Вы находитесь на странице: 1из 20

Categorization of electric motors

The classic division of electric motors has been that of Direct Current
(DC) types vs Alternating Current (AC) types. This is more a de facto
convention, rather than a rigid distinction. For example, many classic
DC motors run happily on AC power.
The ongoing trend toward electronic control further muddles the
distinction, as modern drivers have moved the commutator out of the
motor shell. For this new breed of motor, driver circuits are relied upon
to generate sinusoidal AC drive currents, or some approximation of.
The two best examples are: the brushless DC motor and the stepping
motor, both being polyphase AC motors requiring external electronic
control.
There is a clearer distinction between a synchronous motor and
asynchronous types. In the synchronous types, the rotor rotates in
synchrony with the oscillating field or current (eg. permanent magnet
motors). In contrast, an asynchronous motor is designed to slip; the
most ubiquitous example being the common AC induction motor which
must slip in order to generate torque.

DC motors

A DC motor is designed to run on DC electric power. Two examples of


pure DC designs are Michael Faraday's homopolar motor (which is
uncommon), and the ball bearing motor, which is (so far) a novelty. By
far the most common DC motor types are the brushed and brushless
types, which use internal and external commutation respectively to
create an oscillating AC current from the DC source -- so they are not
purely DC machines in a strict sense.

Brushed DC motors

The classic DC motor design generates an oscillating current in a


wound rotor with a split ring commutator, and either a wound or
permanent magnet stator. A rotor consists of a coil wound around a
rotor which is then powered by any type of battery.
Many of the limitations of the classic commutator DC motor are due to
the need for brushes to press against the commutator. This creates
friction. At higher speeds, brushes have increasing difficulty in
maintaining contact. Brushes may bounce off the irregularities in the
commutator surface, creating sparks. This limits the maximum speed
of the machine. The current density per unit area of the brushes limits
the output of the motor. The imperfect electric contact also causes
electrical noise. Brushes eventually wear out and require replacement,
and the commutator itself is subject to wear and maintenance. The
commutator assembly on a large machine is a costly element,
requiring precision assembly of many parts.
Brushless DC motors

Some of the problems of the brushed DC motor are eliminated in the


brushless design. In this motor, the mechanical "rotating switch" or
commutator/brushgear assembly is replaced by an external electronic
switch synchronised to the rotor's position. Brushless motors are
typically 85-90% efficient, whereas DC motors with brushgear are
typically 75-80% efficient.
Midway between ordinary DC motors and stepper motors lies the realm
of the brushless DC motor. Built in a fashion very similar to stepper
motors, these often use a permanent magnet external rotor, three
phases of driving coils, one or more Hall effect sensors to sense the
position of the rotor, and the associated drive electronics. The coils are
activated, one phase after the other, by the drive electronics as cued
by the signals from the Hall effect sensors. In effect, they act as three-
phase synchronous motors containing their own variable-frequency
drive electronics. A specialized class of brushless DC motor controllers
utilize EMF feedback through the main phase connections instead of
Hall effect sensors to determine position and velocity. These motors
are used extensively in electric radio-controlled vehicles. When
configured with the magnets on the outside, these are referred to by
modelists as outrunner motors.
Brushless DC motors are commonly used where precise speed control
is necessary, as in computer disk drives or in video cassette recorders,
the spindles within CD, CD-ROM (etc.) drives, and mechanisms within
office products such as fans, laser printers and photocopiers. They
have several advantages over conventional motors:

