Introduction To ElectricalMachine
Introduction To ElectricalMachine
COLLEGE OF ENGINEERING
Objectives
1
Chapter One
Field properties, materials, saturation & hysteresis, magnetic circuits, iron loses, Production of
an EMF and Production of electromagnetic force-torque.
1.1 Background
Magnetism plays an integral part in almost every electrical device used today in industry, research,
or the home. Generators, motors, transformers, circuit breakers, televisions, computers, tape
recorders, and telephones all employ magnetic effects to perform a variety of important tasks.
The dynamic age of electricity began with the work of Hans Christian Oersted (1771-1851), who
demonstrated in the year 1819 that a current-carrying conductor produced a magnetic field. This
was the first time that a relationship was shown to exist between electricity and magnetism. His
discovery set off a chain of experiments all across Europe which culminated in the discovery by
Michael Faraday (1791-1867) of his law of electromagnetic induction in 1831.Faraday showed
that it was possible to produce an electric current by means of a magnetic field. This led; in a very
short time, to the development of electrical generators, motors, and transformers, and opened up
our modern electrical era.
All electromagnetic devices make use of magnetic fields in their operation. These magnetic fields
may be produced by permanent magnets or electromagnets. Magnetic fields are created by
alternating- and direct-current sources to provide the necessary medium for developing generator
action and motor action. Throughout this book we will be studying the application of magnetic
fields to electromechanical energy conversion processes as demonstrated in rotating electric
machinery. Also, transformers provide energy transfer from one electric circuit to another via the
changing magnetic field. It will become apparent that there is both transfer and storage of energy
in the magnetic fields of the various electromagnetic devices. Hence all electromagnetic devices
are constructed with appropriate magnetic circuits.
2
1.2 Magnetic Field
The oldest magnetic instrument is a suspended permanent magnet, called a compass. We can define
a magnetic field as a region in space in which a compass needle is acted upon. In a region where
there are no large magnetic objects, the compass needle points in a general north-south longitudinal
direction, with the "north" pole of the compass pointing to the earth's north magnetic pole.
However, we know that similar to the law of electric charges, unlike magnetic poles attract and
like magnetic poles repel. In spite of the fact that the attracting poles of the compass and earth
must be of opposite magnetic polarity, this north-seeking pole of the compass is defined as the
north pole. Similarly, it would be correct to describe the other (unmarked) pole of the compass as
the south-seeking pole. For brevity this pole is called the south pole.
It is well known that a bar of iron can be magnetized by placing it in contact with a strong magnet.
By observing the direction of the compass needle at many points around the magnetized bar, a map
of the magnetic field can be traced. A map of these lines can be obtained by the familiar method
of sprinkling iron filings on a sheet of paper held over the magnetized bar. When this is done, the
pattern of Fig. 1-1 is produced.
The map of Fig.1-1 should not be interpreted too literally. The iron filings are just a local
manifestation of the direction of the magnetic field at that point in apace. Each particle of iron has
in effect become a small magnet and is aligned with the magnetic field of the larger magnet (the
magnetized bar). Although this map seems to show "lines of force." the lines do not actually exist
in space. They can, however, be conceptualized and treated as if they had physical reality. This
visualization of magnetic lines of force which was developed by Faraday will be of great value in
our understanding of electromagnetic principles.
Property 1. Magnetic lines of force are directed from north to south outside a magnet. The direction
is determined by the north pole of a small magnet held in the field.
3
Figure 1.1: Magnetic field pattern near a magnet
Property 3. Magnetic lines of force enter or leave a magnetic surface at right angles.
Property 5. Magnetic lines of force in the same direction tend to repel each other.
These properties can be seen in the field map of Fig. 1.2. Because of properties 1 and 2, we must
assume that there is a magnetic field within the bar; the direction of the field is south to north. The
lines are perpendicular to the magnetic surface because of property 3. The lines spread out because
of properties 4 and 5, and they assume their shape because of the interaction of properties 4, 5, 6,
and 7. Because of property 6, if the lines "find" an easy magnetic path (e.g., through iron), they
will prefer this to a more difficult path-through air, as seen in Fig. 1.2, which is the field map
around a magnet when a piece of iron is brought near it. The iron bar is an "easier" path than the
air, hence the lines tend to concentrate around this part of the circuit. We say that the reluctance of
the iron is less than the reluctance of air, hence the iron is an easier path for the flux lines.
4
Reluctance of a magnetic circuit may be described as magnetic resistance which tends to oppose
the establishment of magnetic flux lines.
A magnetic field is always associated with a current-carrying conductor, as illustrated in Fig. 1.3.
Exploring the magnetic field by means of a compass, we observe the following:
3. If we reverse the direction of current flow, the direction of the magnetic field also changes.
4. The field is strongest near the wire and decreases as we move farther from it. (We can
obtain a measure of field strength by trying to deflect the magnet needle from the position
it has assumed in the field. At a point where the field is strong, it will be more difficult to
deflect it than at a point where it is weak.)
5. If we look at a single current-carrying conductor end on, and draw it as in Fig 1.3, where
the symbol indicates current flowing into the page, it is easier to draw the magnetic field.
If we reverse the current, we have the symbol for current coming out of the page, and we
have the situation depicted in Fig. 1.3. The dot and cross symbols, respectively, represent
the head and tail of an arrow.
6. If we grasp the conductor with our right hand, the thumb pointing in the direction of the
current, our fingers will point in the same direction as the north pole of the compass. This
5
method of determining the directions of current flow in a conductor and the surrounding
lines of force is called Ampere's right-hand rule as illustrated in Fig. 1.4.
Current-carrying conductor
Typical magnetization or 𝐵-𝐻 curves for sheet steel, cast steel, and cast iron are plotted in Fig. 1.5.
The nonlinear relationship between magnetic flux density 𝐵 (Tesla) and magnetic field intensity
𝐻 (ampere-turns per meter) is illustrated. It is observed that the magnetic flux density increases
almost linearly with an increase in the magnetic field intensity up to the knee of the magnetization
curve. Beyond the knee, a continued increase in the magnetic field intensity results in a relatively
small increase in the magnetic flux density. When ferromagnetic materials experience only a slight
increase in magnetic flux density for a relatively large increase in magnetic field intensity, the
materials are said to be saturated. Magnetic saturation occurs beyond the knee of the
magnetization curve.
6
The characteristic of saturation is present only in ferromagnetic materials. An explanation of
magnetic saturation is based on the theory that magnetic materials are composed of very many tiny
magnets (magnetic domains) that are randomly positioned when the material is totally
demagnetized. Upon application of a magnetizing force (𝐻), the tiny magnets will tend to align
themselves in the direction of this force. In the lower part of the magnetizing curve, the alignment
of the randomly positioned tiny magnets increases proportionately to the magnetic field intensity
until the knee of the curve is reached. Beyond the knee of the curve, fewer tiny magnets remain to
be aligned, and therefore large increases in the magnetic field intensity result in only small
increases in magnetic flux density. When there are no more tiny magnets to be aligned, the
ferromagnetic material is completely saturated. In the saturation region of the curve, the magnetic
flux density increases linearly with magnetic field intensity, just as it does for free space or
nonmagnetic materials. From the origin of the 𝐵-𝐻 curve there is a slight concave curvature beyond
which is the essentially linear region. We shall see that the nonlinear characteristics of the
magnetization curve have practical implications in the operation of electrical machines.
1.4 Hysteresis
Hysteresis is the name given to the "lagging" of flux density 𝐵 behind the magnetizing force 𝐻.
when a specimen of ferromagnetic material is taken through a cycle of magnetization. If the
specimen has been completely demagnetized and the magnetizing force H is increased in steps
from zero, the relationship between flux density 𝐵 and 𝐻 is represented by the curve OAC (Fig.
1.6) which is the normal magnetization curve. If the value of 𝐻 is now decreased, the trace of 𝐵 is
higher than OC and follows the curve CD until 𝐻 is reduced to zero. Thus, when 𝐻 reaches zero,
there is a residual flux density referred to as remnant flux density denoted by Br. In order to reduce
𝐵 to zero, a negative field strength OE must be applied. The magnetic field intensity OE required
to wipe out the residual magnetism Br is called coercive force. As 𝐻 is further increased in the
negative direction, the specimen becomes magnetized with the opposite polarity as shown by the
curve EF. If 𝐻 is varied backwards from LO to OK, the flux density curve follows a path FGC,
which is similar to the curve CDEF. The closed loop CDEFGC thus traced out is called the
hysteresis loop of the specimen. The term remnant flux density Br is also called retentivity and the
term coercive force is often called coercivity.
7
The shape of the hysteresis loop will depend upon the nature of magnetic material. Steel alloyed
with 4 % silicon has a very narrow hysteresis loop.
Hysteresis in magnetic materials results in dissipation of energy, which is proportional to the area
of the hysteresis loop. Hence the following conclusions can be drawn:
iii Energy loss is proportional to the area of hysteresis loop and depends upon the quality of
the magnetic material.
The "quantity of magnetism" which exists in a magnetic field is the magnetic line of force, or more
simply, the magnetic flux. In the SI system magnetic flux is measured in units called Weber,
abbreviated Wb, and its symbol is (the Greek lowercase letter phi). The weber is defined in terms
of an induced voltage, so that the definition of the unit will be postponed until we study
electromagnetic induction. Although there is no actual flow of magnetic flux, we will consider
flux to be analogous to current in electric circuits.
The total magnetic flux that comes out of the magnet is not uniformly distributed, as can be seen
in Fig. 1.2. A more useful measure of the magnetic effect is the magnetic flux density, which is
the magnetic flux per unit cross-sectional area. We will consider two equal areas through which
the magnetic flux penetrates at right angles near one end of the permanent magnet along its
centerline. From the illustration it becomes apparent that there is a greater amount of magnetic
8
flux passing through an area that is nearer the magnet pole. In other words, the magnetic flux
density increases as we approach closer to the end of the magnet. However, it must be noted that
the magnetic flux density inside the magnet is uniformly constant. Magnetic flux density is
measured in units of tesla (T) and is given the symbol 𝐵. One tesla is equal to 1 weber of magnetic
flux per square meter of area. We can state that
𝜙
𝐵= (1.1)
𝐴
ϕ = magnetic flux, Wb
Note that this magnetic flux density exists only at the immediate end of the magnet. As we move
away from the end of the magnet, the magnetic flux spreads out, and therefore the magnet flux
density decreases.
We have seen that an increase in the magnitude of current in a coil or a single conductor result in
an increase in the magnetic flux. If the number of turns in a coil are increased (with the current
remaining constant), there is an increase in magnetic flux. Therefore, the magnetic flux is
proportional to the products of amperes and turns. This ability of a coil to produce magnetic flux
is called the magnetomotive force. Magnetomotive force is abbreviated MMF and has the units of
ampere-turns (At). The magnetomotive force is given the symbol Fm. Strictly speaking, the units
of MMF are amperes because turns are dimensionless quantities. However, from a pedagogical
standpoint, we prefer and shall use throughout this book the units of ampere-turns (At) for MMF.
We may write
𝐹𝑚 = 𝑁𝐼 (1.2)
9
Magnetomotive force in the magnetic circuit is analogous to electromotive force in an electric
circuit.
we have seen that doubling the driving force (MMF) in the circuit results in a doubling of the
output quantity (magnetic flux). We consider this ratio of MMF to magnetic flux:
Fm
ℜm = (1.3)
ϕ
ϕ = magnetic flux, Wb
Transposing, we have
𝐹𝑚 = ℜ𝑚ϕ
which shows us that the magnetic flux is directly proportional to the magnetomotive force. This
equation represents Ohm's law of magnetic circuits. The proportionality factor ℜ𝑚, is called the
reluctance of the magnetic circuit and is obviously, analogous to resistance in an electric circuit.
Assuming that a coil has fixed turns and a constant excitation current, the amount of magnetic flux
produced will depend on the material used in the core of the coil. A much larger amount of flux
can be produced in an iron-core coil than in an air-core coil. Thus, we see that the reluctance of
the magnetic circuit depends on the material properties of the magnetic circuit. For our purposes,
the materials are classified as either magnetic or nonmagnetic. Only the ferrous (irons and steels)
group of metals, including cobalt and nickel, are magnetic materials. All other materials, such as
air, insulators, wood, paper, plastic, brass, and bronze, including vacuum, are nonmagnetic
materials. The strength and pattern of the magnetic field in nonmagnetic materials would be
identical to that of air or vacuum (free space). In our discussions we will assume that the magnetic
properties of air and vacuum are the same. We consider some of the peculiar characteristics of
magnetic materials in subsequent sections.
10
The reluctance of a homogeneous magnetic circuit may be expressed in terms of its physical
dimensions and magnetic property as follows:
𝑙
ℜm = Aµ (1.4)
Reluctance is in essence magnetic resistance, that is, the property of a magnetic circuit which is
reluctant or unwilling to set up magnetic flux. The reciprocal of reluctance is termed as permeance,
which is analogous to conductance in electric circuits.
Permeability(µ)
Permeability is the magnetic property that determines the characteristics of magnetic materials and
nonmagnetic materials. The permeability of free space and nonmagnetic materials has the
following symbol and constant value in SI units:
μ0 = 4π10−7 H/m
As we can see, the reluctance of magnetic materials μ𝑟 is much lower than that of air or
nonmagnetic materials μ0 . From the inverse relationship of reluctance and permeability, we
determine that the total permeability of magnetic materials is much greater than that of air.
However, the value of permeability varies with the degree of magnetization of the magnetic
material and, of course, the type of material. Since the permeability of magnetic materials μ𝑟 is
variable, we must employ magnetic saturation (𝐵-𝐻) curves to perform magnetic circuit
calculations. Permeability in magnetic circuits is somewhat analogous to conductivity in electric
circuits.
Example 1.3 In Fig. 1.7 below we assume that the magnetic flux is practically uniform in the
cross-sectional area of the toroid. The mean path length is 0.314m and the cross-sectional area
through which the flux exists is 78.5 x 10-6 m2. Calculate the number of ampere-turns required to
set up magnetic flux of 1Wb.
11
Figure:1.7 Toroid coil
Solution:
𝑙 0.314
ℜm = = = 3.18109 At/W𝑏
Aµ 4π10 78.510−6
−7
𝐹𝑚 = ℜ𝑚Φ=3.18109 1=3.18109 𝐴𝑡
This is obviously a very large number and we may conclude that the path reluctance is very high.
This means that it is comparatively difficult to establish a large magnetic flux in air. For this reason,
when we need high flux densities, it becomes necessary to use materials having high values of
permeability (such as iron or steel) for large portions of the magnetic paths.
One other important magnetic quantity is the magnetomotive force gradient per unit length of
magnetic circuit, or more commonly, the magnetic field intensity. Its symbol is 𝐻 and from the
definition,
ℱ𝑚
𝐻= (1.5)
𝑙
the unit is ampere-turns per meter (At/m). The former name for magnetic field intensity was
magnetizing force. We have seen that more ampere-turns (MMF) are required to set up the same
magnetic flux in magnetic circuits of air than in iron of similar configuration. Hence the magnetic
field intensity for the air path is much larger than for the iron path. In the toroid of Fig. 1.7, a
magnetomotive force of 10 At acts along the mean path of 0.314 m. The magnetic field intensity
𝑁𝐼 10
is 𝐻= = 0.314=31.8At/m
𝑙
12
Eq.(1.5) transposed, Hl = NI is one form of Ampere's circuital law applied to a simple magnetic
circuit. Magnetic field intensity in magnetic circuits is analogous to potential or voltage gradient
in electric circuits.
We can derive a useful relationship for magnetic circuits by summarizing the equations developed
so far.
ϕ
𝐵= , ℱ𝑚 = 𝑁𝐼
𝐴
ℱm 𝑙
ϕ = ℜm , ℜm = Aµ
ℱ𝑚
𝐻=
𝑙
ϕ ℱ𝑚 Aµ𝑙𝐻
𝐵= = ℜm𝐴 = = µ𝐻
𝐴 𝐴𝑙
𝐵 = µ𝐻 (1.6)
Eq.(1.6) shows that the magnetic flux density is directly dependent on both permeability and
magnetic field intensity. Only in air or free space is the permeability μ0 constant , and thus a linear
relationship between 𝐵 and 𝐻 exists. In the next section we consider ferromagnetic materials in
which the absolute permeability is not a constant but depends on the degree of magnetization.
Transposition of Eq.(1.6) gives the absolute permeability as the ratio of the magnetic flux density
to the corresponding magnetic field intensity:
𝐵
µ=
𝐻
Thus, we can obtain the values of absolute permeability of ferromagnetic materials from the
magnetization (𝐵-𝐻) curves. Another method of obtaining the absolute permeability would be to
take the slope (differential) of the curve at various points. Although the differential method may
be more realistic, for our purposes in this book, the simpler method of ratios to obtain the absolute
permeability will be acceptable.
If we wish to compare the permeability of magnetic materials with that of air, we may use the
relative permeability μ𝑟 , which is defined by the equation
13
µ
μ𝑟 = μ (1.7)
0
μ𝑟 = relative permeability
From the typical magnetization curves of Fig. 1.5, we can calculate the value of absolute and
relative permeabilities for any magnetic operating condition. When we do this, we observe that the
value of relative permeability is not a constant but obtains a maximum value at about the knee of
the 𝐵-𝐻 curve.
Example 1.4 Calculate the absolute and relative permeabilities of cast steel operating at magnetic
flux densities of 0.7 T and 1.0 T.
Solution:
From the saturation curve for cast steel, the values of 𝐻 are 400 At/m and 800 At/m. respectively.
For 0.7T:
0.7
µ = 400 = 1.75 × 10 −3𝐻/𝑚 𝑜𝑟 𝑇/𝐴𝑡/m
For 1.0T:
1
µ = 800 = 1.25 × 10 −3𝐻/𝑚
For 0.7T:
1.75 × 10 −3
μ𝑟 = = 1392.61
4π10−7
For 1.0T:
1.25 × 10 −3
μ𝑟 = = 994.72
4π10−7
Thus, we see that cast steel has at least 1000 times more ability to set up magnetic flux
lines than do nonmagnetic materials.
