC applied voltage tests are often conducted in the field because DC test sets are smaller and
more readily available than AC applied voltage sets. With DC tests, the leakage current can be
measured and is often taken as a quantitative measure. However, DC leakage current can vary
considerably from test to test because of creepage across the complex surfaces between windings
and between windings and ground.

The use of AC voltage is preferable since the transformer insulation structures were designed,
constructed, and tested with the application of AC voltage intended.
Impedance testing. An impedance test may be useful in evaluating the condition of transformer
windings, specifically for detecting mechanical damage following rough shipment or a service
fault on the output side that caused high fault currents to flow through the transformer windings.
Mechanical distortion of the windings will cause a change in their impedance. To maximize the
effectiveness of this test, you should take a measurement during the transformer's initial
installation to establish a benchmark value.
An impedance test is performed by electrically connecting the secondary terminals together with
a conductor capable of carrying at least 10% of the line current and applying a reduced voltage to
the primary windings. This is easily accomplished by applying a single-phase voltage to each
phase in succession. The applied voltage is measured at the primary terminals and the current
measured in each line.
You should record these values and then calculate the ratio of voltage to current for each phase.
This ratio should be within 2% for each phase and should not vary more than 2% between tests.
A variation of more than 2% indicates the possibility of mechanical distortion of the winding
conductors, which should be investigated as soon as possible.
Forensic investigators continually encounter instances where transformers have failed following
an extended period of proper maintenance. Some of these failures were unpredictable but many
could have been prevented if available data had been carefully reviewed and results properly
interpreted.
The final section of this article, scheduled for a future issue, will address the interpretation of
maintenance records and how to use recorded data to identify trends that may indicate
developing problems with any specific transformer.
SUGGESTED READING
EC&M Books:
Practical Guide to Applying, Installing, and Maintaining Transformers. For ordering information,
call 800-654-6776.
EC&M Articles:
"Sensible Transformer Maintenance - Part 1" June '94 issue. For copies, call 913-967-1801.
Standards:
ANSI/IEEE C57.12.01, General Requirements for Dry-Type Distribution and Power
Transformers

ANSI/IEEE C57.94, Recommended Practice for Installation, Application, Operation, and


Maintenance of Dry-Type General Purpose Distribution and Power Transformers
ANSI/IEEE C57.12.91, Standard Test Code for Dry-Type Distribution and Power Transformers
For ordering information, call 800-678-IEEE.
Charles T. Raymond is a professional engineer registered in N.Y. and is manager of transformer
services for G.E. Co. in Schenectady, N.Y.
Email1 0 5

Sizing manholes.
Feb 1, 1995 12:00 PM
1 Comment ShareThis3
We were asked to comment on an installation similar to the one pictured in the drawing. Half of
the conductors coming in from the 24-duct bank turn out toward the reader, and the other half go
the other way. In other words, the manhole is being used as a very large tee. Note that with the
voltage over 600V, traditionally this type of work has belonged to the electric utilities. In this
case, however, the service point is on the line side of this installation and the work appeared to
be all premises wiring covered by the NEC.
The EC&M Panel's analysis
We think that the installation is squarely within the scope of the NEC as covered in Sec. 90-2(a)
(1) and (3). The fact that the voltage is over 600V is irrelevant to whether or not it is within the
scope of the NEC. Nevertheless, the NEC does not include all the answers. A manhole is unique
in that, quite clearly, personnel must be able to enter and work within its confines. This isn't true
of any other pull box, and the NEC does not include provisions that directly address this point, at
least in terms of dimensional size. For example, Sec. 110-12(b) requires accessibility to be
maintained, but without giving any minimum dimensions. As with the racking requirement in
Sec. 370-28(b) (generalized to medium voltage by Sec. 370-70), this requirement applies at all
voltages. If there were terminations that must be worked hot, however, then the installation
would need to comply with the clearances given in Table 110-16(a) or Table 110-34(a) as
applicable.
If, as in the case at issue, there is nothing that would trigger workspace requirements, then there
is no minimum dimension limit in the NEC that directly applies to workspace access. There is,
however, a workspace rule in the National Electrical Safety Code (NESC) that does apply. Sec.
323 of that code covers, in great detail, the required strength, dimensions, access dimensions and