• Compared to AC fans using shaded-pole motors, they are very efficient, running
much cooler than the equivalent AC motors. This cool operation leads to much-
improved life of the fan's bearings.
• Without a commutator to wear out, the life of a DC brushless motor can be
significantly longer compared to a DC motor using brushes and a commutator.
Commutation also tends to cause a great deal of electrical and RF noise; without a
commutator or brushes, a brushless motor may be used in electrically sensitive
devices like audio equipment or computers.
• The same Hall effect sensors that provide the commutation can also provide a
convenient tachometer signal for closed-loop control (servo-controlled)
applications. In fans, the tachometer signal can be used to derive a "fan OK"
signal.
• The motor can be easily synchronized to an internal or external clock, leading to
precise speed control.
• Brushless motors have no chance of sparking, unlike brushed motors, making
them better suited to environments with volatile chemicals and fuels.
• Brushless motors are usually used in small equipment such as computers and are
generally used to get rid of unwanted heat.
• They are also very quiet motors which is an advantage if being used in equipment
that is affected by vibrations.
Modern DC brushless motors range in power from a fraction of a watt
to many kilowatts. Larger brushless motors up to about 100 kW rating
are used in electric vehicles. They also find significant use in high-
performance electric model aircraft.

Coreless DC motors

Nothing in the design of any of the motors described above requires


that the iron (steel) portions of the rotor actually rotate; torque is
exerted only on the windings of the electromagnets. Taking advantage
of this fact is the coreless DC motor, a specialized form of a brush or
brushless DC motor. Optimized for rapid acceleration, these motors
have a rotor that is constructed without any iron core. The rotor can
take the form of a winding-filled cylinder inside the stator magnets, a
basket surrounding the stator magnets, or a flat pancake (possibly
formed on a printed wiring board) running between upper and lower
stator magnets. The windings are typically stabilized by being
impregnated with Electrical epoxy potting systems. Filled epoxies that
have moderate mixed viscosity and a long gel time. These systems are
highlighted by low shrinkage and low exotherm. Typically UL 1446
recognized as a potting compound for use up to 180C (Class H) UL File
No. E 210549.
Because the rotor is much lighter in weight (mass) than a conventional
rotor formed from copper windings on steel laminations, the rotor can
accelerate much more rapidly, often achieving a mechanical time
constant under 1 ms. This is especially true if the windings use
aluminum rather than the heavier copper. But because there is no
metal mass in the rotor to act as a heat sink, even small coreless
motors must often be cooled by forced air.
These motors were commonly used to drive the capstan(s) of magnetic
tape drives and are still widely used in high-performance servo-
controlled systems, like radio-controlled vehicles/aircraft, humanoid
robotic systems, industrial automation, medical devices, etc.