14
1.6 Magnetic Circuits
If we construct a coil of many turns, we can increase the magnetic field strength very greatly, as
shown in Fig. 1.8. We can also increase the magnetic field strength by increasing the magnitude
of current in the coil. A cylindrical coil closely wound with a large number of turns of insulated
wire is called solenoid. Thus, we see that the magnetic field strength is proportional to both the
number of turns and the current.
We can determine the direction of the magnetic field in a cylindrical coil of many turns of insulated
wire by using our right hand. If we grasp the coil with our right hand with the fingers pointing in
the direction of the current, the thumb will point in the direction of the north pole. This method of
determining directions of current flow in a coil and magnetic fields of force is another form of
Ampere's right-hand rule. Andre Marie Ampere (1775-1836), pursuant to the experimental work
of Oersted, developed extensively the foundations of electromagnetic theory. Refer to Fig 1.8.
Several practical magnetic circuits are illustrated in Figure 1.9.
15
(c) Vertical contactor (d) Electrical Machine
A toroid of homogeneous magnetic material, such as iron or steel, is wound with a fixed number
of turns of insulated wire as shown in Fig. 1.7. The magnetic flux (ϕ) and the excitation current (I)
are related by Eq.(1.6):
ϕ
𝐵 = µ𝐻 = 𝐴
ϕ µ𝑁𝐼
Thus, = that is ϕ = (𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡)µ𝐼
𝐴 𝑙
where the constant is 𝑁𝐼/𝑙. At the outset, the sample of ferromagnetic material in the toroid was
totally demagnetized. In experimental measurements, the excitation current is varied and the
corresponding values of magnetic flux recorded. Then the calculated values of 𝐵 and 𝐻 are plotted
on linear scales as illustrated in Fig. 1.5.
By definition, a series magnetic circuit contains magnetic flux, which is common throughout the
series magnetic elements. These series magnetic elements may consist of composite sectors of
ferromagnetic materials of different lengths and cross-sectional areas, and of air gaps. The simplest
series magnetic circuit would be of a toroid of homogeneous material and the steel core of a
transformer. More complex series circuits which contain air gaps are illustrated in Fig. 1.9.
Parallel magnetic circuits are defined by the number of paths that the magnetic flux may follow.
Any of these paths or branches may consist of composite sectors of magnetic materials, including
air gaps.
In our discussion so far, we note the following analogous relationships between magnetic
quantities and electric quantities:
16
Electric circuit Magnetic circuit
E (volts) Fm (NI ampere-turns)
I (amperes) ϕ(weber)
R (ohms) ℜm(ampere-turns/weber)
ρ= 1/σ(conductivity) µ(henries/meter)
We can draw useful electrical analogs for the solution of magnetic circuit problems. In an electrical
circuit the driving force is the voltage, the output is the current, and the opposition to establishing
current is the resistance. In the same way, the driving force in the magnetic circuit is the
magnetomotive force, the output is the magnetic flux, and opposition to establishing the flux is the
reluctance.
Thus, we have for the magnetic circuit of Fig. 1.10a the analogous electric circuit and the
analogous magnetic circuit in Fig. 1.10b and c, respectively. The iron and air portions of the
magnetic circuit are analogous to the two series resistors of the electric circuit. Analogous to the
electric circuit, the magnetomotive force must overcome the magnetic potential drops of the two
series reluctances in accordance with Kirchhoff's voltage law applied to magnetic circuits.
Therefore,
17
(a) (b) (c)
Figure: 1.10: Iron-core toroid with air gap: (a) Magnetic circuit; (b) analogous electric circuit; (c)
analogous magnetic circuit.
Given the physical parameters of the series magnetic circuit and the value of magnetic flux or
magnetic flux density, the required magnetomotive force can be calculated in a straightforward
manner using Eq. (1.10).
The general principles of electric circuits embodied in Ohm's and Kirchhoff’s laws are applied
as analogous equivalents to parallel magnetic circuits. With the presence of air gaps, most complex
magnetic circuits are solved using the series parallel equivalent analogs. In analogous equivalents,
Kirchhoff's current law for magnetic circuits states that the sum of magnetic fluxes entering a
junction or node is equal to the sum of magnetic fluxes leaving the junction or node. Needless to
a say, magnetic flux must not be perceived as flowing.
In a series magnetic circuit containing an air gap, there is a tendency for the airgap flux to spread
out (i.e., to create a bulge) as shown in Figure 1.11. This spreading effect, termed fringing, reduces
the net flux density in the air gap.
Fringing flux
Useful flux
18
Magnetic Core (Iron) Losses
It will be shown later that the magnetic flux within the armature of dc machines changes direction
as rotation occurs past the magnetic field poles. This change in direction of the armature magnetic
flux is effectively an alternating flux. This results in core losses, which are treated in more detail
in forthcoming chapters. Magnetic core losses consist of hysteresis losses and eddy-current losses.
Core loss: The core loss Pc occurring in the Magnetic core iron, consists of two components,
hysteresis loss Ph and eddy current loss Pe i.e.
Pc = Ph + Pe
The hysteresis and eddy current losses in the core can be expressed by: -
𝑥
𝑃ℎ = 𝐾ℎ 𝑓𝐵𝑚
𝑎𝑛𝑑 }
2 2
𝑃𝑒 = 𝐾𝑒 𝑓 𝐵 𝑚
Where Kh = proportionality constant which depends upon the volume and quality of the core
Ke = Proportionality constant whose value depends on the volume and resistivity of the core
The value of the exponent x (called Steinmetz’s constant) varies from 1.5 to 2.5 depending upon
the magnetic properties of the core material. Therefore, the total core loss is
Pc = KhfB1.6m + Kef2B2m
Ampere turns for various parts of the magnetic circuit will be calculated separately. To calculate
the ampere turns for a particular part, the following procedure is followed in general:
1. The reluctance of the part is calculated using Eq. (1.4) as the case may be.
2. The magnetic flux established in that part is calculated using Eq. (1.10).
3. Cross-sectional area of the part is calculated from the given dimensions.
19
4. Magnetic flux density is found by dividing the flux by the cross-sectional area, i.e.
ϕ
𝐵= 𝐴
5. Ampere turns per meter of the magnetic flux path length in that part at the flux density
calculated above is found by using the magnetization curve for the magnetic material of
that part.
6. Length of the magnetic flux path in that part is estimated from the given dimensions.
7. Total ampere turns for the part are obtained by multiplying ampere turns per meter by the
length of the flux path.
8. General procedure is now applied to various parts of the magnetic circuit.
9. Total ampere turns for the complete magnetic circuit can now be found by adding
algebraically the ampere turns needed by the various parts of the magnetic circuit.
𝑙𝑔
Reluctance for air-gap, ag (for which r=1) = µ
0𝐴𝑔
𝑙𝑔 𝜙 𝑙𝑔 1 1
thus, F𝑎𝑔 = 𝜙 µ = 𝐴 µ = µ 𝐵𝑔 𝑙𝑔 = μ −7
𝐵𝑔 𝑙𝑔 (1.11)
0𝐴𝑔 𝑔 0 0 0 =4π10
Hence to calculate the ampere turns for the air gap, the following general procedure may be
followed:
using Eq. (1.11), calculate the ampere turns needed for the air gap.
Example 1.5 Illustrates the method of solution for a simple one-material series circuit.
20
The circuit of Fig. 1.11 is a magnetic core made of cast steel. A coil of N turns is wound on it. For
a flux of 560µWb, calculate the necessary current, neglecting any fringing effects. The cross-
sectional area A is constant.
56010−6
Solution, 𝐵 = = 14010−2 = 1.4𝑇
410−4
For B = 1.4 T, H = 2200 At/m (from the 𝐵-𝐻 curve of Figure 1.5). The average or mean length
of the magnetic path is 20 + 12 + 20 + 12 cm = 64 = 0.64 m. Therefore,
Hl=NI=22000.64At
22000.64At
𝐼= = 2.56𝐴
550
Fig.1.13a shows a parallel magnetic circuit. There are NI ampere-turns on the center leg. The flux
that is produced by the MMF in the center leg exists in the center leg and then divides into two
parts, one going in the path afe and the other in the path bcd. If we assume for simplicity that afe
= bcd, the flux is distributed evenly between the two paths. Now
Eq (1.12) is actually the analog of Kirchhoff's current law, but now we can say that the amount of
flux entering a junction is equal to the amount of flux leaving the junction.
21
Another observation that we may make on this circuit is that the MMF drops around a circuit are
the same no matter what path we take. Thus, the MMF drop around afe must be equal to the MMF
drop around bcd. This can be stated more precisely as
(a)
(b) (c)
Figure 1-13: Magnetic circuit with center leg: (a) Magnetic circuit; (b) equivalent magnetic
circuit; (c) analogous electric circuit.
The drop in MMF around either path afe or bcd must also be equal to the MMF drop along path g.
But g also has an "active source," the NI ampere-turns of the coil. The actual MMF existing
between X and Y is the driving force NI minus the drop Hglg in path g. Then we can write
22
E - RgIg = Ibcd (Rb + Rc + Rd) (1.15)
In the analogous magnetic circuit, note that NI is drawn in series with Rmg, although physically the
coil surrounds the central magnetic path.
lg = lf = lc = 12 cm
la = lb = le = ld = 14 cm
Aa = Ab = Ac = Ad = Ae = Af = 1 cm2
Ag = 3 cm2
The material is sheet steel. The flux density in the center leg is 0.9 T. Calculate the MMF required
to produce this flux density.
Solution
The total flux in the center leg is 𝜙 = 𝐴𝐵 = 0.9310−4 = 2.710−4 𝑊𝑏. The flux divides into
two parts, the left-hand path through afe and the right-hand path through bcd i.e 𝜙𝑎𝑓𝑒 = 𝜙𝑏𝑐𝑑 =
2.710−4
= 𝑓𝑙𝑢𝑥 𝑖𝑛 𝑡ℎ𝑒 𝑐𝑒𝑛𝑡𝑒𝑟 𝑙𝑒𝑔(𝑔)/2 . The flux density in path g is Bg = 0.9 T and therefore
2
2.710−4
2
𝐵𝑎 = = 1.35𝑇
110−4
and therefore
Ha = 950 At/rn
Ha = Hb = Hc =Hd =He= Hf
Therefore,
23
1.7 Permanent Magnets
Permanent magnets are commonly used as compasses and magnetic lifts. Today, there is a
substantial increase in the application of permanent magnets for electromagnetic devices such as
instruments. magnetic clutches and brakes, loudspeakers and relays, as well as small generators
and motors.
Modern Permanent magnets materials are alloys composed of nickel, aluminum, and iron,
described by the trade name Alnico. Current research has developed rare-earth materials for
permanent magnets having extremely high values of residual flux density. A wide variety of
powdered-composition permanent magnets called ferrites are useful for relatively high-frequency
applications. The composition materials of ferrites are usually barium and ceramic.
Oersted at Copenhagen in 1820 discovered a very important phenomenon giving the relationship
between magnetism and electricity. As per this relationship, a conductor carrying a current i is
surrounded all along its length by a magnetic field, the lines of magnetic flux being concentric
circles in planes at right angles to the conductor. This phenomenon of a magnetic field being
associated with a current carrying conductor lead to the question whether the converse of the above
is possible, i.e. can a magnetic field generate a current? Michael Faraday, on 29 Aug. 1831,
succeeded in generating an electric current with the aid of magnetic flux.
From his experiments, Faraday concluded that a current was generated in a coil so long as the lines
of force bearing through the conductor changed. The current thus generated is called the induced
current and the emf that gives rise to this induced current is called the induced emf. This
phenomenon of generating an induced current in a closed circuit by changing the magnetic field
through it, is called electromagnetic induction. The operation of electrical equipments like motors,
generators, transformers, etc. is mainly based upon the laws formulated by Faraday.
Faraday conducted the following experiment to obtain an electric current with the aid of magnetic
flux.
24
Fig.1.14 shows a coil connected to a galvanometer G. When the magnet was kept inside the coil
nothing happened as shown in Fig. 1.14 (a). But when the North Pole of the magnet was inserted
in the coil as shown in Fig. 1.14(b), the galvanometer pointer was deflected momentarily on one
side and the direction of current was found to be anticlockwise. When the magnet was withdrawn
as shown in Fig. 1.14(c), the pointer of the galvanometer deflect on the other side and the direction
of current was found to be clockwise. Similar results were obtained when the South Pole of the
magnet was inserted or withdrawn, but the direction of current in this case was reverse to that
obtained with the North Pole.
Faraday summed up the results of the experiments described above in the form of following two
laws, known as Faraday's laws of electromagnetic induction.
Faraday's first law: states that whenever the magnetic flux associated or linked with a closed
circuit is changed, or alternatively, when a conductor cuts or is cut by the magnetic flux, an emf is
induced in the circuit resulting in an induced current. This emf is induced so long as the magnetic
flux changes.
Faraday's second law: states that the magnitude of the induced emf generated in a coil is directly
proportional to the rate of change of magnetic flux.
These two basic laws discovered by Faraday changed the course of electrical engineering and led
to the development of generators, transformers, etc.
The change of flux as discussed in the Faraday's laws can be produced in two different ways: (i)
by the motion of the conductor or the coil in a magnetic field, i.e. the magnetic field is stationary
and the moving conductors cut across it. The emf generated in this way is normally called
dynamically induced emf; (ii) by changing the current (either increasing or decreasing) in a circuit.
thereby changing the flux linked with stationary conductors, i.e. the conductors or coils remain
stationary and the flux linking these conductors is changed. The emf is termed statically induced
emf. Statically induced emf can be further subdivided into
The concept of dynamically induced emf gave rise to the development of generators, whereas
statically induced emf was helpful in developing transformers.
25
1.8.1 Direction of Induced emf
In case of dynamically induced emf, Fleming's right-hand rule is used to obtain the direction of
induced emf, whereas Lenz's law is normally used to fix the direction of statically induced emf.
(a) Fleming's Right-Hand Rule: Stretch the forefinger, the middle finger and the thumb of the
right hand in three mutually perpendicular directions as shown in Fig.1.15. If the forefinger
points in the direction of the magnetic flux, the thumb points in the direction of motion of the
conductor relative to the magnetic field, then the middle finger represents the direction of the
induced emf.
(b) Lenz's Law: The direction of statically induced emf can be obtained with the help of Lenz's
law which states: "the direction of the induced emf is always such that it tends to set up a
current opposing the change of flux responsible for producing that emf.
Lenz’s law is further clarified by using it with reference to Fig. 1.14. When the north pole of the
magnet is inserted in the coil, an emf is induced in it due to the motion of the magnet, thereby
generating induced current. According to Lenz's law, the direction of this induced current
generated in the coil should be such that the motion of the magnet is opposed, which is possible
only when the upper end of the coil behaves as a north pole. For this to happen, the current
generated in the coil should be in the anticlockwise direction as was observed by Faraday. In a
similar way, the direction of the induced emf can be determined for any case utilizing Lenz's law.
S N
S N
Motion
Motion
Flux
e Flux
EMF
direction
emf of induced emf
26
1.8.2 Magnitude of Induced Emf in a Coil
Let a coil consist of N number of turns over it. Assume that the flux through the coil changes from
its initial value 𝜙1 to 𝜙2 in an interval t second.
The term flux linkages used over here simply means the product of flux in Weber and the number
of turns with which the flux is linked. Now as per Faraday's laws of electromagnetic induction,
induced emf in the above coil due to a change of flux is given by,
𝑁𝜙1 −𝑁𝜙2
𝐼𝑛𝑑𝑢𝑐𝑒𝑑 𝑒𝑚𝑓 = 𝑉
𝑡
Based on the above, the instantaneous value of emf induced in the coil can be represented as,
𝑑(𝑁𝜙)
𝑒𝑖𝑛𝑑 = − 𝑉 or
𝑑𝑡
𝑁𝑑𝜙
𝑒𝑖𝑛𝑑 = − (1.16)
𝑑𝑡
The negative sign in the Eq.(1.16) above equation signifies that the induced emf generates a current
tending to oppose the increase of flux through the coil. The relation expressed by the above
equation can be called Faraday's law.
Dynamically induced emf is produced by the movement of the conductor in a magnetic field.
Fig.1.16 shows a uniform magnetic field of flux density 𝐵 Tesla, in which the conductor is moving
in the direction shown and cuts the flux at right angles.
27
B
v
A
The dynamically induced emf is the rate of change of flux linkages, i.e.
𝐵𝑙𝑑𝑥 𝑑𝑥
Dynamically induced emf, i.e 𝑒𝑖𝑛𝑑 = = 𝐵𝑙 𝑑𝑡
𝑑𝑡
𝑑𝑥
As =𝑣
𝑑𝑡
When the conductor or coil remains stationary and the flux linking with these conductors or coil
undergo a change, an emf is induced in the conductors. Such an induced emf is termed as statically
induced emf. Statically induced emf can be further classified as (i) self-induced emf and (ii)
mutual induced emf.
i. Self-induced emf
Any electrical circuit in which the change of current is accompanied by the change of flux, and
therefore by an induced emf, is said to be inductive or to possess self-inductance. Thus, the
property of the coil which enables to induce an emf in it whenever the current changes is called
self-induction.
28
termed as flux linkages. Now if the current flowing in the coil is changed, then the number of lines
linking the coil also changes. As such emf is induced in the coil according to Faraday's laws of
electromagnetic induction. This emf is termed as statically self-induced emf or the emf of self-
induction. The phenomenon of self-induction is felt only when the current is changing, either
increasing or decreasing.
As per Faraday's laws of electromagnetic induction, this induced emf is given by,
𝑁𝑑𝜙
𝑒𝑖𝑛𝑑 = − ,V (1.18)
𝑑𝑡
The coil in question is wound on an iron core, whose permeability is constant. Thus, flux is
proportional to the current through the coil, i.e.