locations, cover supports, ladder requirements, drainage, ventilation, mechanical protection, and
identification of manholes. We think that the requirements in this section should be applied in
concert with those in the NEC, as indicated in Sec. 90-2(a)(1) (FPN):
(FPN): For additional information concerning such installations in an industrial or multibuilding
complex, see the National Electrical Safety Code, ANSI C-2-1990.
Minimum workspace
The NESC, in Part B of Sec. 323, reads in part as follows:
B. Dimensions
Manholes shall meet the following requirements: A clear working space sufficient for performing
the necessary work shall be maintained. The horizontal dimension of the clear working space
shall be not less than 3 ft (900 mm). The vertical dimensions shall be not less than 6 ft (1.80 m)
except in manholes where the opening is less than 1 ft (300 mm), horizontally, of the adjacent
interior side wall of the manhole. EXCEPTION 1: Where one boundary of the working space is
an unoccupied wall and the opposite boundary consists of cables only, the horizontal working
space between these boundaries may be reduced to 30 in. (750mm).
Remember, the idea is to apply the NESC in concert with the NEC, not to replace the NEC. For
example, in cases where NEC clearance rules apply, they should be applied because those rules
are generally more strict. This is appropriate. The NESC is designed to be used by highly
qualified and supervised utility personnel. Its provisions are generally less strict than the NEC.
We think, however, that in instances such as this, where the NEC is silent as to required work
spaces, the NESC should be consulted. This particular NESC rule fits quite well with NEC
provisions and we think it can and should be enforced. As we will show, however, in this case
there probably will be plenty of room at the unoccupied end of the manhole.
Cable bends
With regard to required spaces for routing conductors, it is the NESC that is silent. For example,
Sec. 341-A-1 of the NESC merely requires "Bending of the supply cable during handling,
installation, and operation shall be controlled to avoid damage." There is no specific dimension
rule.
On the other hand, Sec. 370-71(b) of the NEC is very specific. A box used for an angle pull, such
as the, one shown here, must meet the following dimensions:
(b) For Angle or U Pulls. The distance between each cable or conductor entry inside the box and
the opposite wall of the box shall not be less than thirty-six times the outside diameter, over
sheath, of the largest cable or conductor. This distance shall be increased for additional entries by
the amount of the sum of the outside diameters, over sheath, of all other cables or conductor
entries through the same wall of the box.

The distance between a cable or conductor entry and its exit from the box shall be not less than
thirty-six times the outside diameter, over sheath, of that cable or conductor.
We think that if this type of requirement is appropriate for bending allowances in the traditional
sheet-metal pull box, there is no justification for treating a manhole differently, where it is used
for the same purpose. The only difference would be to recognize personnel access and, in doing
so, to additionally check that the NESC minimums were observed.
In this case, however, the strict application of this rule leads to a grotesque result that we find
difficult to support. Using the formula in the rule based on 2 in. (1/6 ft) diameter cable we end up
(rounding up 2-in. to the nearest ft) with a 18-ft by 12-ft manhole (36 x 1/6 ft + 71 x 1/6 ft
[approximately equal to] 18 ft; 36 x 1/6 ft + 35 x 1/6 ft [approximately equal to] 12 ft). We note
that this section does not correlate with Sec. 370-28(a)(2) for 600V installations and below,
which was changed to require considering only the largest single calculated row instead of all the
entries on one wall of the box.
If that principle were carried over to this installation, the result would be large but more
manageable. If the concept of the largest row also includes the largest column, we would end up
(again after rounding up 2 in. to the nearest ft) with a 9-ft by 8-ft manhole (36 x 1/6 ft + 17 x 1/6
ft [approximately equal to] 9 ft; 36 x 1/6 ft + 11 x 1/6 ft [approximately equal to] 8 ft). We think
the Authority Having Jurisdiction should be consulted to see whether a waiver in this direction
might be entertained. Certainly one could point out the inconsistency with Sec. 370-28(a)(2).
Otherwise, there is no alternative to arranging a custom, palatially sized manhole.
Note also, that for medium voltage installations such as this one, the warning-label requirement
in Sec. 370-72(e) will apply. The "DANGER HIGH VOLTAGE - KEEP OUT" marking must be
added to the manhole cover, in addition or in place of the familiar "ELECTRICAL" marking
(Sec. 323-J of the NESC requires this identification but not the warning) on such covers. We are
aware that this is seldom done; however, this is in the same subsection that deals with typical
manhole covers, recognizing that their weight alone secures them to their enclosures.

You might also like