Universal motors

A variant of the wound field DC motor is the universal motor. The name
derives from the fact that it may use AC or DC supply current, although
in practice they are nearly always used with AC supplies. The principle
is that in a wound field DC motor the current in both the field and the
armature (and hence the resultant magnetic fields) will alternate
(reverse polarity) at the same time, and hence the mechanical force
generated is always in the same direction. In practice, the motor must
be specially designed to cope with the AC current (impedance must be
taken into account, as must the pulsating force), and the resultant
motor is generally less efficient than an equivalent pure DC motor.
Operating at normal power line frequencies, the maximum output of
universal motors is limited and motors exceeding one kilowatt are rare.
But universal motors also form the basis of the traditional railway
traction motor in electric railways. In this application, to keep their
electrical efficiency high, they were operated from very low frequency
AC supplies, with 25 Hz and 16 2/3 hertz operation being common.
Because they are universal motors, locomotives using this design were
also commonly capable of operating from a third rail powered by DC.
The advantage of the universal motor is that AC supplies may be used
on motors which have the typical characteristics of DC motors,
specifically high starting torque and very compact design if high
running speeds are used. The negative aspect is the maintenance and
short life problems caused by the commutator. As a result such motors
are usually used in AC devices such as food mixers and power tools
which are used only intermittently. Continuous speed control of a
universal motor running on AC is very easily accomplished using a
thyristor circuit, while stepped speed control can be accomplished
using multiple taps on the field coil. Household blenders that advertise
many speeds frequently combine a field coil with several taps and a
diode that can be inserted in series with the motor (causing the motor
to run on half-wave rectified AC).
Universal motors can rotate at relatively high revolutions per minute
(rpm). This makes them useful for appliances such as blenders,
vacuum cleaners, and hair dryers where high-speed operation is
desired. Many vacuum cleaner and weed trimmer motors exceed
10,000 rpm, Dremel and other similar miniature grinders will often
exceed 30,000 rpm. Motor damage may occur due to overspeed (rpm
in excess of design specifications) if the unit is operated with no
significant load. On larger motors, sudden loss of load is to be avoided,
and the possibility of such an occurrence is incorporated into the
motor's protection and control schemes. Often, a small fan blade
attached to the armature acts as an artificial load to limit the motor
speed to a safe value, as well as provide cooling airflow to the
armature and field windings.
With the very low cost of semiconductor rectifiers, some applications
that would have previously used a universal motor now use a pure DC
motor, sometimes with a permanent magnet field.
AC motors
In 1882, Nikola Tesla identified the rotating magnetic field principle,
and pioneered the use of a rotary field of force to operate machines.
He exploited the principle to design a unique two-phase induction
motor in 1883. In 1885, Galileo Ferraris independently researched the
concept. In 1888, Ferraris published his research in a paper to the
Royal Academy of Sciences in Turin.
Introduction of Tesla's motor from 1888 onwards initiated what is
sometimes referred to as the Second Industrial Revolution, making
possible the efficient generation and long distance distribution of
electrical energy using the alternating current transmission system,
also of Tesla's invention (1888).[2] Before the invention of the rotating
magnetic field, motors operated by continually passing a conductor
through a stationary magnetic field (as in homopolar motors).
Tesla had suggested that the commutators from a machine could be
removed and the device could operate on a rotary field of force.
Professor Poeschel, his teacher, stated that would be akin to building a
perpetual motion machine.[3] Tesla would later attain U.S. Patent
0,416,194 , Electric Motor (December 1889), which resembles the
motor seen in many of Tesla's photos. This classic alternating current
electro-magnetic motor was an induction motor.
Michail Osipovich Dolivo-Dobrovolsky later invented a three-phase
"cage-rotor" in 1890. This type of motor is now used for the vast
majority of commercial applications.

Components

A typical AC motor consists of two parts:

1. An outside stationary stator having coils supplied with AC current to produce a


rotating magnetic field, and;
2. An inside rotor attached to the output shaft that is given a torque by the rotating
field.

Torque motors

A torque motor is a specialized form of induction motor which is


capable of operating indefinitely at stall (with the rotor blocked from
turning) without damage. In this mode, the motor will apply a steady
stall torque to the load (hence the name). A common application of a
torque motor would be the supply- and take-up reel motors in a tape
drive. In this application, driven from a low voltage, the characteristics
of these motors allow a relatively-constant light tension to be applied
to the tape whether or not the capstan is feeding tape past the tape
heads. Driven from a higher voltage, (and so delivering a higher
torque), the torque motors can also achieve fast-forward and rewind
operation without requiring any additional mechanics such as gears or
clutches. In the computer world, torque motors are used with force feedback
steering wheels.

Slip ring

The slip ring or wound rotor motor is an induction machine where the rotor comprises
a set of coils that are terminated in slip rings to which external
impedances can be connected. The stator is the same as is used with a
standard squirrel cage motor.
By changing the impedance connected to the rotor circuit, the
speed/current and speed/torque curves can be altered.
The slip ring motor is used primarily to start a high inertia load or a
load that requires a very high starting torque across the full speed
range. By correctly selecting the resistors used in the secondary
resistance or slip ring starter, the motor is able to produce maximum
torque at a relatively low current from zero speed to full speed. A
secondary use of the slip ring motor is to provide a means of speed
control. Because the torque curve of the motor is effectively modified
by the resistance connected to the rotor circuit, the speed of the motor
can be altered. Increasing the value of resistance on the rotor circuit
will move the speed of maximum torque down. If the resistance
connected to the rotor is increased beyond the point where the
maximum torque occurs at zero speed, the torque will be further
reduced.
When used with a load that has a torque curve that increases with
speed, the motor will operate at the speed where the torque developed
by the motor is equal to the load torque. Reducing the load will cause
the motor to speed up, and increasing the load will cause the motor to
slow down until the load and motor torque are equal. Operated in this
manner, the slip losses are dissipated in the secondary resistors and
can be very significant. The speed regulation is also very poor.
Stepper motors