ϕ I Or
𝜙
= 𝑐𝑜𝑠𝑡𝑎𝑛𝑡
𝐼
ϕ
i.e I = ϕ
I
Now if current is changed at a certain rate, the flux also changes at the same rate. Thus, the rate
ϕ ϕ
of change of flux = rate of change of current ( I = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡) . Substituting this in Eq. (1.18),
I
𝜙
𝑒 = −𝑁 𝐼 rate of change of current Or,
𝑁𝜙 𝑑𝑖
𝑒=− 𝑑𝑡 (1.19)
𝐼
𝑁𝜙
The term i.e. flux linkages/ampere is generally called the self-inductance of the coil or the
𝐼
coefficient of self-induction and is denoted by a symbol L. With this replacement, Eq, (1.19)
becomes,
𝑑𝑖
𝑒 = −𝐿 𝑑𝑡 (1.20)
𝑁𝜙
Where, 𝐿= 𝐼
henry
29
The negative sign in Eq.(1.20) indicates that it is an emf opposing the change, i.e. if the current is
increasing, this emf will oppose the increase in current (emf will be opposite to the applied
voltage), in case the current is decreasing, the induced emf tends to prevent the decrease of current
and its direction is therefore the same as that of current or the applied voltage. It also indicates that
the energy is being absorbed from the electric circuit and stored as magnetic energy in the coil.
The coefficient of self-induction L of the circuit is thus defined as the magnetic flux linked with
the coil when a unit current flows through it. It is also numerically equal to the induced emf due
to unit rate of change of current in the coil. The practical unit of inductance is henry.
The phenomenon of generation of induced emf in a circuit by changing the current in a neighboring
circuit is called mutual induction. Consider two coils P and S such that P is connected to a cell
through switch K and S to a galvanometer as shown in Fig. 1.18. When the switch K is closed
suddenly to start current in the coil P, the galvanometer gives a sudden "kick" in one direction.
Now when K is opened, the galvanometer again shows a deflection but in the opposite direction.
The above observations indicate clearly that an induced current is set up in the coil S when the
current is changed in the coil P, though the coil S is not connected physically to coil P. Two coils
possessing this property are said to have mutual inductance. The unit of mutual inductance is also
henry. It is denoted by M. Two coils are said to possess a mutual inductance of 1 henry when
current changing at the rate of 1 ampere per second in one coil induces an emf of 1 volt in the
other.
P S
K
G
𝜙2
= 𝐾 Or
𝜙1
𝜙2 = 𝐾𝜙1
Also, ϕ I
30
𝜙
Or = 𝑐𝑜𝑠𝑡𝑎𝑛𝑡
𝐼
𝜙2
Now I = 𝜙2
I
𝐾𝜙1
Or 𝜙2 = current
I
When current is changed at a certain rate, 𝜙2 also changes at the same rate. Thus,
𝐾𝜙1
Rate of change of 𝜙2 = rate of change of current (1.21)
I
According to Faraday's laws of electromagnetic induction the emf induced in S is given by,
𝐾𝜙1
𝑒S = 𝑁2 rate of change of current
I
𝐾𝜙1 𝑑𝑖
Or 𝑒S = 𝑁2 𝑑𝑡
I
𝑑𝑖
𝑒S = 𝑀 𝑑𝑡
𝑁2 𝐾𝜙1 𝜙2
Where, 𝑀= = 𝑁2
I I
The constant M in the above equation, which is equal to the flux linkages of coil S per ampere of
current in coil P, is called the coefficient of mutual induction or mutual inductance.
Hence the coefficient of mutual induction is defined as the number of lines of force passing through
the secondary coil S when unit current changes in the primary coil P. It is also numerically equal
to the induced emf in one circuit due to a unit rate of change of current in the other circuit.
31
electric field, inductors store energy in a magnetic field. While capacitors prevent voltage from
changing instantaneously, inductors, as we shall see, prevent current from changing
instantaneously.
Consider a coil of wire carrying some current creating a magnetic field within the coil. As shown
in Fig. 1.19, if the coil has an air core, the flux can pretty much go where it wants to, which leads
to the possibility that much of the flux will not link all of the turns of the coil. To help guide the
flux through the coil, so that flux leakage is minimized, the coil might be wrapped around a
ferromagnetic bar or ferromagnetic core as shown in Fig.1.20. The lower reluctance path provided
by the ferromagnetic material also greatly increases the flux . We can easily analyze the magnetic
circuit in which the coil is wrapped around the ferromagnetic core in Fig.1.20(a). Assume that all
of the flux stays within the low-reluctance pathway provided by the core, and apply Eq.(1.3):
F 𝑁𝐼
𝜙 = ℜm = ℜ (1.23)
𝑚 𝑚
Figure 1.19 A coil with an air core will have considerable leakage flux.
ϕ
N
i
e
(a) (b)
Figure 1.20: Flux can be increased and leakage reduced by wrapping the coils around a
ferromagnetic material that provides a lower reluctance path.
The flux will be much higher using the core (a) rather than the rod (b).
From Faraday’s law, changes in magnetic flux create a voltage e, called the electromotive force
(emf), across the coil equal to
𝑑𝜙
𝑒 = 𝑁 𝑑𝑡 (1.24)
32
Substituting Eq.(1.23) into (1.24) gives
𝑑 𝑁𝐼 𝑁 2 𝑑𝑖 𝑑𝑖
𝑒 = 𝑁 𝑑𝑡 (ℜ ) = ℜ = 𝐿 𝑑𝑡
𝑚 𝑚 𝑑𝑡
𝑁2
𝐿 = ℜ heneries (1.25)
𝑚
Notice in Fig. 1.20 (a) that a distinction has been made between e, the emf voltage induced across
the coil, and V, a voltage that may have been applied to the circuit to cause the flux in the first
place. If there are no losses in the connecting wires between the source voltage and the coil, then
e = v and we have the final defining relationship for an inductor:
𝑑𝑖
𝑣 = 𝐿 𝑑𝑡
Energy stored
Consider a coil having a constant inductance of L Henry, in which the current increases by 𝑑𝑖 in
𝑑𝑡 seconds, then induced emf in the coil, 𝑒 becomes
𝑑𝑖
𝑒 = −𝐿 𝑑𝑡
The applied voltage must balance the voltage drop across resistor R and neutralize the above
induced emf, thus,
𝑑𝑖
𝑉 = 𝑖𝑅 + 𝐿 (1.26)
𝑑𝑡
where, 𝑉𝑖𝑑𝑡 is the energy supplied by the source in time 𝑑𝑡 𝑖 2 𝑅𝑑𝑡 the energy dissipated in the
form of heat 𝐿𝑖𝑑𝑖 the energy absorbed by the inductance of the coil in building up the magnetic
field.
33
Thus, energy absorbed by the magnetic field during the time 𝑑𝑡 second
= 𝐿𝑖𝑑𝑡 Joules
Hence total energy absorbed by the magnetic field when the current increases from zero to 𝐼
amperes
𝐼 𝐼 1
Energy stored = ∫0 𝐿𝑖𝑑𝑖 = 𝐿 ∫0 𝑖𝑑𝑖 = 2 𝐿𝐼 2 𝑱 (1.28)
Force on a Conductor
Ampere demonstrated in 1820 that there is a magnetic field associated with a conductor carrying
current. When placed in a transverse magnetic field, this conductor experiences a force that is
proportional to
(a) the strength of the magnetic field,
(b) the magnitude of current in the conductor, and
(c) the length of the conductor in, and perpendicular to, the magnetic field.
In SI units, the electromagnetic force developed on the conductor carrying current in a magnetic
field B is given by
34
Chapter Two
Transformer
Contents of this chapter includes:
Principle of action, construction, ideal & practical models, parameter testing, voltage regulation,
efficiency, 3-phase transformers, connection groups.
2.1 Background
The transformer is a static device that transfers electrical energy from one electrical circuit to
another electrical circuit through the medium of magnetic field and without a change in the
frequency. The electric circuit which receives energy from the supply mains is called primary
winding and the other circuit which delivers electrical energy to the load is called secondary
winding.
Actually, the transformer is an electric energy conversion device, since the energy received by the
primary is converted to useful electrical energy in the other circuits (secondary winding circuit).
If the secondary winding has more turns than the primary winding, then the secondary voltage is
higher than the primary voltage and the transformer is called a step-up transformer. When the
secondary winding has less turns than the primary windings then the secondary voltage is lower
than the primary voltage and the transformer is called step down transformer.
Note that a step-up transformer can be used as a step-down transformer, in which the secondary of
step-up transformer becomes the primary of the step-down transformer. Actually, a transformer
can be termed a step-up or step-down transformer only after it has been put into service.
ii Matching source and load impedances for maximum power transfer in electronic and
35
Transformers are used extensively in ac power systems. AC electrical power can be generated at
one central location, its voltage stepped up for transmission over long distances at very low losses
and its voltage stepped down again for final use.
There are basically two types of transformers, the core-type and the shell-type. These two types
differ from each other by the manner in which the windings are wound around the magnetic core.
The magnetic core is a stock of thin silicon-steel laminations about 0.35mm thick for 50Hz
transformers. In order to reduce the eddy current losses, these laminations are insulated from one
another by thin layer of varnish. In the core-type, the windings surround a considerable part of
steel core as shown in Fig. 2.1(a). In shell-type the steel core surrounds a major part of the windings
as shown in Fig. 2.1(b). For a given output and voltage rating, core-type transformer requires less
iron but more conductor material as compared to a shell-type transformer.
The vertical portions of the core are usually called limbs or legs and the top and bottom portions
are called yoke. This means that for single-phase transformers, core-type has two-legged core
whereas shell type has three-legged core. In core-type transformers, most of the flux is confined
to high permeability core. However, some of the flux leaks through the core legs and non-magnetic
material surrounding the core. The flux called leakage flux, links one winding and not the other.
A reduction in this leakage flux is desirable as it improves the transformer performance
considerably. Consequently, an effort is always made to reduce it. In the core-type transformer,
this is achieved by placing half of the low voltage (LV) winding over one leg and the other half
over the second leg or limb. For the high voltage (HV) winding also, half of the winding is over
one leg and the other half over the second leg, as shown in Fig. 2.1.
H.V.
Winding Yoke
L.V
H.V
Limb or
L.V 2 2 2
Leg 2
H.V
L.V
L.V.
Winding
(a) (b)
Figure 2.1 Constructional details of single-phase (a) core-type & (b) Shell-type transformer
36
Low voltage winding is placed adjacent to the steel core and high voltage winding outside, in order
to minimize the amount of insulation required. In shell-type transformer the low voltage and high
voltage windings are wound over the central limb and are interleaved or sandwiched as shown in
Fig.2.1(b). Note that the bottom and top are low voltage coils.
In core-type transformer, the flux has a single path around the legs or yokes. Fig. 2.1(a). In shell-
type transformer, the flux in the central limb divides equally and returns through the outer two legs
as shown in Fig. 2.1(b).
There are two types of windings employed in transformers. The concentric coils are used for core-
type transformer as shown in Fig. 2.1(a) and interleaved (or sandwiched) coils for shell-type
transformers as shown in Fig. 2.2(b).
One type of laminations for the core and shell type of transformers is illustrated in Fig. 2.2 (a) and
(b) respectively. In both core and shell-type transformers, the individual laminations are cut in the
form of long strips of L's, E’s and I's as shown in Fig. 2.3.
Butt Joints
Butt
Joints
Butt
Joints
(a) (b)
Figure 2.2 two adjacent layers for (a) core and (b) shell type transformers
In order to avoid high reluctance at the joints where the laminations are butted against each other,
the alternative layers are stacked as shown in Fig. 2.4.
37
Butt Joints
During the transformer construction first the primary and secondary winding are wound, then the
laminations are pushed through the coil openings, layer by layer and the steel core is placed. The
laminations are then tightened by means of clamps and bolts.
Low-power transformers are air cooled whereas larger power transformers are immersed in oil for
better cooling. In oil-cooled transformer, the coil serves as a coolant and also as an insulation
medium.
V1 P N1 N2 S
38
2.5 EMF Equation of a Transformer
Let the voltage V1 applied voltage primary be sinusoidal (or sine wave). Then the current Im and,
therefore, the flux will flow with the variations of Im . That is, the flux is in time phase with the
current Im and varies sinusoidally. Let sinusoidal variation of flux be expressed as
= m 𝑆𝑖𝑛𝜔𝑡
Where m is maximum of the magnetic flux in Weber and = 2f is the angular frequency in
rad/sec and f is the supply frequency in Hz.
The emf e1 in volt, induced in the primary of N1 turns by the alternating flux is given by
d
𝑒1 = −𝑁1 dt
= −𝑁1 𝜔 𝑚 𝐶𝑜𝑠𝜔𝑡
𝜋
= 𝑁1 𝜔 𝑚 𝑠𝑖𝑛( 𝜔𝑡 − 2 )
𝜋
Its maximum value, E1max occurs when 𝑆𝑖𝑛 (𝜔𝑡 − 2 )is equal to 1.
𝐸1𝑚 = 𝑁1 𝜔 𝑚
𝜋
and 𝑒1 = 𝐸1𝑚 𝑠𝑖𝑛 (𝜔𝑡 − 2 )
= √2𝜋𝑓𝑁1 𝑚
= 4.44𝑓𝑁1 𝑚 (2.1)
Since the primary winding resistance is negligible hence e1, at every instant, must be equal and
opposite of V1. That is,
𝑑
𝑣1 = −𝑒1 = −𝑁1
𝑑𝑡
or 𝑉1 = −𝐸1
39
𝜋
= 𝑁2 𝜔𝑠𝑖𝑛(𝜔𝑡 − )
2
𝜋
= 𝐸𝑚2 𝑠𝑖𝑛( 𝜔𝑡 − )
2
= 4.44𝑓𝑁2 𝑚 (2.2)
𝐼𝑛𝑝𝑢𝑡𝑉𝐴 = 𝑂𝑢𝑡𝑝𝑢𝑡𝑉𝐴
𝑉1 𝐼1 = 𝑉2 𝐼2
and
𝐼1 𝑉 1
= 𝑉2 = 𝑘
𝐼2 1
Hence, the currents are in the inverse ratio of the (voltage) transformation ratio of Eq. (2.3).
𝐸1 𝐸
Also, the ratio of = 𝑁2 = √2𝜋𝑓𝑚 and this shows that the emf per turn in each of the windings
𝑁1 2
is the same.
Example 2.1 A single phase transformer has 350 primary and 1050 secondary turns. The net
cross-sectional area of the core is 55cm2. If the primary winding be connected to a 400 V, 50 Hz
single phase supply, calculate (i) the maximum value of flux density in the core and (ii) the voltage
induced in the secondary winding.
Solution
40
Voltage applied to the primary = 400 V
Induced emf in the primary, E1 voltage applied to the primary, V1 = 400 V
= 55 10-4 m2
𝐸1 = 4.44𝑓𝑚 𝑁1 = 4.44𝑓𝐵𝑚 𝐴𝑖 𝑁1
Example 2.2 The required no-load voltage ratio in a single phase 50 Hz, core type transformer is
6600/500. Find the number of turns in each winding, if the flux is to be 0.06 Wb.
Solution
6000
No-load voltage ratio = 500
Flux = 0.06 Wb
Frequency f = 50Hz
Induced emf in the low voltage winding (secondary) of the transformer is given by,
𝐸2 = 4.44𝑓𝑚 𝑁2
41
or 500 = 4.44 × 50 × 0.06 × 𝑁2
The number of turns in each winding should be a whole number, moreover each winding in the
core type transformer is accommodated on both the limbs. i.e. half number of turns of each winding
on one limb. As such the number of turns in each winding should be even.
Considering all the factors mentioned above, the number of turns in the high voltage winding
N1=500. Here the number of turns finally taken is 500 and not 502, because the high voltage
winding will be split up into a number of coils. With 250 turns on each limb, high voltage winding
on one limb can be split into 5 coils of 50 turns each.
The equivalent circuit for any electrical engineering devices can be drawn if the equations
describing its behavior are known. If any electrical device is to be analyzed and investigated further
for suitable modifications, its appropriate equivalent circuit is necessary. The equivalent circuit for
electromagnetic devices consists of a combination o1 resistances, inductances, capacitances,
voltages etc. Such an equivalent circuit (or circuit model) can, therefore, be analyzed and studied
easily by the direct application of electric circuit theory.
As stated above equivalent circuit is simply a circuit representation of the equations describing the
performance of the device. In the equivalent circuit of Fig. 2.6(a) (rl +jx1) and (r2 + jx2) are the
leakage impedances of the primary and secondary windings respectively. The voltage 𝑉1′ is treated
as a voltage drop in the direction of I1. Recall that the magnitude of 𝑉1′ does not change appreciably
from no load to full load in large transformers. The magnitude of 𝑉1′ depends on f ,N1 and m, since
|𝑉1′ | = |𝐸1 |.
42
The primary current I1 consists of two components. One component 𝐼1′ is the load component and
counteracts the secondary m.m.f. I2N2 completely. The other component is exciting current Ie
which is composed of Ic and Im. The current Ic is in phase with 𝑉1′ and product V1' Ic gives core loss.
The resistance Rc parallel with 𝑉1′ represents the core loss Pc, such that.
2
(𝑉1′ )
𝑃𝑐 = 𝐼𝑐2 𝑅𝑐 = 𝑉1′ 𝐼𝑐 = 𝑅𝑐
𝑉1′
And 𝑅𝑐 = 𝐼𝑐
The current Im lags V1' by 90° and this can, therefore, be represented in the equivalent circuit by a
reactance Xm, such that
𝑉′
𝑋𝑚 = 𝐼 1
𝑚
Rc and Xm are shown in Fig. 2.6 (b), which is the exact equivalent circuit of a transformer. The
resistance Rc and reactance Xm are called core-loss resistance and magnetizing reactance,
respectively.
For minor changes in supply voltage and frequency, which is common under normal operation, Rc
and Xm are treated constant.
In Fig. 2.6 (a) and (b) , the ideal transformer has been introduced to show the transformation of
voltage and current between primary and secondary windings . Even at this stage the transformer
magnetization curve is assumed linear, since the effect of higher order harmonic can't be
represented in the equivalent circuit.
R1 jX1 jX2 R2
+
+ + +
I1 I2
V1 E1 E2
V2
_ _ _ _
43
jx2 r2
r1 jx1 I1' N1 N2
+
+ + +
I1 Ie I2
Ic Im
V1 E2
E1 Rc jXm
V2
_ _ _
_
Ideal
transformer
_ _ _
_
Ideal
transformer
+ + N
I e 1 I2
N N2
I1 1
N2 I'c I'm
N
V1 2 N V2
V1 N1 E1 2 Rc jXm E2
N1
_ _
_
Ideal
transformer
In transformer analysis, it is usual to transfer the secondary quantities to primary side or primary
quantities to secondary side. Secondary resistance drop I2 r2 when transferred to primary side must
be multiplied by the turns ratio N1 /N2.