Closely related in design to three-phase AC synchronous motors are


stepper motors, where an internal rotor containing permanent magnets
or a large iron core with salient poles is controlled by a set of external
magnets that are switched electronically. A stepper motor may also be
thought of as a cross between a DC electric motor and a solenoid. As
each coil is energized in turn, the rotor aligns itself with the magnetic
field produced by the energized field winding. Unlike a synchronous
motor, in its application, the motor may not rotate continuously;
instead, it "steps" from one position to the next as field windings are
energized and de-energized in sequence. Depending on the sequence,
the rotor may turn forwards or backwards.
Simple stepper motor drivers entirely energize or entirely de-energize
the field windings, leading the rotor to "cog" to a limited number of
positions; more sophisticated drivers can proportionally control the
power to the field windings, allowing the rotors to position between the
cog points and thereby rotate extremely smoothly. Computer controlled
stepper motors are one of the most versatile forms of positioning
systems, particularly when part of a digital servo-controlled system.
Stepper motors can be rotated to a specific angle with ease, and hence
stepper motors are used in pre-gigabyte era computer disk drives,
where the precision they offered was adequate for the correct
positioning of the read/write head of a hard disk drive. As drive density
increased, the precision limitations of stepper motors made them
obsolete for hard drives, thus newer hard disk drives use read/write
head control systems based on voice coils.
Stepper motors were upscaled to be used in electric vehicles under the
term SRM (switched reluctance machine).
Linear motors

A linear motor is essentially an electric motor that has been "unrolled"


so that, instead of producing a torque (rotation), it produces a linear
force along its length by setting up a traveling electromagnetic field.
Linear motors are most commonly induction motors or stepper motors.
You can find a linear motor in a maglev (Transrapid) train, where the
train "flies" over the ground, and in many roller-coasters where the
rapid motion of the motorless railcar is controlled by the rail.

Doubly-fed electric motor

Doubly-fed electric motors have two independent multiphase


windings that actively participate in the energy conversion process
with at least one of the winding sets electronically controlled for
variable speed operation. Two is the most active multiphase winding
sets possible without duplicating singly-fed or doubly-fed categories in
the same package. As a result, doubly-fed electric motors are
machines with an effective constant torque speed range that is twice
synchronous speed for a given frequency of excitation. This is twice the
constant torque speed range as singly-fed electric machines, which
have only one active winding set.
A doubly-fed motor allows for a smaller electronic converter but the
cost of the rotor winding and slip rings may offset the saving in the
power electronics components. Difficulties with controlling speed near
synchronous speed limit applications.[4]

Singly-fed electric motor

Singly-fed electric machines incorporate a single multiphase winding


set that is connected to a power supply. Singly-fed electric machines
may be either induction or synchronous. The active winding set can be
electronically controlled. Induction machines develop starting torque at
zero speed and can operate as standalone machines. Synchronous
machines must have auxiliary means for startup, such as a starting
induction squirrel-cage winding or an electronic controller. Singly-fed
electric machines have an effective constant torque speed range up to
synchronous speed for a given excitation frequency.
The induction (asynchronous) motors (i.e., squirrel cage rotor or wound
rotor), synchronous motors (i.e., field-excited, permanent magnet or
brushless DC motors, reluctance motors, etc.), which are discussed on
the this page, are examples of singly-fed motors. By far, singly-fed
motors are the predominantly installed type of motors.

Nanotube nanomotor
Researchers at University of California, Berkeley, recently developed
rotational bearings based upon multiwall carbon nanotubes. By
attaching a gold plate (with dimensions of the order of 100nm) to the
outer shell of a suspended multiwall carbon nanotube (like nested
carbon cylinders), they are able to electrostatically rotate the outer
shell relative to the inner core. These bearings are very robust; devices
have been oscillated thousands of times with no indication of wear.
These nanoelectromechanical systems (NEMS) are the next step in
miniaturization that may find their way into commercial aspects in the
future.
See also:

• Molecular motors
• Electrostatic motor

Materials

There is an impending shortage of many rare raw materials used in the


manufacture of hybrid and electric cars (Nishiyama 2007) (Cox 2008).
For example, the rare earth element dysprosium is required to
fabricate many of the advanced electric motors used in hybrid cars
(Cox 2008). However, over 95% of the world's rare earth elements are
mined in China (Haxel et al. 2005), and domestic Chinese consumption
is expected to consume China's entire supply by 2012 (Cox 2008). [citation
needed]

A few non-Chinese sources such as Thor Lake and Hoidas Lake in


Canada, as well as Mt Weld in Australia are currently under
development (Lunn 2006). However, it is not known if they will be
online in time to supply sufficient production by the time shortage
hits.[citation needed]

Motor standards

The following are major design and manufacturing standards covering


electric motors:

• International Electrotechnical Commission: IEC 60034 Rotating Electrical


Machines
• National Electrical Manufacturers Association (USA): NEMA MG 1 Motors and
Generators
• Underwriters Laboratories (USA): UL 1004 - Standard for Electric Motors

Power factor

The power factor of an AC electric power system is defined as the


ratio of the real power to the apparent power, and is a number
between 0 and 1 (frequently expressed as a percentage, e.g. 0.5 pf =
50% pf). Real power is the capacity of the circuit for performing work in
a particular time. Apparent power is the product of the current and
voltage of the circuit. Due to energy stored in the load and returned to
the source, or due to a non-linear load that distorts the wave shape of
the current drawn from the source, the apparent power can be greater
than the real power. Low-power-factor loads increase losses in a power
distribution system and result in increased energy costs.
Power factor in linear circuit
Instantaneous and average power calculated from AC voltage and current with a unity
power factor (φ=0, cosφ=1)
Instantaneous and average power calculated from AC voltage and current with a zero
power factor (φ=90, cosφ=0)
Instantaneous and average power calculated from AC voltage and current with a lagging
power factor (φ=45, cosφ=0.71)
In a purely resistive AC circuit, voltage and current waveforms are in
step (or in phase), changing polarity at the same instant in each cycle.
Where reactive loads are present, such as with capacitors or inductors,
energy storage in the loads result in a time difference between the
current and voltage waveforms. This stored energy returns to the
source and is not available to do work at the load. Thus, a circuit with a
low power factor will have higher currents to transfer a given quantity
of real power than a circuit with a high power factor.
Circuits containing purely resistive heating elements (filament lamps,
strip heaters, cooking stoves, etc.) have a power factor of 1.0. Circuits
containing inductive or capacitive elements (lamp ballasts, motors,
etc.) often have a power factor below 1.0. For example, in electric
lighting circuits, normal power factor ballasts (NPF) typically have a
value of (0.4 - 0.6). Ballasts with a power factor greater than (0.9) are
considered high power factor ballasts (HPF).
The significance of power factor lies in the fact that utility companies
supply customers with volt-amperes, but bill them for watts. Power
factors below 1.0 require a utility to generate more than the minimum
volt-amperes necessary to supply the real power (watts). This
increases generation and transmission costs. For example, if the load
power factor were as low as 0.7, the apparent power would be 1.4
times the real power used by the load. Line current in the circuit would
also be 1.4 times the current required at 1.0 power factor, so the losses
in the circuit would be doubled (since they are proportional to the
square of the current). Alternatively all components of the system such
as generators, conductors, transformers, and switchgear would be
increased in size (and cost) to carry the extra current.
Utilities typically charge additional costs to customers who have a
power factor below some limit, which is typically 0.9 to 0.95. Engineers
are often interested in the power factor of a load as one of the factors
that affect the efficiency of power transmission.