𝑁
Secondary resistance drop, when transferred to primary =(𝐼2 𝑟2 ) 𝑁1
2
𝑁 𝑁 𝑁
= (𝐼1 ⋅ 𝑁1 ⋅ 𝑟2 ) 𝑁1 , (𝑝𝑢𝑡𝑡𝑖𝑛𝑔𝐼2 = 𝐼1 ⋅ 𝑁1)
2 2 2
𝑁1 2
= 𝐼1 [( ) 𝑟2 ] = 𝐼1 𝑟2′
𝑁2
𝑁 2
Where 𝑟2′ = 𝑟2 ⋅ (⋅ 𝑁1)
2
44
If resistance r2' is placed in the primary circuit, then the relation between voltage V1 and V2 is
unaffected. This resistance r2' is called the secondary resistance referred to primary. Therefore,
the total resistance in the primary circuit is
𝑁1 2
𝑟𝑒1 = 𝑟1 + 𝑟2 ⋅ ( ) = 𝑟1 + 𝑟2′
𝑁2
Hence re1 is called the transformer equivalent (or total) resistance referred to primary winding.
𝑁 2
Similarly, the primary resistance referred to secondary is 𝑟1 ⋅ (⋅ 𝑁2) and the equivalent (or total)
1
𝑁2 2
𝑟𝑒2 = 𝑟2 + 𝑟1 ⋅ (⋅ ) = 𝑟2 + 𝑟1′
𝑁1
r1 x1 I1 I2 x2 r2
+
+ Ie
I2
Ic Im
V1 E1 E2 V2
Rc jXm
_
_
𝑁1 𝑁1 2
𝐼2 𝑥2 ( ) = 𝐼1 [( ) 𝑥2 ] = 𝐼1 𝑥2′
𝑁2 𝑁2
The quantity x '2 is called the secondary leakage reactance referred to primary. Total primary
leakage reactance is
𝑁1 2
𝑥𝑒1 = 𝑥1 + 𝑥2 ⋅ ( ) = 𝑥1 + 𝑥2′
𝑁2
Where xe1, is called the equivalent or total leakage reactance referred to primary. Likewise, the
equivalent or total leakage reactance referred to secondary is
𝑁2 2
𝑥𝑒2 = 𝑥2 + 𝑥1 ⋅ ( ) = 𝑥2 + 𝑥1′
𝑁1
45
𝑧𝑒1 = 𝑟𝑒1 + 𝑗𝑥𝑒1
𝑁 2 𝑁 2
𝑧𝑒1 = (𝑁1) 𝑧𝑒2 and 𝑧𝑒2 = (𝑁2) 𝑧𝑒1
2 1
In general, when values are referred to either circuit the following conditions should be kept in
mind
i. The energy condition (i.e. the active and reactive) power should be remain unchanged
ii. The phase angle between voltage and current i.e. power factor, should be remain the same
and
iii. The referring factor must be the same for all values of the same type.
Simplification of the exact equivalent circuit: The equivalent circuit of Fig. 2.6 (b) can be
simplified by referring all the quantities to primary or secondary and at the same time, moving the
ideal transformer to one side. If the secondary quantities are referred to primary, the equivalent
circuit of Figure2.6 (c) is obtained. Since it is usual to omit the ideal transformer, it is shown dotted
for the sake of completeness. When the primary quantities are referred to the secondary side, the
𝑁 2 𝑁 2
equivalent circuit of Fig. 2.6(d) is obtained. Note that 𝑅𝑐′ = 𝑅𝑐 (𝑁2) and 𝑋𝑚
′
= 𝑋𝑚 (𝑁2 ) . The
1 1
exact equivalent circuits of Fig. 2.6(c) and (d) are known as T-circuits for a transformer, referred
to primary and secondary windings respectively.
In the equivalent circuits of Fig. 2.6 (c) and (d), the referred quantities with suitable notation, have
been used. A more general equivalent circuit can be drawn as shown in Fig. 2.6(e), where for
simplicity (i) a particular notation for referred-quantities has been dropped (ii) the complex
notation (bar over I, j with reactance etc.) has been given up and (iii) the ideal transformer is not
shown. If the general equivalent circuit refers to the primary, one has to keep in mind that the
secondary quantities have been referred to the primary side. On the other hand, if the general
equivalent circuit refers to the secondary, then the primary quantities must be referred to the
secondary side. Thus in the general equivalent circuit of a transformer, one has merely to keep in
mind about the side to which all the quantities have been referred.
46
It may be interesting at this stage to draw the phasor diagram for the equivalent circuit of Fig.
2.6(e) from a knowledge of the electric circuit theory. Assume that the secondary load voltage V2
load current I2 and angle 2, by which I2 lags V2 are known. First of all, draw I2 lagging V2 by an
angle 2 and then add I2 (r2 + jx2) to V2 to obtain E2, Fig. 2. It is obvious from Fig.2.6 (e) that
current Im due to voltage E2, must lag it by 90° and further Ic must be in phase with E2. The phasor
sum of Ic and Im gives Ie and phasor sum of I2 and Ie gives I1. The voltage drop I1 (r1 + jx1) is
now added to E2 to obtain V1 as shown in Fig. 2. The secondary p.f. is cos2 lagging and the
primary p.f. is cos 1 lagging.
The voltage drops I1 (r1 + jx1) and I2 (r2 + jx2) have been drawn to a much larger scale, in
comparison with V1 or V2 for the sake of clarity.
jI1 x
V1 1
I1r1
jI2 x
2
E2
I2r2
V2
I1
I2
1 2
Ic Ie
Im
Figure 2.7: Phasor diagram for equivalent circuit of Figure 2.6 (e)
Approximate Equivalent Circuit: Approximate equivalent circuit is obtained from the exact
equivalent circuit Fig. 2.6(e), if the shunt branch (Rc and Xm in parallel) is moved to the primary
or secondary terminals as shown in Fig. 2.8(a) and (b) respectively. It may be seen from Fig. 2.8
(a) that the exciting current Ie does not flow through rl and x1, whereas Ie does flow through r1 and
x1 in the exact equivalent circuit. Thus the primary leakage impedance drop due to the exciting
current, i.e. Ie( r1 +jx1) has been neglected in Fig. 2.8 (a), though it is not so actually.
It may also be seen from Fig. 2.8 (b) that Ie flows through r2 and x2, whereas Ie does not flow
through r2 and x2 in the exact equivalent circuit. Thus the secondary leakage impedance drop due
to Ie, i.e. Ie (r2 + jx2) has been included, though Ie (r2 + jx2) is actually zero.
47
x1+x2=xe r1+r2=re
I1 I2
+ +
Ie
Ic Im
V1 V2
Rc jXm
_ _
(a)
r1+r2=re x1+x2=xe
+ +
Ie I2
I1
Ic Im
V1 V2
Rc jXm
_ _
(b)
x1+x2=xe
r1+r2=re x1+x2=xe
+
+ +
+
I1=I2 I1=I2
V1 V1 V2
V2
_ _
_ _
(c) (d)
Fig. 2.8 (a) and (b) Approximate equivalent circuits of a transformer (c) and (d) Simplified forms
of the approximate equivalent circuit.
Since the exciting current is only about 2 to 6 per cent of the rated winding current in power and
distribution transformers, the error introduced by neglecting Ie (r1 +jx1) or including Ie (r2 + jx2) is
insignificant. However, the computational labor involved is reduced considerably by the use of
approximate equivalent circuits of Fig. 2.8(a) and (b). As before, one must keep in mind about the
side to which all the equivalent-circuit quantities have been referred.
Still further simplification is achieved by neglecting the shunt branch Rc and Xm in Fig. 2.8 (a) and
(b) and this leads to equivalent circuit of Fig. 2.8(c). This simplification is tantamount to neglecting
exciting current Ie in comparison with rated currents, which is almost justifiable in large
transformers, say over 100 KVA or so. For transformers having ratings near 500 KVA or more,
the equivalent resistance re is quite small as compared with equivalent leakage reactance xe.
Consequently, re may be neglected, leading to the equivalent circuit of Figure3 (d). Thus, when a
48
large power system is studied, a transformer is usually replaced by its equivalent circuit of the
form shown in Fig. 2.8(d).
The equivalent circuit Fig. 2.6(e) should be used only when the exciting current is a large
percentage of the rated current e.g., in audio-frequency transformers used in electronic circuits, in
transformers used for relaying and measurement purposes etc. For high voltage surge
investigations, the transformer equivalent circuit must be modified to include the effects of inter-
turn and turn to earth capacitances.
The equivalent circuit parameters can also be obtained from the physical dimensions of the
transformer core and its winding details. Complete analysis of the transformer can be carried out,
once its equivalent circuit parameters are known. The power required during these two tests is
equal to the appropriate power loss occurring in the transformer.
The circuit diagram for performing open circuit test on a single-phase transformer is given in Fig.
2.9 (a). In this diagram, a voltmeter, wattmeter and an ammeter are shown connected on the low
voltage side of the transformer. The high voltage side is left open circuited. The rated frequency
voltage applied to the primary, i.e. low voltage side, is varied with the help of a variable ratio auto-
transformer. When the voltmeter reading is equal to the rated voltage of the L.V. winding , all
three instrument readings are recorded.
Auto-
transformer Ie
A + Ie
W Ic Im
V1=E1 R Xm
c
V H.V L.V _
(a) (b)
49
Figure 2.9 (a) Circuit diagram for open-circuit test on a transformer and (b) approximate equivalent
circuit at no load
The ammeter records the no-load current or exciting current Ie. Since Ie is quite small (2 to 6%) of
rated current), the primary leakage impedance drop is almost negligible, and for all practical
purposes, the applied voltage V1 is equal to the induced emf E1. Consequently, the equivalent
circuit of Figure 2.6 (e) gets modified to that shown in Figure 2.9( b).
The input power given by the wattmeter reading consists of core loss and ohmic loss. The exciting
current being about 2 to 6 percent of the full load current, the ohmic loss in the primary ( = I e2 r1 )
2 2
varies from 0.04 percent 100 to 0.36 percent of the full-load primary ohmic loss.
100 100
In view of this fact, the ohmic loss during open circuit test is negligible in comparison with the
normal core loss (approximately proportional to the square of the applied voltage). Hence the
wattmeter reading can be taken as equal to transformer core loss.
Pc = core loss
Then 𝑃𝑐 = 𝑉1 𝐼𝑒 𝑐𝑜𝑠 𝜃𝑜
𝑃
No load p.f. = 𝑐𝑜𝑠 𝜃𝑜 = 𝑉 𝑐𝐼
1 𝑒
𝑉1 𝑉1
Core loss resistant 𝑅𝐶𝐿 = =𝐼
𝐼𝑐 𝑒 𝑐𝑜𝑠 𝜃𝑜
𝑉12 𝑉12 1
=𝑉𝐼 =
1 𝑒 𝑐𝑜𝑠 𝜃𝑜 𝑃𝑐 𝑜
50
Magnetizing reactance
𝑉 𝑃𝑐
𝑋𝑚𝐿 = 𝐼 1 = 𝐼
𝑚 𝑒 𝑠𝑖𝑛 𝜃𝑜
The subscript L with Rc and Xm is used merely to emphasize that these values are for the L.V.
side.
It must be kept is mind that the values of Rc and Xm, in general, refer to the side, in which the
instruments are placed (the L.V. side in the present case). A voltmeter is sometimes, used at the
open-circuited secondary terminals, in order to determine the turns ratio.
ii the shunt branch parameters of the equivalent circuit, i.e. Rc and Xm and
The low voltage-side of the transformer is short-circuited and the instruments are placed on the
high voltage side, as illustrated in Fig. 2.10 (a).
Auto-
transformer
A
W
Short circuit
H.V L.V
Figure 2.10 (a) connection diagram for short circuit test on a transformer
r1 x1 x2 r2
Short circuit
Rc Xm
Figure 2.10 (b) Equivalent circuit with short-circuit on the secondary side
r1 x1 x2 r2
Isc
Short circuit
Vsc
51
Figure 2.10 (c) Transformer equivalent circuit with secondary short-circuited
The applied voltage is adjusted by auto-transformer, to circulate rated current in the high voltage
side. In a transformer, the primary m.m.f. is almost equal to the secondary m.m.f., therefore, a
rated current in the H.V. winding causes rated current to flow in the L.V. winding.
A primary voltage of 2 to 12% of its rated value is sufficient to circulate rated currents in both
primary and secondary windings. From Fig. 2.10 (b), it is clear that the secondary leakage
impedance drop appears across the exciting branch (RC and Xm in parallel). About half (1 to 6%)
of the applied voltage appears across the secondary leakage impedance and, therefore, across the
exciting branch. The core flux induces the voltage across the exciting branch and since the latter
is 1 to 6% of rated voltage, the core flux is also 1 to 6% of its rated value. Hence the core loss,
being approximately proportional to the square of the core flux, is 0.01 percent
1 1 6 6
= 100 to 0.36 percent = 100 of its value at rated voltage. The
100 100 100 100
wattmeter, in short circuit test, records the core loss and the ohmic loss in both windings. Since
the core loss has been proved to be almost negligible in comparison with the rated voltage core
loss, the wattmeter can be taken to register only the ohmic losses in both windings.
At rated voltage, the exciting Current is 2 to 6% of full load current. When the voltage across the
exciting branch is 1 to 6% of rated voltage, the exciting current may be 0.02 percent
2 1 6 6
= 100 to 0.36% percent = 100 of its full-load current and can,
100 100 100 100
therefore, be safely ignored. As a result of this the equivalent circuit of Fig.2.6(e), with the
secondary short-circuited, gets modified to that shown in Fig. 2.10 (c)
Let VSC, ISC and PSC be the voltmeter, ammeter and wattmeter readings; then from Fig. 2.10 (c),
equivalent leakage impedance referred to H.V. side,
𝑉𝑆𝐶
𝑍𝑒𝐻 = 𝐼𝑆𝐶
𝑃𝑆𝐶
equivalent resistance referred to H.V. side, 𝑟𝑒𝐻 = 2 and equivalent leakage reactance referred to
𝐼𝑆𝐶
2 2
H.V. side, 𝑋𝑒𝐻 = √𝑍𝑒𝐻 − 𝑟𝑒𝐻
52
In reH, XeH and ZeH„, the subscript H is used to indicate that these quantities are referred to H.V.
side. These parameters can however, be referred to the L.V. side, if required.
In the analysis of transformer equivalent circuit, the values of equivalent resistance and equivalent
leakage reactance referred to either side are used. However, if the leakage impedance parameters
for both primary and secondary are required separately, then it is usual to take r1 = r2=½ re) and x1
= x2=½ xe, referred to the same side.
Voltage regulation of a transformer can be determined from the data obtained from short-circuit
test. Data of both open-circuit and short-circuit tests is necessary
How can a wattmeter connected on the H.V. side, record the ohmic in the L.V. winding also?
When rated current is made to flow in the H.V. winding, the L.V. winding must also carry rated
current, because the transformer action requires I1N1= I2N2. The flow of rated current in the L.V.
winding causes ohmic loss, which must be supplied from somewhere. The only way to provide
L.V. winding loss is from the input to H.V. side. But the entire input power to H.V. side is recorded
by the wattmeter, therefore, the ohmic losses in both windings are given by the wattmeter reading.
It has already been stated that open-circuit and short-circuit tests should be performed on the L.V.
side and H.V. side respectively only for the sake of convenience. This can he illustrated by
considering a 3300/220V, 33KVA, single-phase transformer.
For open-circuit test on low voltage side, the ranges of voltmeter, ammeter and wattmeter are 220V
(rated value), 6A ( 2 to 6% of rated current of 150A) and 6A, 220V respectively. These are the
standard ranges for ordinary instruments and therefore, more accurate readings can be obtained. If
the open circuit test is performed on the H.V. side, a source of 3300V may not be readily available.
At the same time, the instrument ranges are 3300V, 0.4A and 0.4A , 3300V which are which are
53
not within the range of ordinary instruments and the results obtained may not be so accurate . Also
it may not be safe to work on the high voltage side.
For a short-circuit test on the H.V. side, the instrument ranges are 165V (2 to 12% of rated voltage
of 3300V), l0A (rated current) and 10A, 165V, which are well within the range of the ordinary
instruments. On the other hand, instrument ranges, for a short-circuit test on L.V. side are 11V,
150A, and 150A. 11V. Instruments of such ranges and auto-transformer capable of handling 150A
may not be readily available and at the same time, the results may not be so accurate. It is for these
reasons that the open-circuit and short-circuit tests are conducted on L.V. and H.V. sides
respectively.
On the primary side of a two-winding transformer, one terminal is positive with respect to the other
terminal at any one instant. At the same instant, one terminal of the secondary winding is positive
with respect to the other terminal. These relative polarities of the primary and secondary terminals
at any instant must be known if the transformers are to be operated in parallel or are to be used in
a polyphase circuit.
E1- E2 E1+ E2
V V
A2 a2 A2 a1
- - - +
E1 E2 E1
E2
+ + + -
A1 a1 A1 a2
(a) (b)
Figure 2.11 Polarity test on a two-winding transformer (a) subtractive polarity and (b) additive
polarity
When viewed from the H.V. side, the terminals are marked A1 and A2, the former, i.e. A1 being on
the extreme right. Terminals A1 and A2 marked plus and minus arbitrarily in Figure 6. Now
terminal A1 is connected to one end of the secondary winding and a voltmeter is connected between
A2 and the other end of the secondary winding. A voltage of suitable value is now applied to the
H.V. winding. Let E1 and E2 be the e.m.fs induced on H.V. and L.V. sides respectively. If the
voltmeter reading is equal to E1–E2 then secondary terminal connected to A1 is positive and is
marked a1, the L.V. terminal connected to A2 through the voltmeter is negative and is marked a2
54
as shown in Figure 2.11(a). If voltmeter reading is equal to E1+E2, then the terminals connected to
A1 and A2 are negative and positive and are marked a2 and a1 respectively as shown in Figure
2.11(b). The subscript numbers 1,2 on the H.V. and L.V. windings are so arranged that when A 2
is negative with respect to A1. a2 is also negative with respect to a1 at the same instant. In other
words, if the instantaneous emf is directed from A2 to A1 in H.V. winding, it is at the same time
directed from a2 to a1 in the L.V. winding.