Definition and calculation


AC power flow has the three components: real power (P), measured in
watts (W); apparent power (S), measured in volt-amperes (VA); and
reactive power (Q), measured in reactive volt-amperes (VAr).
The power factor is defined as:
.
In the case of a perfectly sinusoidal waveform, P, Q and S can be
expressed as vectors that form a vector triangle such that:
If is the phase angle between the current and voltage, then the power
factor is equal to , and:
Since the units are consistent, the power factor is by definition a
dimensionless number between 0 and 1. When power factor is equal to
0, the energy flow is entirely reactive, and stored energy in the load
returns to the source on each cycle. When the power factor is 1, all the
energy supplied by the source is consumed by the load. Power factors
are usually stated as "leading" or "lagging" to show the sign of the
phase angle, where leading indicates a negative sign.
If a purely resistive load is connected to a power supply, current and
voltage will change polarity in step, the power factor will be unity (1),
and the electrical energy flows in a single direction across the network
in each cycle. Inductive loads such as transformers and motors (any
type of wound coil) consume reactive power with current waveform
lagging the voltage. Capacitive loads such as capacitor banks or buried
cable generate reactive power with current phase leading the voltage.
Both types of loads will absorb energy during part of the AC cycle,
which is stored in the device's magnetic or electric field, only to return
this energy back to the source during the rest of the cycle.
For example, to get 1 kW of real power, if the power factor is unity, 1
kVA of apparent power needs to be transferred (1 kW ÷ 1 = 1 kVA). At
low values of power factor, more apparent power needs to be
transferred to get the same real power. To get 1 kW of real power at
0.2 power factor, 5 kVA of apparent power needs to be transferred
(1 kW ÷ 0.2 = 5 kVA). This apparent power must be produced and
transmitted to the load in the conventional fashion, and is subject to
the usual distributed losses in the production and transmission
processes.
It is often possible to adjust the power factor of a system to very near
unity. This practice is known as power factor correction and is achieved
by switching in or out banks of inductors or capacitors. For example the
inductive effect of motor loads may be offset by locally connected
capacitors. When reactive elements supply or absorb reactive power
near the point of reactive loading, the apparent power draw as seen by
the source is reduced and efficiency is increased. The reactive
elements can create voltage fluctuations and harmonic noise during
connection and disconnection procedures, and they will supply or sink
reactive power regardless of whether there is a corresponding load
operating nearby, increasing the system's no-load losses. In a worst
case, reactive elements can interact with the system and with each
other to create resonant conditions, resulting in system instability and
severe overvoltage fluctuations. As such, reactive elements cannot
simply be applied at will, and power factor correction is normally
subject to engineering analysis.