When the voltmeter reads the difference E1–E2, the transformer is said to possess a subtractive
polarity and when voltmeter reads E1+E2 the transformer has additive polarity. In subtractive
polarity, the voltage between A2 and a2 (or A1 and a1) is reduced. The leads connected to these
terminals and the two windings are, therefore, not subjected to high voltage stress. In additive
polarity the windings and the leads connected to A1, A2, a1 and a2 are subjected to high voltage
stresses. On account of these reasons, subtractive polarity is preferable to additive polarity.
Example 2.3 A 20 kVA, 2500/250 V, 50 Hz, single-phase transformer gave the following test
result
Compute the parameters of the approximate equivalent circuit referred to high-voltage and low-
voltage sides. Also draw the exact equivalent circuit referred to the low -voltage side.
Solution
V 250
hence, R cL = 1 = = 595
I c 0.42
V 250
X mL = 1 = = 187
I m 1.336
55
Alternatively, the value of RcL and XmL can be determined as follows:
𝑉12 (250)2
𝑅𝑐𝐿 = = = 595𝛺
𝑃𝑐 105
𝑉 250
Now 𝐼𝑐 = 𝑅 1 = 595 = 0.42𝐴
𝑐𝐿
𝑃𝑠𝑐 320
𝑟𝑒𝐻 = 2 = = 8𝛺
𝐼𝑠𝑐 82
2
𝑥𝑒𝐻 = √𝑍𝑒𝐻 2
− 𝑟𝑒𝐻 = √132 − 52 = 12𝛺
1 2 1 2 1 2 1 2
𝑟𝑒𝐿 = 𝑟𝑒𝐻 × (𝑘) = 5 (10) = 0.05𝛺; 𝑥𝑒𝐿 = 𝑥𝑒𝐻 × (𝑘) = 12 (10) = 0.12𝛺
Ie Ie
(a) (b)
Figure (a) approximate equivalent circuit referred to L.V. side and exact equivalent circuit
referred to L.V. side.
56
𝑋𝑚𝐻 = 𝑋𝑚𝐿 × (𝑘)2 = 187(10)2 = 18,700𝛺
′ 1 1
𝑟1𝐿 = 𝑟1𝐻 = 2 𝑟𝑒𝐿 = 2 0.05 = 0.025𝛺
′ 1 1
𝑥1𝐿 = 𝑥1𝐻 = 2 𝑥𝑒𝐿 = 2 0.12 = 0.06𝛺
The purpose of first considering an ideal transformer, i.e. a transformer with no core losses, no
winding resistance, no magnetic leakage and constant permeability, is merely to highlight the most
important aspect of transformer action. Such transformer never exists and now the phasor diagrams
of real transformer will be considered.
The magnetic flux m being common to both the primary and secondary is drawn first. The induced
emf E1 and E2 lag m by 900 and are shown accordingly in Figure 2.12. The voltage -E1 is being
replaced by V1 just for convenience.
The core loss (or iron loss) consists of hysteresis loss and eddy current loss. These losses are always
present in the ferromagnetic core of the transformer, since the transformer is an ac operated
magnetic device. The hysteresis loss in the core is minimized by using high grade material such as
cold-rolled-grain-oriented (CRGO) steel and the eddy current loss is minimized by using thin
lamination for the core.
The current in the primary is alternating, therefore, the magnetizing force H is cyclically varying
from one positive value say Hl to a corresponding negative value −Hl, Figure 2.12 (a). When the
magnetizing force is - Hl, the flux density is maximum negative equal to OM. As the magnetizing
57
force decreases from - Hl, the current Ie decreases and becomes zero for a flux density, or flux,
equal to ON. When the current Ie becomes positive and equal to OP, the flux is reduced to zero but
it is going to become positive. The traverse of the loop along the arrows involves time. When Ie is
crossing zero positive (passing through zero and becoming positive), the core flux is negative and
is equal to ON in Figure 2.12(a). This is shown in Figure 2.12(b) at instant tl, where waveforms
are assumed sine waves. When Ie is positive and equal to OP, Figure 2.12(a), the flux is crossing
zero and becoming positive; this is shown in Figure 2.12(b) at instant t2. It is seen from Figure
2.12(b) that exciting current Ie leads the magnetic flux (or lags Ie) by some time angle . This
angle of lead, or lag, being dependent on the hysteresis loop, is called the hysteretic angle. In
Figure 2.12 (c), Ie is shown leading , or is shown lagging Ie, by hysteretic angle .
or ie ,
ie
P
-H1 0
0
O t
P H1 or Ie
N N
M
t 2
t1
(a)
(b)
V1
jI e
x1
Ie r 1
V’1=-E1
0
Ie
Ic
l1
Im
Ex1=-jIex1
E1, E2=V2
(c)
Figure 2.12 (a) Hysteresis loop for transformer core (b) exciting current and core flux waveforms
and (c) no-load phasor diagram of a transformer.
58
The no-load primary current Ie is called the exciting current of the transformer and can be resolved
into two components. The component Im along m is called the reactive or magnetizing current,
since its function is to provide the required magnetic flux m. The second component along V'1 is
Ic and this component is called the core- loss component. When multiplied by V'1 gives the total
core loss Pc.
P
V ' 1 I c = Pc or I c = c Amp .
V' 1
Ie = I m + I 2c
2
Note that in an ideal transformer, core-loss current I c = 0 and therefore exciting current Ie equals
The effect of primary resistance r1 can be accounted for, by adding to V1, a voltage drops equal to
Ie r1 as shown in Figure 2.12 (c). Note that Ier1 is in phase with Ie and is drawn parallel to Ie in the
phasor diagram.
The existence of electrical potential difference is essential for the establishment of current in an
electric circuit. Similarly, the magnetic potential difference is necessary to establish flux in a
magnetic circuit. This magnetic potential difference establishes:
ii the primary leakage flux l1 which links only the primary winding.
The distinctive behavior of the mutual flux m and the primary leakage flux l1The mutual flux m
exists entirely in the ferromagnetic core and, therefore, involves hysteresis loop. The current Ie that
establishes m must lead it by some hysteresis angle. On the other hand, the primary leakage flux
l1 exists largely in air. Although l1 does pass through some iron, the reluctance offered to l1 is
mainly due to air. Consequently l1 does not involve any hysteresis loop and it can be taken to be
59
in phase with the current Ie that produces it, Figure2.12(c). In the primary winding, induces an
emf E1 lagging it by 90°; similarly, the primary leakage flux l1 induces an emf Ex1 in the primary
winding and lagging it (i.e. l1) by 90°. Since Ie leads Ex1 by 90°, it is possible to write EX1= -jIexl.
The primary applied voltage Vl must have a component jIexl, equal and opposite to Exl. Here xl has
the nature of reactance and is referred to as the primary leakage reactance in ohms. It may be noted
that x1 is a fictitious quantity merely introduced to represent the effects of primary leakage flux.
The total voltage drop in primary at no-load is Ie (r1+jx1) = Iez1 where z1 is the primary leakage
impedance. Therefore Fig. 2.12(c) gives the phasor diagram of transformer at no-load, where Nl is
assumed to be equal to N2. The primary voltage equation at no-load can be written as:
The primary leakage impedance drop shown in Figure 2.12(c), is drawn to a larger scale, in
comparison with Vl' or Vl, just for the sake of clarity. At no-load and V'1 and V1 are very nearly
equal. Even at full load primary leakage impedance drop in power transformer is about 2 to 5% of
V1, so that the magnitude of V'1 or E1 does not change appreciably from no-load to full load.
60
When the switch S is closed, secondary current I2 starts flowing from terminal n to the load.
Assume the load to have a lagging power factor so that I2 lags secondary load voltage V2 by an
angle 2. At first V2 is drawn with I2 lagging V2 by the secondary p.f. angle 2, Figure 1.11 (a).
The secondary resistance drop is accounted for, by drawing I2r2 parallel to I2. The secondary m.m.f.
I2N2 gives rise to a leakage flux l2 which links only the secondary and not the primary. The flux
l2 is called the secondary leakage flux and is in phase with I2, for the same reason that l1 is in
phase with Ie in Figure 2.12(c). The secondary leakage flux induces emf Ex2 in the secondary
winding, lagging l2 by 90°. The secondary no load voltage E2 must have a component equal and
opposite to –jx2I2. Thus, the phasor sum of V2, I2r2 and jx2I2 gives the secondary induced emf E2
as shown in Figure 2.15(a).
where z2 is the secondary leakage impedance of the transformer. Further the mutual flux is drawn
leading E2 by 90° and exciting current Ie is drawn leading by the hysteretic angle . Note that
the phasor V2 has purposely been taken to the left of vertical line, so that E2 is vertically downward
and the mutual flux is horizontal.
The component of the primary current which neutralizes the demagnetizing effect of I2 is
I'1 (I1N1 = I2N2) and drawn opposite to I2. The phasor sun of I'1 and Ie gives the total primary
current I1 taken from the supply mains . The primary leakage impedance drop I1(r1+jx1) is depicted
as explained earlier.
The voltage equation for primary circuit under load can be written as
where z1 is the primary leakage impedance of transformer. Note that the angle 1 between V1 and
I1 is the primary power factor angle under load.
If the secondary load current I2 leads the voltage V2 such that the load p.f. is leading, then the
phasor diagram for the transformer is as shown in Figure 2.15 (b). The entire procedure for drawing
the phasor diagram is the same as explained for Figure 2.15 (a).
61
V1
jI1 x
jI 1
x1
1
V1
I1r1
I 1 r1
V’1=-E1
V’1=-E1
I1
I’1
1
I1
I’1
Ex2 1
=-jIex1 Ie Ie
l 2
2
I2 2
I2r2 V2 I2
V2
jI 2
E1, E2 E1, E2
x2
I 2 r2
jI 2x 2
Figure 2.15 Transformer phasor diagram for (a) lagging p.f. load and (b) leading p.f. load
Constant voltage is the characteristics of most domestic, commercial and industrial loads. It is
therefore, necessary that the output voltage of a transformer must remain within narrow limits as
the load and its power factor vary. This requirement is more stringent in distribution transformers
as these directly feed the load centers. The voltage drop in a transformer on load is chiefly
determined by its leakage reactance which must be kept as low as design and manufacturing
techniques would permit.
𝐸2 − 𝑉2
𝑉𝑜𝑙𝑡𝑎𝑔𝑒𝑟𝑒𝑔𝑢𝑙𝑎𝑡𝑖𝑜𝑛 = 𝑖𝑛𝑝. 𝑢
𝑠𝑒𝑐𝑜𝑛𝑑𝑎𝑟𝑦 𝑟𝑎𝑡𝑒𝑑 𝑣𝑜𝑙𝑡𝑎𝑔𝑒
It is stipulated that the secondary rated voltage of a transformer is equal to the secondary terminal
voltage at no load, i.e. E2.
𝐸2 −𝑉2 𝐸2 −𝑉2
𝑉𝑜𝑙𝑡𝑎𝑔𝑒𝑟𝑒𝑔𝑢𝑙𝑎𝑡𝑖𝑜𝑛 = 𝑖𝑛𝑝. 𝑢 = × 100 𝑖𝑛 𝑝𝑒𝑟𝑐𝑒𝑛𝑡𝑎𝑔𝑒
𝐸2 𝐸2
At no-load, the primary leakage impedance drop is almost negligible, therefore, the secondary no-
𝑁
load voltage 𝐸2 = 𝑉1 𝑁2 . The expression for voltage regulation can also be written as
1
62
𝑁 𝑁
𝑉1 𝑁2 − 𝑉2 𝑉1 − 𝑉2 𝑁1
1 2
× 100𝑖𝑛𝑝𝑒𝑟𝑐𝑒𝑛𝑡𝑎𝑔𝑒 = × 100𝑖𝑛𝑝𝑒𝑟𝑐𝑒𝑛𝑡𝑎𝑔𝑒
𝑁 𝑉1
𝑉1 𝑁2
1
The change in secondary terminal voltage with load current is due to the primary and secondary
leakage impedances of the transformer. The magnitude of this change depends on the load power
factor, load current, total resistance and leakage reactance of a transformer.
A distribution transformer should have a small value of voltage regulation (i.e. good voltage
regulation) so that the terminal voltage at the consumers does not vary widely as the load changes.
For a transformer of large voltage regulation (i.e. poor voltage regulation), the voltage at the
consumers' terminals will fall appreciably with increase in load. This has a detrimental effect on
the operation of fluorescent tubes, T.V. sets, refrigeration motors, etc since these are designed to
operate satisfactorily at a constant voltage. Thus, distribution transformer should be designed to
have a low value of leakage impedances.
The voltage regulation of a transformer can be obtained from its approximate equivalent circuit
referred to primary or secondary. Figure 2.16 (a) illustrates the approximate equivalent circuit of
a transformer referred to the secondary side and the phasor diagram for this circuit is drawn in
Figure 2.16 (b) for a lagging power factor load. For the calculation of voltage regulation, draw an
arc of radius OD meeting the extension of line OA in F. It may be seen from Figure 2.16 (b) that
OF (= E2) is approximately equal to OC.
𝐸2 = 𝑂𝐶 = 𝑂𝐴 + 𝐴𝐵 + 𝐵𝐶(𝑜𝑟𝐵′𝐶′)
= 𝑂𝐴 + 𝐴𝐵′ 𝑐𝑜𝑠 𝜃2 + 𝐷𝐵′ 𝑠𝑖𝑛 𝜃2
= 𝑉2 + 𝐼2 𝑟𝑒2 𝑐𝑜𝑠 𝜃2 + 𝐼2 𝑥𝑒2 𝑠𝑖𝑛 𝜃2
63
F
D C’
2 C
E2 I2xe2
re2 xe2 2
+ + B B’
2
I2re2
E2 V2 A
V2
_ _
I2
2
Figure 2.16 (a) approximate equivalent circuit of a 2-winding transformer, referred to secondary;
(b) the phasor diagram of the circuit of Figure 2.16(a) for lagging power factor load.
Note carefully that E2-V2 is not equal to AD i.e. I2ze2. The change in secondary terminal voltage is
equal to the magnitude of E2 minus the magnitude of V2.
𝐼2 𝑟𝑒2 2 𝑟
𝐼2𝑟 𝑂ℎ𝑚𝑖𝑐𝑙𝑜𝑠𝑠𝑎𝑡𝑟𝑎𝑡𝑒𝑑𝑐𝑢𝑟𝑟𝑒𝑛𝑡
𝑒2
Also 𝜀𝑟 = = =
𝐸2 𝐸2 𝐼2𝑟 𝑅𝑎𝑡𝑒𝑑𝑉𝐴
𝐼2 ⋅𝑥𝑒2
Similarly, for rated current I2, = 𝜀𝑥
𝐸2
From Eq. (2.2), the per unit voltage regulation at rated current is given by
64
(𝜀𝑟 𝑐𝑜𝑠 𝜃2 + 𝜀𝑥 𝑠𝑖𝑛 𝜃2 ) × 100 (2.3b)
It should be noted that Eqs. (2.1) to (2.3) are valid for lagging power factors only. For leading
power factor loads, the phasor diagram of Figure reveals that
Therefore, secondary terminal voltage drop, for any load current I2, is
In case I2 is the rated (or full-load)current, then p.u. voltage regulation is given by
𝜀𝑟 𝑐𝑜𝑠 𝜃2 − 𝜀𝑥 𝑠𝑖𝑛 𝜃2
Condition for zero voltage regulation: It can be seen from Eq. (2.3) that voltage regulation varies
with load power factor. If load power factor is varied with constant values of load current and
secondary emf, then zero voltage regulation will occur when
𝜀𝑟 𝑐𝑜𝑠 𝜃2 + 𝜀𝑥 𝑠𝑖𝑛 𝜃2 = 0
𝜀 𝐼2 𝑟𝑒2 𝑟
𝑡𝑎𝑛 𝜃2 = − 𝜀𝑟 = − 𝐼 𝑥 = − 𝑥𝑒2
𝑥 𝐸2 2 𝑒2 𝑒2
𝐸2
𝑥𝑒2
magnitude of the load p.f. , 𝑐𝑜𝑠 𝜃2 = 𝑟𝑒2
The negative value of tan2 indicates a leading power factor. Therefore, zero voltage regulation
𝑥𝑒2 𝑥𝑒2
occurs when load power factor is leading. For leading p.f.s. greater than , the voltage
𝑧𝑒2 𝑧𝑒2
regulation will be negative, i.e. the voltage will rise from its no load value, as the transformer load
is increased.
= 𝜀𝑟 𝑐𝑜𝑠 𝜃2 + 𝜀𝑥 𝑠𝑖𝑛 𝜃2 .
65
The condition for maximum voltage regulation is obtained by dedifferentiating the above
expression with respect to 2 and equating the results to zero. Here again the load current and
secondary emf are assumed to remain constant.
𝑑
(𝑝. 𝑢. 𝑟𝑒𝑔𝑢𝑙𝑎𝑡𝑖𝑜𝑛) = −𝜀𝑟 𝑠𝑖𝑛 𝜃2 + 𝜀𝑥 𝑐𝑜𝑠 𝜃2 = 0
𝑑𝜃2
𝜀 𝑥𝑒2
Or 𝑡𝑎𝑛 𝜃2 = 𝜀𝑥 =
𝑟 𝑟𝑒2
𝑟
Any 𝑐𝑜𝑠 𝜃2 = 𝑧𝑒2
𝑒2
Here tan2 is positive; therefore, maximum voltage regulation occurs at lagging load p.f. equal to
𝑟𝑒2
. In other words, maximum voltage regulation occurs when load power-factor angle 2 is equal
𝑧𝑒2
𝐼2 2 2 ) 𝐼2 𝑧𝑒2
=𝐸 (𝑟𝑒2 + 𝑥𝑒2 = = 𝑧𝑒2 𝑝𝑢
2 𝑧𝑒2 𝐸2
E2
2 2
I2 r 2 =
I2 r I 2 re2 cos 2 e2 I2xe2
e2
x
I2
e2
I 2 x e2 cos 2
e2
V2
I2 x
V2 V2
E2
C E2 = V2
E2 I2re2
E2 < V2
E2 > V2
I2
I2
2 2
2 90
I2
O O O
Figure 2.17 Phasor diagram for 1-phase transformer for (a) negative voltage regulation (V.R);(b)
zero V.R and (c) maximum V.R
Thus, the magnitude of maximum voltage regulation is equal to the p.u value equivalent leakage
impedance of the transformer. For example, if a transformer has ze2 = 0.054. then magnitude of
maximum possible voltage regulation is 5.4%.