Non-sinusoidal components
In circuits having only sinusoidal currents and voltages, the power
factor effect arises only from the difference in phase between the
current and voltage. This is narrowly known as "displacement power
factor". The concept can be generalized to a total, distortion, or true
power factor where the apparent power includes all harmonic
components. This is of importance in practical power systems which
contain non-linear loads such as rectifiers, some forms of electric
lighting, electric arc furnaces, welding equipment, switched-mode
power supplies and other devices.
A particularly important example is the millions of personal computers
that typically incorporate switched-mode power supplies (SMPS) with
rated output power ranging from 250 W to 750 W. Historically, these
very-low-cost power supplies incorporated a simple full-wave rectifier
that conducted only when the mains instantaneous voltage exceeded
the voltage on the input capacitors. This leads to very high ratios of
peak-to-average input current, which also lead to a low distortion
power factor and potentially serious phase and neutral loading
concerns.
Regulatory agencies such as the EU have set harmonic limits as a
method of improving power factor. Declining component cost has
hastened acceptance and implementation of two different methods.
Normally, this is done by either adding a series inductor (so-called
passive PFC) or adding a boost converter that forces a sinusoidal input
(so-called active PFC). For example, SMPS with passive PFC can
achieve power factor of about 0.7–0.75, SMPS with active PFC, up to
0.99 power factor, while a SMPS without any power factor correction
has a power factor of only about 0.55–0.65.
To comply with current EU standard EN61000-3-2, all switched-mode
power supplies with output power more than 75 W must include
passive PFC, at least. 80 PLUS power supply certification requires a
power factor of 0.9 or more.[1]
A typical multimeter will give incorrect results when attempting to
measure the AC current drawn by a non-sinusoidal load and then
calculate the power factor. A true RMS multimeter must be used to
measure the actual RMS currents and voltages (and therefore apparent
power). To measure the real power or reactive power, a wattmeter
designed to properly work with non-sinusoidal currents must be used.
Measuring power factor
Power factor in a single-phase circuit (or balanced three-phase circuit)
can be measured with the wattmeter-ammeter-voltmeter method,
where the power in watts is divided by the product of measured
voltage and current. The power factor of a balanced polyphase circuit
is the same as that of any phase. The power factor of an unbalanced
polyphase circuit is not uniquely defined.
A direct reading power factor meter can be made with a moving coil
meter of the electrodynamic type, carrying two perpendicular coils on
the moving part of the instrument. The field of the instrument is
energized by the circuit current flow. The two moving coils, A and B,
are connected in parallel with the circuit load. One coil, A, will be
connected through a resistor and the second coil, B, through an
inductor, so that the current in coil B is delayed with respect to current
in A. At unity power factor, the current in A is in phase with the circuit
current, and coil A provides maximum torque,driving the instrument
pointer toward the 1.0 mark on the scale. At zero power factor, the
current in coil B is in phase with circuit current, and coil B provides
torque to drive the pointer towards 0. At intermediate values of power
factor, the torques provided by the two coils add and the pointer takes
up intermediate positions.[2]
Another electromechanical instrument is the polarized-vane type. [3] In
this instrument a stationary field coil produces a rotating magnetic
field (connected either directly to polyphase voltage sources or to a
phase-shifting reactor if a single-phase application). A second
stationary field coil carries a current proportional to current in the
circuit. The moving system of the instrument consists of two vanes
which are magnetized by the current coil. In operation the moving
vanes take up a physical angle equivalent to the electrical angle
between the voltage source and the current source. This type of
instrument can be made to register for currents in both directions,
givng a 4-quadrant display of power factor or phase angle.
Digital instruments can be made that either directly measure the time
lag between voltage and current waveforms and so calculate the
power factor, or by measuring both true and apparent power in the
circuit and calculating the quotient. The first method is only accurate if
voltage and current are sinusoidal; loads such as rectifiers distort the
waveforms from the sinusoidal shape.

Mnemonics
English-language power engineering students are advised to
remember: "ELI the ICE man" or "ELI on ICE" – the voltage E leads the
current I in an inductor L, the current leads the voltage in a capacitor
C.
Or even shorter: CIVIL – in a Capacitor the I (current) leads Voltage,
Voltage leads I (current) in an inductor L.

Principle of operation and comparison to synchronous motors

A 3-phase power supply provides a rotating magnetic field in an induction motor.


The basic difference between an induction motor and a synchronous
AC motor is that in the latter a current is supplied onto the rotor. This
then creates a magnetic field which, through magnetic interaction,
links to the rotating magnetic field in the stator which in turn causes
the rotor to turn. It is called synchronous because at steady state the
speed of the rotor is the same as the speed of the rotating magnetic
field in the stator.
By way of contrast, the induction motor does not have any direct
supply onto the rotor; instead, a secondary current is induced in the
rotor. To achieve this, stator windings are arranged around the rotor so
that when energised with a polyphase supply they create a rotating
magnetic field pattern which sweeps past the rotor. This changing
magnetic field pattern can induce currents in the rotor conductors.
These currents interact with the rotating magnetic field created by the
stator and the rotor will turn.
However, for these currents to be induced, the speed of the physical
rotor and the speed of the rotating magnetic field in the stator must be
different, or else the magnetic field will not be moving relative to the
rotor conductors and no currents will be induced. If by some chance
this happens, the rotor typically slows slightly until a current is re-
induced and then the rotor continues as before. This difference
between the speed of the rotor and speed of the rotating magnetic
field in the stator is called slip. It is unitless and is the ratio between
the relative speed of the magnetic field as seen by the rotor to the
speed of the rotating field. Due to this an induction motor is sometimes
referred to as an asynchronous machine.
Types:

1. Based on type of phase supply


1. three phase induction motor (self starting in nature)
2. single phase induction motor (not self starting)
2. Other
1. Squirrel cage induction motor
2. Slip ring induction motor

Formulas

The relationship between the supply frequency, f, the number of pole


pairs, p, and the synchronous speed, n, is given by f = p*n.
From this relationship:
Speed of rotating field (n) = f/P (revs.s-1)
Speed of rotor = n(1-S) (rev.s-1)
where S is the slip.
Slip is calculated using:
% slip = (n - r) / n * 100
where r is the rotor speed
In contrast, a synchronous motor always runs at either a constant
speed N=(120f)/P or zero.
N = speed of synchronous motor
P = number of poles
f = frequency

Construction

The stator consists of wound 'poles' that carry the supply current that
induces a magnetic field in the conductor. The number of 'poles' can
vary between motor types but the poles are always in pairs (i.e. 2,4,6
etc). There are two types of rotor:

1. Squirrel-cage rotor
2. Slip ring rotor

The most common rotor is a squirrel-cage rotor. It is made up of bars of


either solid copper (most common) or aluminum that span the length
of the rotor, and are connected through a ring at each end. The rotor
bars in squirrel-cage induction motors are not straight, but have some
skew to reduce noise and harmonics.
The motor's phase type is one of two types:

1. Single-phase induction motor


2. 3-phase induction motor

Speed control

The rotational speed of the rotor is controlled by the number of pole


pairs (number of windings in the stator) and by the frequency of the
supply voltage. Before the development of cheap power electronics, it
was difficult to vary the frequency to the motor and therefore the uses
for the induction motor were limited.
There are various techniques to produce a desired speed. The most
commonly used technique is PWM (Pulse Width Modulation), in which a
DC signal is switched on and off very rapidly, producing a sequence of
electrical pulses to the inductor windings. The duty cycle of the pulses,
also known as the mark-space ratio, determines the average power
input to the motor. For example, a 100 V DC signal that is cut into on-
and off- pulses of equal width, has an average voltage of 50 V. If the
on- pulses are a third of the duration of the off pulses, the average
would be 25 V. The frequency of the pulses determines the motor
speed.
The general term for a power electronic device that controls the speed
as well as other parameters is called an 'inverter'. A typical unit will
take the mains AC supply, rectify and smooth it into a "link" DC
voltage, and, by using the method described above, converts it into
the desired AC waveform.
Because the induction motor has no brushes and is easy to control,
many older DC motors are being replaced with induction motors and
accompanying inverters in industrial applications.

Starting of induction motor

In a three phase induction motor, the induced emf in the rotor circuit
depends on the slip of the induction motor and the magnitude of the
rotor current depends upon this induced emf (electromotive force).
When the motor is started, the slip is equal to 1 as the rotor speed is
zero, so the induced emf in the rotor is large. As a result, a very high
current flows through the rotor. This is similar to a transformer with the
secondary coil short circuited, which causes the primary coil to draw a
high current from the mains. Similarly, when an induction motor starts,
a very high current is drawn by the stator, on the order of 5 to 9 times
the full load current. This high current can damage the motor windings
and because it causes heavy line voltage drop, other appliances
connected to the same line may be affected by the voltage fluctuation.
To avoid such effects, the starting current should be limited. A soft start
starter is a device which limits the starting current by providing
reduced voltage to the motor. Once the rotor speed increases, the full
rated voltage is given to it.

Types of starters

1. Direct on line starter


2. Autotransformer starter
3. Star Delta starter
4. Stator Resistance starter

Вам также может понравиться