Phasor diagrams for a single-phase transformer for different operating power factors are
illustrated in Figure 2. In Figure 2.17 (a), E2 < V2 voltage regulation (V.R.) is therefore negative.
66
In Figure 2.17 (b) E2 = V2. V.R is zero. Figure 2.17 (c) is drawn under the condition of maximum
V.R, because here load power-factor angle 2 = leakage-impedance angle of the transformer
𝑥 𝑟
where 𝜙 = 𝑡𝑎𝑛−1 ( 𝑒2⁄𝑟𝑒2 ) = 𝑐𝑜𝑠 −1 ( 𝑒2⁄𝑧𝑒2 )
Example 2.4 A 6600/440 V, single-phase transformer has an equivalent resistance of 0.02 p.u.
and an equivalent reactance of 0.05 p.u. Find the full-load voltage regulation at 0.8 pf lagging, if
the primary voltage is 6600 V. Find also the secondary terminal voltage at full load.
Solution
V2 = 440 + 20.25 V
Example 2.5 A short-circuit test, when performed on the H.V. side of a 10 kVA, 2000/400 V,
single-phase transformer gave the following data:
60 V, 4 A, 100 W
If the L.V. side is delivering full load (or rated) current at 0.8 p.f. lag and at 400 V, find the voltage
applied to H.V. side.
Solution
𝑉𝑠𝑐 60
𝑍𝑒𝐻 = = = 15𝛺
𝐼𝑠𝑐 4
67
2
𝑥𝑒𝐻 = √𝑍𝑒𝐻 2
− 𝑟𝑒𝐻 = √152 − 6.252 = 13.61𝛺
1 2 1 2
𝑟𝑒𝐿 = 𝑟𝑒𝐻 × (𝑘) = 6.25 (5) = 0.25𝛺
1 2 1 2
𝑥𝑒𝐿 = 𝑥𝑒𝐻 × (𝑘) = 13.61 (5) = 0.544𝛺
Equipment is desired to operate at a high efficiency. Fortunately, losses in transformers are small.
Because the transformer is a static device, there are no rotational losses such as windage and
friction losses in a rotating machine. In a well-designed transformer the efficiency can be as high
as 99%.
ii ohmic loss
Core loss The core loss Pc occurring in the transformer iron, consists of two components,
hysteresis loss Ph and eddy current loss Pe i.e.
Pc = Ph + Pe
The hysteresis and eddy current losses in the core can be expressed by :-
68
𝑥
𝑃ℎ = 𝐾ℎ 𝑓𝐵𝑚
𝑎𝑛𝑑 }
𝑃𝑒 = 𝐾𝑒 𝑓 2 𝐵 2 𝑚
Where Kh = proportionality constant which depends upon the volume and quality of the core
material and units used.
Ke = Proportionality constant whose value depends on the volume and resistivity of the core
material, thickness of laminations and the units employed
The value of the exponent x (called Steinmetz’s constant) varies from 1.5 to 2.5 depending upon
the magnetic properties of the core material. Therefore, the total core loss is
Pc = KhfB1.6m + Kef2B2m
Ohmic Loss When a transformer is loaded, ohmic loss (I2R) occurs in both the primary and
secondary winding resistances. Since the standard operating temperature of electrical machines is
750C. The ohmic loss should be calculated at 750C.
The efficiency of a transformer (or any other device) is defined as the ratio of the output power to
input power. Thus
𝑂𝑢𝑡𝑝𝑢𝑡𝑝𝑜𝑤𝑒𝑟
𝐸𝑓𝑓𝑖𝑐𝑖𝑒𝑛𝑐𝑦𝜂 =
𝐼𝑛𝑝𝑢𝑡𝑝𝑜𝑤𝑒𝑟
𝑉2 𝐼2 𝑐𝑜𝑠 𝜃2
𝜂=𝑉𝐼 2
2.4
2 2 𝑐𝑜𝑠 𝜃2 +𝑃𝑐 +𝐼 2 𝑅
69
Since the efficiencies of power and distribution transformers are usually very high, it is therefore,
more accurate to determine the efficiency from measurement of losses than from the measurement
of output.
In Eq. (*), Pc is constant and the load voltage V2 remains practically constant. A specified values
𝑑𝜂 𝑑𝜂
of load p.f. cos2, the efficiency will be maximum when 𝑑𝐼 = 0. Therefore, 𝑑𝐼 = 0 for Eq. (2.4)
2 2
is
𝑑𝜂 (𝑉2 𝐼2 𝑐𝑜𝑠 𝜃2 +𝑃𝑐 +𝐼22 𝑟𝑒2 )(𝑉2 𝑐𝑜𝑠 𝜃2 )−(𝑉2 𝐼2 𝑐𝑜𝑠 𝜃2 )(𝑉2 𝑐𝑜𝑠 𝜃2 +2𝐼2 𝑟𝑒2 )
= 2 =0
𝑑𝐼2 (𝑉2 𝐼2 𝑐𝑜𝑠 𝜃2 +𝑃𝑐 +𝐼22 𝑟𝑒2 )
Or (𝑉2 𝐼2 𝑐𝑜𝑠 𝜃2 + 𝑃𝑐 + 𝐼22 𝑟𝑒2 )(𝑉2 𝑐𝑜𝑠 𝜃2 ) = (𝑉2 𝐼2 𝑐𝑜𝑠 𝜃2 )(𝑉2 𝑐𝑜𝑠 𝜃2 + 2𝐼2 𝑟𝑒2 )
Hence the maximum efficiency occurs when the variable ohmic loss I 22 re 2 is equal to the fixed
core loss Pc. From Eq.(2.5) the load current I2 at which maximum efficiency occurs is given by
𝑃 𝑃
𝐼2 = √𝑟 𝑐 = 𝐼𝑓𝑙 √𝐼2 𝑟𝑐 2.6
𝑒2 𝑓𝑙 𝑒2
E2
If both sides of above equation are multiplied by , we get
1000
𝐸2 𝐼2 𝐸2 𝐼
𝑓𝑙 𝑐 𝑃
= 1000 √
1000 𝐹𝑢𝑙𝑙𝑙𝑜𝑎𝑑𝑜ℎ𝑚𝑖𝑐𝑙𝑜𝑠𝑠𝑒𝑠
𝐶𝑜𝑟𝑒𝑙𝑜𝑠𝑠
= (𝑟𝑎𝑡𝑒𝑑𝑡𝑟𝑎𝑛𝑠𝑓𝑜𝑟𝑚𝑒𝑟𝑘𝑉𝐴) × √
𝑂ℎ𝑚𝑖𝑐𝑙𝑜𝑠𝑠𝑒𝑠𝑎𝑡𝑟𝑎𝑡𝑒𝑑𝑐𝑢𝑟𝑟𝑒𝑛𝑡
𝑃
Or (𝑘𝑉𝐴)𝑚𝑎𝑥⋅𝜂 = (𝑘𝑉𝐴)√ 2 𝑐 2.7
𝐼 𝑟 𝑓𝑙 𝑒2
Thus, the maximum efficiency, for a constant load current, occurs at unity power factor (i.e. at
purely resistive load). It is seen from Eq. (2.6) that the load current at which maximum efficiency
70
occurs does not depend upon the load power factor because Pc and re2 are almost unaffected by a
variation in the load power factor.
A reduction in the load power factor reduces the transformer output and therefore the transformer
efficiency is also reduced accordingly. Figure 2.18 illustrates the effect of p.f. on efficiency. Note
that transformer efficiency is maximum at the same load current regardless of variation in the load
power factor.
100
Pf=1.0
Pf=0.8
75
Percent efficiency
Pf=0.2
50
max
25
0.5 1.0
Per unit load current
The manufacturer of transformers fixes a name plate on the transformer, on which are recorded
the rated output, the rated voltages, the rated frequency etc. of a particular transformer. A typical
name plate rating of a single-phase transformer is as follows: 20 KVA, 3300/220V, 50Hz. Here 20
KVA is rated output at the secondary terminals. Note that the rated output is expressed in kilo-
volt-ampere (KVA) rather than in kilowatt (KW). This is due to the fact that rated transformer
output is limited by the heating and hence by losses in the transformer. The two types of losses in
a transformer are core loss and ohmic ( I2r) loss. The core loss depends on transformer voltage and
ohmic loss on transformer current. As these losses depend on transformer voltage (V) and current
(I) and are almost unaffected by the load power factor, the transformer rated output is expressed
in VA (VI) or in kVA and are not in kW. For example, a transformer working on rated voltage
and rated current with load pf equal to zero has rated losses and rated kVA output but delivers zero
power to load. This shows that transformer must be expressed in kVA.
71
Rated input in kVA at Rated 0utput in kVA at
the primary ter min als (cos1 ) = the sec ondary ter min als (cos2 ) + Losses
Since a transformer operates at very high efficiency, losses may be ignored. Furthermore, the
primary power factor Cos1 and secondary power factor Cos2 is nearly equal. Therefore, the rated
KVA marked on the name plate of a transformer refers to both windings i.e. rated KVA of primary
winding and secondary winding are equal.
The voltage 3300/220 V refers to the design voltages of the two windings. Either one may serve
as primary or secondary. If it is step-down transformer, then 3300 V is rated primary voltage and
refers to the voltage applied to the primary winding. The 220V is rated secondary voltage and
refers to the voltage developed between output terminals at no-load with rated voltage applied to
the primary terminals.
Rated primary and secondary currents are calculated from the rated KVA and the corresponding
rated voltages. Thus
20,000
Rated (or full load) primary current = = 6.06𝐴
3300
20,000
Rated (or full load) secondary current = = 90.91𝐴
220
Note that the rated primary and secondary currents refer to the currents for which the windings are
designed.
Rated frequency refers to the frequency for which the transformer is designed to operate.
Example 2.6 A 100 kVA, 1000/10000 V, 50 Hz, single phase transformer has an iron loss of
1100 W. The copper loss with 5 A in the high voltage winding is 400 W. Calculate the efficiencies
at (i) 25 %, (ii) 50 % and (iii) 100 % of normal load for power factors of (a) 1.0 and (b) 0.8. The
output terminal voltage being maintained at 10000 V. Find also the load for maximum efficiency
at both power factors.
Solution
72
Secondary full load current, I2
100×1000
𝐼2 = 10000
= 10𝐴
2.5 2
Copper losses at 25% full load =( 5 ) × 400
= 100 W
= 25 000 W
= 95.4%
= 94.34%
= 50 000 W
= 97.65%
= 97.1%
73
Efficiency at 100 % full load, unity pf:
10 2
Copper losses at 100% full load =( 5 ) × 400
= 1600 W
= 97.37%
= 96.73%
x = 0.829
Load for maximum efficiency = 0.829 100 = 82.9 kVA
Example 2.7 A single phase transformer working at unity power factor has an efficiency of 90%
at both one-half load and at the full load of 500 W. Determine the efficiency at 75 % of full load.
Solution
74
Let the iron losses of the transformer be = x watts
Hence,
500
0.9 = 500+𝑥+𝑦
= 10𝐴
1 2 𝑦
Total copper losses at half of full load =(2) × 𝑦 = 4
250
Thus, 0.9 = 𝑦
250+𝑥+ ⁄4
y = 37 W
and x = 18.53 W
= 0.75 x 500
= 375 W
= (0.75)2 x 37
= 20.8 W
75
Efficiency at 75 percent of full load
375
𝜂𝑎𝑡75% = 375+18.53+20.8 × 100
= 90.5%
When electric power is supplied to a locality a single transformer, capable of handling the required
power demand, is installed. In some cases, it may be preferable to install two or more transformers
in parallel, instead of one large unit. Though two or more transformers may be expensive than one
large unit, yet this scheme possess certain advantages described below.
1. With two or more transformers, power system becomes more reliable. For instance, if one
transformer develops a fault, it can be removed and the other transformers can maintain the
flow of power, though at reduced load.
2. Transformers can be switched on or off, depending upon the power demand. In this manner,
the transformer losses decrease and the system becomes more economical and efficient in
operation.
3. The cost of standby (or spare) unit is much less when two or more transformers are
installed.
In any case, in the long run, electric power demand may become more than rated KVA capacity
of already existing transformer or transformers. Under such circumstances, the need for extra
transformer arises; the extra unit must be connected in parallel.
Note that the parallel operation of transformers requires that their primary windings, as well as
secondary windings are connected in parallel. In this section only the parallel operation of single-
phase transformers is considered.
The various conditions which must be fulfilled for the satisfactory parallel operation of two or
more single-phase transformers are as follows:
a. The transformer must have the same voltage ratios, i.e with the primaries connected to the same
voltage sources, the secondary voltage of all transformers should be equal in magnitude.
76
b. The equivalent leakage impedance in ohms must be inversely proportional to their respective
KVA ratings. In other words, per unit (pu) leakage impedance of transformers based on their
KVA rating must be equal.
c. The ratio of equivalent leakage reactance to equivalent resistance i.e. Xe/re should be the same for all
transformers.
d. The transformer must be connected properly as far as their polarities are concerned.
Out of the conditions listed above, condition(d) must be strictly fulfilled. If the secondary terminals
are connected with wrong polarities, large circulating currents will flow and the transformers may
get damaged.
Condition (a) should be satisfied as accurately as possible; since different secondary voltages
would give rise to undesired circulating currents. For conductions (b) and (c), some deviation is
permissible. Thus, the fulfillment of condition (d) is essential whereas the fulfillment of the other
conditions is desirable.
A1 A2 A1 A2
A B
+ - + -
a1 a2 a1 a2
V
Figure 2.19 shows single-phase transformers in parallel, connected to some voltage source on the
primary side. Zero voltmeter reading indicates proper polarities. If the voltmeter reads the sum of
two secondary voltages, the polarities are improper and can be corrected by reversing the
secondary terminals of any one of the transformers.
Generation, transmission and distribution of electric energy is invariably done through the use of
three-phase systems because of its several advantages over single-phase systems. As such, a large
number of three-phase transformers are inducted in a 3-phase energy system for stepping-up or
stepping – down the voltage as required. For 3-phase up or down transformation, three units of 1-
77
phase transformers or one unit of 3-phase transformer may be used. When three identical units of
1-phase transformers are used as shown in Figure 2.20(a), the arrangement is usually called a bank
of three transformers or a 3-phase transformer bank. A single 3-phase transformer unit may
employ 3–phase core-type construction Figure 2.20(b) or three phase shell type construction.
Input
A B C
P P P
P S P S P S
I II III
S S S
a b c
Output
(a) (b)
Figure 2.20 (a)Three-phase transformer bank, both windings in star;(b) three-phase core-type
transformer
A single-unit 3-phase core-type transformer uses a three-limbed core, one limb for each phase
winding as shown in Figure 2.20(b). Actually, each limb has the L.V. winding placed adjacent to
the laminated steel core and then H.V. winding is placed over the 1.v. winding. Appropriate
insulation is placed in between the core and 1.v. winding and also in between the two windings.
A 3-phase core-type transformer costs about 15% less than a bank of three 1-phase transformers.
Also, a single unit occupies less floor space than a bank.
These connections are shown in Figures 2.21 and 2.22, where V and I are taken as input line
voltage and line current respectively. Primary and secondary windings of one phase are drawn
parallel to each other. With phase turns ratio from primary to secondary as
78
N1/N2= a, the voltages and current in the windings and lines are shown in Figures 2.21 and 2.22.
The various connections are now described briefly.
(a) Star-delta (Y-) Connection This connection is commonly used for stepping down the voltage
from a high level to a medium or low level. The insulation on the h.v. side of the transformer is
𝑁 𝑉 𝑉
Secondary phase voltage, 𝑉2 = 𝑁2 . = 𝑎.√3
1 √3
𝑉
Secondary line voltage = secondary phase voltage = 𝑎.√3
𝑉 𝑉
Input VA = 3 . I = output VA = 3.𝑎.√3 . 𝑎𝐼 = √3𝑉𝐼
√3
Phase and line values for voltages and currents on both primary and secondary sides of star-delta
transformer are shown in Figure 2.21(a)
I 3aI aI
I 3
V
V I
3 3a aI V
V aI I 3 a 3V
V
3 a
(a) (b)
Figure 2.21 (a) Star-delta connection and (b) delta-star connection of 3-phase transformers
79
(b) Delta-Star (-Y) connection: - This type of connection is used for stepping up the voltage to a
high level. For example, these are used in the beginning of h.v. transmission lines so that insulation
is stressed to about 57.74% of line voltage
Delta-star transformers are also generally used as distribution transformers for providing mixed
line to line voltage to high-power equipment and line to neutral voltage to 1-phase low-power
equipment. For example, 11kV/400V, delta-star distribution transformer is used to distribute
power to consumers by 3-phase four-wire system. Three-phase high–power equipment is
connected to 400V, three-line wires, whereas 1-phase low-power equipment is energized from 231
V line to neutral circuits.
Phase and line values for voltages and currents on primary as well as secondary sides of a 3-phase
delta-star transformer are shown in Figure 2.21(b).
In case a bank of three transformers is used, then one transformer can be removed for maintenance
purposes while the remaining two transformers (called an open-delta or V-connection) can still
deliver 58% of the power delivered by the original 3-phase transformer bank.
Phase and line value for voltages and currents on both primary and secondary sides of a 3-phase
delta-delta transformer is shown in Figure 2.22 (a).
I kI kI
V V I V
k 3 kI V
I kI V 3k
V k
3 3
(a) (b)
Figure 2.22 (a) Delta-delta connection and (b) Star-star connection of three-phase transformers.
Phase and line values of voltages and currents on both sides of a star-star transformer are shown
in Figure 2.22(b)
81
Example 2.8 A 3-phase transformer is used to step-down the voltage of a 3-phase, 11kV feeder
line. Per-phase turns ratio is 12. For a primary line current of 20A, calculate the secondary line
voltage, line current and output KVA for the following Connections:
(a) star-delta (b) delta-star (c) delta-delta (d) star-star. Neglect losses.
Solution
(a) Three-phase transformer with star-delta connection is shown in Figure 2.23(a)
𝑉𝐿1 11000
𝑝ℎ𝑎𝑠𝑒 𝑣𝑜𝑙𝑡𝑎𝑔𝑒 𝑜𝑛 𝑝𝑟𝑖𝑚𝑎𝑟𝑦, 𝑉𝑃1 = = 𝑉
√3 √3
𝑝ℎ𝑎𝑠𝑒 𝑐𝑢𝑟𝑟𝑒𝑛𝑡 𝑜𝑛 𝑝𝑟𝑖𝑚𝑎𝑟𝑦, 𝐼𝑃1 = 𝐼𝐿1 = 20𝐴
𝑉 𝑉
𝐻𝑒𝑟𝑒, 𝑃1 = 𝑃2 𝑎𝑛𝑑𝐼𝑝1 × 12 = 𝐼𝑝2 × 1
12 1
11000
∴ 𝑃ℎ𝑎𝑠𝑒 𝑣𝑜𝑙𝑡𝑎𝑔𝑒 𝑜𝑛 𝑠𝑒𝑐𝑜𝑛𝑑𝑎𝑟𝑦, 𝑉𝑝2 = = 529.25𝑉
√3𝑥12
𝐿𝑖𝑛𝑒 𝑣𝑜𝑙𝑡𝑎𝑔𝑒 𝑜𝑛 𝑠𝑒𝑐𝑜𝑛𝑑𝑎𝑟𝑦, 𝑉𝐿2 = 𝑉𝑝2 = 529.25𝑉
𝑃ℎ𝑎𝑠𝑒 𝑐𝑢𝑟𝑟𝑒𝑛𝑡 𝑜𝑛 𝑠𝑒𝑐𝑜𝑛𝑑𝑎𝑟𝑦, 𝐼𝑝2 = 12𝐼𝑝1 = 12 × 20 = 240𝐴
𝐿𝑖𝑛𝑒 𝑐𝑢𝑟𝑟𝑒𝑛𝑡 𝑜𝑛 𝑠𝑒𝑐𝑜𝑛𝑑𝑎𝑟𝑦, 𝐼𝐿2 = √3𝐼𝑝2 = √3 × 240 = 415.68𝐴
3𝑉𝑝2 .𝐼𝑝2 11000 1
𝑂𝑢𝑡𝑝𝑢𝑡𝐾𝑉𝐴 = = 3. × 240𝑥 1000 = 381.04𝐾𝑉𝐴
1000 √3𝑥12
82
VL1 11000 Vp1
Phase voltage on sec ondary, VP 2 = = V = 916.7V =
12 12 12
Line voltage on sec ondary, VL2 = VP 2 = 916.7V
20
Phase current on primary, I P1 = A
3
20
Phase current on sec ondary, I P 2 = 12 I P1 = 12 A
3
12 20
Line current on sec ondary, I L2 = 3 I P 2 = 3 . = 240A
3
11000 12x 20 1
Output KVA = 3 = 381.04KVA .
12 3 1000
I L1 = 20A I L2
12:1
I L1 = 20A
VP1 I P1 VP 2 = VL2 12:1
VL1 = 11000 V I P2
I P2 VP 2
VL1 = 11000 V I P1 VL2
(a)
(b)
I L1 = 20A I L2 = 3I P2 I L2
12:1 12:1
VP 2 = VL2 VP1 I P1
VP2
VL1 = 11000 V I P2
VP1 = 11000 V I P2 VL2 = 3VP2
I P1
(c) (d)
83
Example 2.9 An 11000/415V, delta-star transformer feeds power to a 30 kW, 415V, 3-phase
induction motor having an efficiency of 90% and full-load pf 0.833. Calculate the transformer
rating and phase and line currents on both high and low voltage sides.
30
Transformer kVA rating = = 40KVA
0.9 x 0.833
Total load in VA 40,000
Solution Line current on l.v. side of transformer = = = 55.65A
3 x line voltage 3 x 415
.
For star connected 1.v. winding, phase current in 1.v. winding = line current on 1.v side = 55.65A.
40,000
Line current on HV, side of transformer = = 2.1A
3 x 11000
For delta connected HV winding, phase current in HV winding:
1 1
= (line current on h.v. side) = x 2.1 = 1.212A
3 3
Phase Shift
Some of the three-phase transformer connections will result in a phase shift between the primary
and secondary line-voltages. Consider the phase voltages, shown in Figure 2.24, for the Y-
connections. The phases VAN and Va are aligned, but line voltage VAB of the primary leads the line
voltage Vab of the secondary by 300. It can be shown the -Y connection also provides a 300 phase
shift in there line –to-line voltage. This property of phase shift in Y- or -Y connections can be
used advantageously in some applications.
a
Vc
c Va = Vab
VAN VAB
VCN
Vb
VBN B b
C
VAB
VAN Va Vab
-VBN
84
V-Connection
It was stated earlier that in the - connection of three single-phase transformers, one transformer
can be removed and the system can still deliver three-phase power to a three-phase load. This
configuration is known as an open-delta or V connection. It may be employed in an emergency
situation when one transformer must be removed for repair and continuity of service is required.
A a Ia
c
C
Ib
n
B b
Ic
(a)
VAB Vab
Van Ia
Vcb
30
Ic
Vcn
30
(b)
Consider Figure 2.25(a) in which one transformer, shown dotted is removed. For simplicity the
load is considered to by Y- connected. Figure 2.25(b) shows the phase diagram for voltages and
currents. Here VAB, VBC and VCA represent the line-to line voltage of the primary Vab, Vbc and Vcb
secondary and Van, Vbn and Vcn represent the phase voltages of the load. For an inductive load the
load currents Ia, Ib and Ic will lag the corresponding voltages Van, Vbn and Vcn by the load phase
angle θ.
85
Ia = Ic = I current rating of the transformer secondary winding and = 0 for a resistive
load. Power delivered to the load by the V connection is
P = 3VI 2.9
The V connection is capable of delivering 58% power without overloading the transformer (i.e.,
not exceeding the current rating of the transformer winding).
II b c
a + b + c = 0
(a)
a
b
(b) (c)
86
(d)
However, if section II is pushed in between sections I and III by removing its yokes, a common
magnetic structure shown in Figure 2.26(c), is obtained.
This core structure can be built using stacked laminations as shown in Figure 2.26(d). Both primary
and secondary windings of a phase are placed on the same leg. Note that the magnetic paths of
legs a and c are somewhere longer than that of leg b (Figure 2.26 (c). This will result in some
imbalance in the magnetizing currents. However, this imbalance is not significant.
Figurer 2.27 shows a picture of a three-phase transformer of this type. Such a transformer weight
less , costs less, and requires less space than a three-phase transformer bank of the same rating.
The disadvantage is that if one phase breaks down, the whole transformer must be removed for
repair.
Transformers are usually air-cooled even if placed in metal cases. Larger sizes are placed in tanks
with special transformer oil. The oil has a dual function; it insulates while providing cooling. Still
87
larger sizes have tanks with corrugated sides or cooling fins or radiators to dissipate the heat to the
surrounding air. Figure 2.28 shows a typical self-cooled transformer. The oil moves around by
natural convection, since warmer oil flows up. It flows down again through the radiator, which
gives up this heat to the surrounding air. In larger oil-cooled units, the oil must be pumped around
to maintain acceptable temperature levels. No matter what size of transformer is dealt with, they
all operate on the same principle.
2.16 Autotransformers
In principle and in general construction, the autotransformer does not differ from the conventional
two-winding transformer so far discussed. It does differ from it. however, in the way the primary
and secondary windings are interrelated.
It will be recalled that in discussing the transformer principles of operation, it was pointed out that
a counter emf was induced in the winding, which acted as a primary to establish the excitation
ampere turns. The induced voltage per turn was the same in each and every turn linking with the
common flux of the transformer. Therefore, fundamentally it makes no difference in the operation
whether the secondary induced voltage is obtained from a separate winding linked with the core
or from a portion of the primary turns. The same voltage transformation results in the two
88
situations. When the primary and secondary voltage are derived from the same winding. the
transformer is called an autotransformer.
a
a
I1
I2 b
Load
b V2
V1 I1
I2 - I1
Load
V2 RL
I 2 - I1
c
c
(a) (b)
In Figure 2.29(a) the input voltage V1 is connected to the complete winding (a-c) and the load RL
is connected across a portion of the winding, that is, (b-c).The voltage V2 is related to V1 as in the
conventional two-winding transformer, that is,
𝑁
𝑉2 = 𝑉1 × 𝑁𝑏𝑐 2.10
𝑎𝑐
where Nbc and Nac are the number of turns on the respective windings. The ratio of voltage
transformation in an autotransformer is the same as that for an ordinary transformer, namely,
𝑁 𝑉 𝐼
𝑘 = 𝑁𝑎𝑐 = 𝑉1 = 𝐼2 2.11
𝑏𝑐 2 1
89
𝑃 = 𝑉2 𝐼2 2.12
Note that I1 flows in the portion of winding ab, whereas the current (I2 – I1) flows in the remaining
portion bc.The resulting current flowing in the winding bc is always the arithmetic difference
between I1 and I2 , since they are always in opposite sense. Remember that the induced voltage in
the primary opposes the primary voltage. As a result. the current caused by the induced voltage
flows opposite to the input current.
𝐼1 + (𝐼2 − 𝐼1 ) = 𝐼2 2.13
𝑁𝑎𝑐
Hence the ampere-turns due to section bc, where the substitutions 𝐼2 = 𝑘𝐼1 and 𝑁𝑏𝑐 = 𝑘
are made
Thus the ampere-turns due to sections bc and ab balance each other, a characterstic of all
transformer actions.
Equation (2.12) gives the power determined by the load. To see how this power is delivered, we
can write the equation in a slightly modified forn. By substituting Eq. (2.13) into Eq. (2.12), we
obtain
𝑃 = 𝑉2 𝐼2 = 𝑉2 [𝐼1 + (𝐼2 − 𝐼1 )]
= 𝑉2 𝐼1 + 𝑉2 (𝐼2 − 𝐼1 )𝑊 2.14
This indicates that the load power consists of two parts. The first part is
𝑃𝑐 = 𝑉2 𝐼1 ≡ 𝑐𝑜𝑛𝑑𝑢𝑐𝑡𝑒𝑑𝑝𝑜𝑤𝑒𝑟𝑡𝑜𝑙𝑜𝑎𝑑𝑡ℎ𝑟𝑜𝑢𝑔ℎ𝑎𝑏 2.15
90
We will see in the following examples that most of the power to the load is directly conducted by
winding ab. The remaining power is transferred by the common winding bc. To show these powers
Pc and Ptr in terms of the total power P. ue proceed as follows:
𝑃𝑐 𝑉𝐼 𝐼 1
= 𝑉2𝐼1 = 𝐼1 = 𝑘
𝑃 2 2 2
and
𝑃𝑡𝑟 𝑉2 (𝐼2 −𝐼1 ) (𝐼2 −𝐼1 ) 𝑘−1
= = =
𝑃 𝑉2 𝐼2 𝐼2 𝑘
𝑃 𝑃(𝑘−1)
Thus 𝑃𝑐 = 𝑘 and 𝑃𝑡𝑟 = with a > 1 for a step-down autotransformer.
𝑘
Solution
Therefore,
𝐼2 = 2.174 + 𝐼1 = 23.914𝐴
91
and that transformed is
(𝑘−1) 1.1−1
𝑃𝑡𝑟 = 𝑃 = 55,000 = 5.0𝑘𝑉𝐴
𝑘 1.1
Consider now the step-up transformer of Figure 2.29(b). Following reasons similar to those above,
it follows that
𝑃 = 𝑉1 𝐼1 = 𝑉1 [𝐼2 + (𝐼1 − 𝐼2 )]
= 𝑉1 𝐼2 + 𝑉1 (𝐼1 − 𝐼2 )𝑊 2.17
where we made the substitution of I1 from Eq. (2.13), which really is kirchhoff's current law
applied to point b. To show this, note at point b we have
𝐼1 + (𝐼2 − 𝐼1 ) = 𝐼2
so that
𝐼1 = 𝐼2 − (𝐼2 − 𝐼1 ) = 𝐼2 + (𝐼1 − 𝐼2 )
Again, Eq. (2.17) shows us that the power supplied to the load consists of two parts,
and
and
𝑃𝑡𝑟 𝑉1 (𝐼1 −𝐼2 ) (𝐼1 −𝐼2 )
= = = 1−𝑘 2.21
𝑃 𝑉1 𝐼1 𝐼1
Thus for the step-up transformer with a < 1, we obtain 𝑃𝑐 = 𝑘𝑃 and 𝑃𝑡𝑟 = 𝑃(1 − 𝑘) As before, Pc
is the power directly conducted to the load and Ptr is the portion that is transformed.
Example 2.11 Repeat the problem of Example 2.10 for a 2300 V-to-2530 V step-up connection as
shown in Figure 2.29 (b).
Solution
92
As calculated in Example 2.10, the current rating of the winding ab is I2 = 21.74 A, which also is
the load current. The output voltage is 2530 V; thus the volt-ampere rating of the autotransformer
is
The examples given make it clear that an autotranaformer of given physical dimensions can handle
𝑘
much more load power than an equivalent two-winding transformer; in fact,1−𝑘 times its rating as
1
two-winding transformer for the step-down autotransformer or 𝑘−1 for the step-up arrangement. A
5-kVA transformer is capable of taking care of 11 times its rating. These great gains are possible
since an autotransformer transforms, by transformer action, only a fraction of the total power; the
power that is not transformed is conducted directly to the load.
It should be noted that an autotransformer is not suitable for large percentage voltage reductions
as is a distribution transformer. This is due to the required turns ratio becoming too large; hence
the power-handling advantage would be minimal. Furthermore, in the unlikely but possible
event that the connections to the low-voltage secondary were to fail somewhere below point b in
Figure 2.29(a), the winding bc would be deleted from the circuit. This implies that the load would
see the full high line voltage. Autotransformers are not used for these reasons where large voltage
changes are encountered. In situations where autotransformers can be used to their full advantage,
it will be found that they are cheaper than a cowentional two-winding transformer of similar rating.
They also have better regulation (i.e., the voltage does not drop so much for the same load), and
they operate at higher efficiency. In all applications using autotransformer it should be realized
that the primary and secondary circuits are not electrically isolated, since one input terminal is
common with one output terminal.
93
Objective Tests on Single-Phase Transformer
1. A transformer transforms
(a) frequency (c) current
(b) voltage (d) voltage and current
2. Which of the following is not a basic element of a transformer?
(a) core (c) secondary winding
(b) primary winding (d) mutual flux
3. In an ideal transformer,
(a) windings have no resistance (c) core has infinite permeability
(b) core has no losses (d) all of the above
4. The main purpose of using core in a transformer is to
(a) decrease iron losses (d) decrease reluctance of the common
(b) prevent eddy current loss magnetic circuit
95
(a) copper loss (c) magnetizing current and no-load loss
(b) magnetizing current (d) efficiency of the transformer
18. The main purpose of performing open-circuit test on a transformer is to measure its
(a) Cu loss (c) total loss
(b) core loss (d) insulation resistance
19. During short-circuit test, the iron loss of a transformer is negligible because
(a) the entire input is just sufficient to meet nor- mal flux
Cu losses only (c) iron core becomes fully saturated
(b) flux produced is a small fraction of the (d) supply frequency is held constant
20. The iron loss of a transformer at 400 Hz is 10W. Assuming that eddy current and hysteresis
losses vary as the square of flux density, the iron loss of the transformer at rated voltage but at 50
Hz would be…. watt.
(a) 80 (b) 640 (c) 1.25 (d) 100
21. In operating a 400 Hz transformer at 50 Hz
(a) only voltage is reduced in the same (c) both voltage and kVA rating are reduced
proportion as the frequency in the same proportion as the frequency
(b) only kVA rating is reduced in the same (d) none of the above
proportion as the frequency
22. The voltage applied to the h.v. side of a transformer during short-circuit test is 2% of its rated
voltage. The core loss will be…. percent of the rated core loss.
(a) 4 (b) 0.4 (c) 0.25 (d) 0.04
23. Transformers are rated in kVA instead of kW because
(a) load power factor is often not known (c) total transformer loss depends on volt-
(b) kVA is fixed whereas kW depends on ampere
load p.f. (d) it has become customary
24. When a 400-Hz transformer is operated at 50 Hz its kVA rating is
(a) reduced to 1/8 (c) unaffected
(b) increased 8 times (d) increased 64 times
25. At relatively light loads, transformer efficiency is low because
96
(a) secondary output is low output
97
(c) energy is transferred both inductively (d) secondary current is increased
and conductivity
34. The saving in Cu achieved by converting a 2-winding transformer into an autotransformer is
determined by
(a) voltage transformation ratio (c) magnetic quality of core material
(b) load on the secondary (d) size of the transformer core
35. An autotransformer having a transformation ratio of 0.8 supplies a load of 3 kW. The power
transferred conductively from primary to secondary is….kW.
(a) 0.6 (b) 2.4 (c) 1.5 (d) 0.27
36. The essential condition for parallel operation of two 1-ϕ transformers is that they should have the
same
(a) polarity (c) voltage ratio
(b) kVA rating (d) percentage impedance
37. If the impedance triangles of two transformers operating in parallel are not identical in shape and
size, the two transformers will
(a) share the load unequally even when unloaded
(b) get heated unequally (d) run with different power factors
(c) have a circulatory secondary current
38. Two transformers A and B having equal outputs and voltage ratios but unequal percentage
impedances of 4 and 2 are operating in parallel. Transformer A will be running over-load by .......
percent.
(a) 50 (b) 66 (c) 33 (d) 5
Objective Tests on Three-Phase Transformer
1. Which of the following connections is best suited for 3-phase, 4-wire service?
(a) ∆-∆ (b) Y-Y (c) ∆-Y (d) Y-∆
2. In a three-phase Y-Y transformer connection, neutral is fundamental to the
(a) suppression of harmonics (c) provision of dual electric service
(b) passage of unbalanced currents due to (d) balancing of phase voltages with
unbalanced loads respect to line voltages
98
3. As compared to ∆-∆ bank, the capacity of the bank of transformers is…. percent.
(a) 57.7 (b) 66.7 (c) 50 (d) 86.6
4. If three transformers in a ∆-∆ are delivering their rated load and one transformer is removed, then
overload on each of the remaining transformers is… . percent.
(a) 66.7 (b) 173.2 (c) 73.2 (d) 58
5. When a V-V system is converted into a ∆-∆ system, increase in capacity of the system is
……. percent.
99
12. Statement: An auto-transformer is more efficient in transferring energy from primary to
secondary circuit. Reason: Because it does so both inductively and conductively.
Key
(a) statement is false, reason is correct and (c) both statement and reason are correct
relevant and are connected to each other as cause
(b) statement is correct, reason is correct and effect
but irrelevant (d) both statement and reason are false
13. Out of the following given choices for poly phase transformer connections which one will you
select for three-to-two phase conversion?
(a) Scott (c) double Scott
(b) star/star (d) star/double-delta
14. A T-T transformer cannot be paralleled with…. transformer.
(a) V-V (b) Y-∆ (c) Y-Y (d) ∆-∆
15. Instrument transformers are used on a.c. circuits for extending the range of
(a) ammeters (c) wattmeter
(b) voltmeters (d) all of the above
16. Before removing the ammeter from a current transformer, its secondary must be short- circuited
in order to avoid
(a) excessive heating of the core (c) increase in iron losses
(b) high secondary e.m.f. (d) all of the above
100
Chapter three
Induction Machine
Contents of this chapter includes:
Revolving field, construction, synchronous speed & slip, rotor & equivalent circuit models,
determine parameter of the equivalent circuit, torque equation, stall & starting torque, efficiency,
torque-speed curves and parameter measurement.
3.1 Background
As a general rule, conversion of electrical power into mechanical power takes place in the rotating
part of an electric motor. In dc motors, the electric power is conducted directly to the armature
(i.e., rotating part) through brushes and commutator. Hence, in this sense, a dc motor can be called
a conduction motor. However, in ac motors, the rotor does not receive electric power by conduction
but by induction in exactly the same way as the secondary of a 2-winding transformer receives its
power from the primary. That is why such motors are known as induction motors. In fact, an
induction motor can be treated as a rotating transformer i.e., one in which primary winding is
stationary but the secondary is free to rotate.
Figure 3.1: Squirrel cage AC induction motor opened to show the stator and rotor
construction, the shaft with bearings, and the cooling fan.
Of all the ac motors, the polyphase induction motor is the one which is extensively used for various
kinds of industrial drives. It has the following main advantages and also some dis-advantages:
Advantages:
i It has very simple and extremely rugged, almost unbreakable construction (especially squirrel cage
type).
101
ii Its cost is low and it is very reliable.
iii It has sufficiently high efficiency. In normal running condition, no brushes are needed, hence
frictional losses are reduced. It has a reasonably good power factor.
iv It requires minimum of maintenance.
v It starts up from rest and needs no extra starting motor and has not to be synchronized. Its starting
arrangement is simple especially for squirrel-cage type motor.
Disadvantages:
An induction motor consists essentially of two main parts: (a) a stator and (b) a rotor.
(a) Stator
The stator of an induction motor is, in principle, the same as that of a synchronous motor or
generator. It is made up of a number of stampings, which are slotted to receive the windings Fig.
3.2(a). The stator carries a 3-phase winding Fig.3.2 (b) and is fed from a 3-phase supply. It is
wound for a definite number of poles; the exact number of poles being determined by the
requirements of speed. Greater the number of poles, lesser the speed and vice versa. It will be
shown that the stator windings, when supplied with 3-phase currents, produce a magnetic flux,
which is of constant magnitude but which revolves (or rotates) at synchronous speed (given by Ns
= 120 f/P). This revolving magnetic flux induces an e.m.f. in the rotor by mutual induction.
(a) (b)
Fig. 3.2 (a) Unwound stator with semi-closed slots& (b) Completely wound stator
Laminations are of high-quality low-loss silicon steel. for an induction motor
102
(b) Rotor
i Squirrel-cage rotor: Motors employing this type of rotor are known as squirrel-cage
induction motors.
ii Phase-wound or wound rotor: Motors employing this type of rotor are variously known as
‘phase-wound’ motors or ‘wound’ motors or as ‘slip-ring’ motors.
Squirrel-cage rotor
Almost 90 per cent of induction motors are squirrel-cage type, because this type of rotor has the
simplest and most rugged construction imaginable and is almost indestructible. The rotor consists
of a cylindrical laminated core with parallel slots for carrying the rotor conductors which, it should
be
(a) (b)
Figure 3.3: (a) Squirrel-cage rotor with copper bars and alloy brazed end-rings &(b) Rotor with
shaft and bearings
noted clearly, are not wires but consist of heavy bars of copper, aluminum or alloys. One bar is
placed in each slot, rather the bars are inserted from the end when semi-closed slots are used. The
rotor bars are brazed or electrically welded or bolted to two heavy and stout short-circuiting end-
rings, thus giving us, what is so picturesquely called, a squirrel-case construction (Fig. 3.3). It
should be noted that the rotor bars are permanently short-circuited on themselves, hence it is not
possible to add any external resistance in series with the rotor circuit for starting purposes. The
rotor slots are usually not quite parallel to the shaft but are purposely given a slight skew (Fig.
3.4). This is useful in two ways: (i) it helps to make the motor run quietly by reducing the magnetic
hum and (ii) it helps in reducing the locking tendency of the rotor i.e. the tendency of the rotor
teeth to remain under the stator teeth due to direct magnetic attraction between the two. In small
motors, another method of construction is used. It consists of placing the entire rotor core in a
mould and casting all the bars and end-rings in one piece. The metal commonly used is an
aluminium alloy. Another form of rotor consists of a solid cylinder of steel without any conductors
or slots at all. The motor operation depends upon the production of eddy currents in the steel rotor.
103
Figure 3.4: skewed rotor bar structure
This type of rotor is provided with 3-phase, double-layer, distributed winding consisting of coils
as used in alternators. The rotor is wound for as many poles as the number of stator poles and is
always wound 3-phase even when the stator is wound two-phase. The three phases are starred
internally. The other three winding terminals are brought out and connected to three insulated
slip-rings mounted on the shaft with brushes resting on them [Fig.3.5a]. These three brushes are
further externally connected to a 3-phase star-connected rheostat [Fig.3.5b]. This makes possible
the introduction of additional resistance in the rotor circuit during the starting period for
increasing the starting torque of the motor, as shown in Fig. 3.6 and for changing its speed-
torque/current characteristics. When running under normal conditions, the slip-rings are
automatically short-circuited by means of a metal collar, which is pushed along the shaft and
connects all the rings together. Next, the brushes are automatically lifted from the slip-rings to
reduce the frictional losses and the wear and tear. Hence, it is seen that under normal running
conditions, the wound rotor is short-circuited on itself just like the squirrel-case rotor.
(a) (b)
Figure 3.5: (a) Slip-ring motor with slip-rings, brushes and short-circuiting devices&(b) three
phase star-connected rheostat
104
Figure 3.6: slip rings connected with external resistors
Components of the induction machines
1. Frame. Made of close-grained alloy cast iron.
2. Stator and Rotor Core. Built from high-quality low-loss silicon steel laminations
and flash-enamelled on both sides.
3. Stator and Rotor Windings. Have moisture proof tropical insulation embodying mica
and high-quality varnishes. Are carefully spaced for most effective air circulation and
are rigidly braced to withstand centrifugal forces and any short-circuit stresses.
4. Air-gap. The stator rabbets and bore are machined carefully to ensure uniformity of
air-gap.
5. Shafts and Bearings. Ball and roller bearings are used to suit heavy duty, toruble-free
running and for enhanced service life.
6. Fans. Light aluminium fans are used for adequate circulation of cooling air and are
securely keyed onto the rotor shaft.
7. Slip-rings and Slip-ring Enclosures. Slip-rings are made of high-quality phosphor-
bronze and are of moulded construction.
Fig. 3.7 (a) shows the disassembled view of an induction motor with squirrel-cage rotor.
According to the labelled notation (a) represents stator (b) rotor (c) bearing shields (d) fan (e)
ventilation grill and (f) terminal box. Similarly, Fig. 3.7 (b) shows the disassembled view of a
slip-ring motor where (a) represents stator (b) rotor (c) bearing shields (d) fan (e) ventilation grill
(f) terminal box (g) slip-rings(h) brushes and brush holders.
(a) (b)
105
Figure 3.7:(a) disassembled view of an induction motor with squirrel-cage rotor&(b) the
disassembled view of a slip-ring motor
Applications: Squirrel cage induction motor is used in lathes, drilling machine, fan, blower
printing machines etc.
106
✓ The construction is simple and robust and it is cheap as compared to slip ring induction
motor
✓ Due to its simple construction and low cost. The squirrel cage induction motor is widely
used
✓ Less rotor copper losses and hence high efficiency
✓ Speed control by rotor resistance method is not possible
✓ Squirrel cage induction motor is used in lathes, drilling machine, fan, blower printing
machines etc.
Stator of three phase induction motor is made up of numbers of slots to construct a 3-phase winding
circuit which is connected to 3 phase AC source. The three-phase winding are arranged in such a
manner in the slots that they produce a rotating magnetic field after three phase AC supply is given
to them.
Rotor of three phase induction motor consists of cylindrical laminated core with parallel slots that
can carry conductors. Conductors are heavy copper or aluminum bars which fits in each slots &
they are short circuited by the end rings. The slots are not exactly made parallel to
the axis of the shaft but are slotted a little skewed because this arrangement reduces magnetic
humming noise & can avoid stalling of motor.
The stator of the motor consists of overlapping winding offset by an electrical angle of 120°. When
the primary winding or the stator is connected to a 3 phase AC source, it establishes a rotating
magnetic field which rotates at the synchronous speed. Secrets Behind the Rotation: According to
Faraday’s law an emf induced in any circuit is due to the rate of change of magnetic flux linkage
through the circuit. As the rotor winding in an induction motor are either closed through an external
resistance or directly shorted by end ring, and cut the stator rotating magnetic field, an emf is
induced in the rotor copper bar and due to this emf a current flow through the rotor conductor.
Here the relative speed between the rotating flux and static rotor conductor is the cause of current
generation; hence as per Lenz's law the rotor will rotate in the same direction to reduce the cause
i.e. the relative velocity.
107
120𝑓𝑒
𝑛𝑠𝑦𝑛𝑐 = (3.1)
𝑃
where fe is the system frequency in hertz and P is the number of poles in the machine.
This rotating magnetic field Bs passes over the rotor bars and induces a voltage in them.
𝑒𝑖𝑛𝑑 = (𝑣 × 𝐵) ∙ 𝑙 (3.2)
It is the relative motion of the rotor compared to the stat or magnetic field that produces induced
voltage in a rotor bar. The velocity of the upper rotor bars relative to the magnetic field is to the
right, so the induced voltage in the upper bars is out of the page, while the induced voltage in the
lower bars is into the page. This results in a current flow out of the upper bars and into the lower
bars. However, since the rotor assembly is inductive, the peak rotor current lags behind the peak
rotor voltage (see Figure 3.8b). The rotor current flow produces a rotor magnetic field BR.
the resulting torque is counterclockwise. Since the rotor induced torque is counterclockwise, the
rotor accelerates in that direction.
There is a finite upper limit to the motor's speed, however. If the induction motor's rotor were
turning at synchronous speed, then the rotor bars would be stationary relative to the magnetic field
and there would be no induced voltage. If 𝑒𝑖𝑛𝑑 were equal to 0, then there would be no rotor current
and no rotor magnetic field. With no rotor magnetic field, the induced torque would be zero, and
the rotor would slow down as a result of friction losses. An induction motor can thus speed up to
near-synchronous speed, but it can never exactly reach synchronous speed.
Note that in normal operation both the rotor and stator magnetic fields BR and Bs rotate together at
synchronous speed 𝑛𝑠𝑦𝑛𝑐 , while the rotor itself turns at a slower speed.
108
Figure 3.8: The development of induced torque in an induction motor. (a) The rotating stator
field BS induces a voltage in the rotor bars; (b) the rotor voltage produces a rotor current flow,
which lags behind the voltage because of the inductance of the rotor; (c) the rotor current
produces a rotor magnetic field BR lagging 90° behind itself, and BR interacts with Bnet to produce
a counterclockwise torque in the machine.
Thus, from the working principle of three phase induction motor it may observed that the rotor
speed should not reach the synchronous speed produced by the stator. If the speeds equals, there
would be no such relative speed, so no emf induced in the rotor, & no current would be flowing,
and therefore no torque would be generated. Consequently, the rotor cannot reach the synchronous
speed. The difference between the stator (synchronous speed) and rotor speeds is called the slip.
The rotation of the magnetic field in an induction motor has the advantage that no electrical
connections need to be made to the rotor.
When the induction motor starts the rotor will runs in the same direction of the rotating magnetic
field at a speed less than the synchronous speed. The rotor never attains the synchronous speed
but always runs at a speed less than Ns. The relative speed between the synchronously rotating
field and the actual speed of the rotor is called as Slip Speed.
The voltage induced in a rotor bar of an induction motor depends on the speed of the rotor relative
to the magnetic fields. Since the behavior of an induction motor depends on the rotor's voltage and
current, it is often more logical to talk about this re lative speed. Two terms are commonly used to
de fine the relative motion of the rotor and the magnetic fields. One is slip speed, defined as the
difference between synchronous speed and rotor speed:
109
where 𝑛𝑠𝑙𝑖𝑝 = slip speed of the machine
The other term used to describe the relative motion is slip, which is the relative speed expressed
on a per-unit or a percentage basis. That is, slip is defined as:
𝑛
𝑠 = 𝑛 𝑠𝑙𝑖𝑝 (× 100%) (3.5)
𝑠𝑦𝑛𝑐
𝑛𝑠𝑦𝑛𝑐 −𝑛𝑚
𝑠= (× 100%) (3.6)
𝑛𝑠𝑦𝑛𝑐
This equation can also be expressed in terms of angular velocity w (radians per second) as:
𝜔𝑠𝑦𝑛𝑐 −𝜔𝑚
𝑠= (× 100%) (4.7)
𝜔𝑠𝑦𝑛𝑐
At normal load the % slip of induction motor varies between 2 to 5 % and at no load it is as
small as 0.5 %.
Notice that if the rotor turns at synchronous speed, s = 0, while if the rotor is stationary, s = 1. All
normal motor speeds fall somewhere between those two limits. It is possible to ex press the
mechanical speed of the rotor shaft in terms of synchronous speed and slip. Solving Equations
(3.6) and (3.7) for mechanical speed yields:
𝑛𝑚 = 𝑛𝑠𝑦𝑛𝑐 (1 − 𝑠) (3.8)
Or 𝜔𝑚 = 𝜔𝑠𝑦𝑛𝑐 (1 − 𝑠) (3.9)
These equations are useful in the derivation of induction motor torque and power relationships.
An induction motor works by inducing voltages and currents in the rotor of the machine, and for
that reason it has sometimes been called a rotating transformer. Like a transformer, the primary
(stator) induces a voltage in the secondary (rotor), but unlike a transformer, the secondary
frequency is not necessarily the same as the primary frequency.
110
If the rotor of a motor is locked so that it cannot move, then the rotor will have the same frequency
as the stator. On the other hand, if the rotor turns at synchronous speed, the frequency on the rotor
will be zero. What will the rotor frequency be for any arbitrary rate of rotor rotation?
At nm = 0 r/min, the rotor frequency fr = fe, and the slip s = 1. At nm = 𝑛𝑠𝑦𝑛𝑐 , the rotor frequency
fr = 0 Hz, and the slip s = 0. For any speed in between, the rotor frequency is directly proportional
to the difference between the speed of the magnetic field 𝑛𝑠𝑦𝑛𝑐 and the speed of the rotor 𝑛𝑚 .
Since the slip of the rotor is defined as:
𝑛𝑠𝑦𝑛𝑐 −𝑛𝑚
𝑠= (3.10)
𝑛𝑠𝑦𝑛𝑐
𝑓𝑟 = 𝑠𝑓𝑒 (3.11)
Several alternative forms of this expression exist that are sometimes useful. One of the more
common expressions is derived by substituting Equation (3.10) for the slip into Equation (3.11)
and then substituting for 𝑛𝑠𝑦𝑛𝑐 in the denominator of the expression:
𝑛𝑠𝑦𝑛𝑐 −𝑛𝑚
𝑓𝑟 = 𝑓𝑒
𝑛𝑠𝑦𝑛𝑐
120𝑓𝑒
But 𝑛𝑠𝑦𝑛𝑐 = [from Equation (3.1 )], so
𝑃
𝑝
𝑓𝑟 = (𝑛𝑠𝑦𝑛𝑐 − 𝑛𝑚 ) 𝑓
120𝑓𝑒 𝑒
Therefore,
𝑝
𝑓𝑟 = 120 (𝑛𝑠𝑦𝑛𝑐 − 𝑛𝑚 ) (3.12)
Example 3.1. Find the % slip of an induction motor with 4 poles and 50 Hz supply when it is
running at 1475 rpm.
120𝑓𝑒 𝑛𝑠𝑦𝑛𝑐 −𝑛𝑚 ((1500 – 1475)
Soln. 𝑛𝑠𝑦𝑛𝑐 = = 120 x 50 / 4 = 1500 rpm, %𝑠𝑙𝑖𝑝 = = = 1.6%
𝑃 𝑛𝑠𝑦𝑛𝑐 1500
Example 3.2. A slip-ring induction motor runs at 290 r.p.m. at full load, when connected to
50-Hz supply. Determine the number of poles and slip.
111
Soln. Since N is 290 rpm; Ns has to be somewhere near it, say 300 rpm. If Ns is assumed as
300 rpm, then 300 = 120 × 50/P. Hence, P = 20. ∴ s = (300 − 290)/300 = 3.33%
Example 3.3. The stator of a 3-ϕ induction motor has 3 slots per pole per phase. If supply
frequency is 50 Hz, calculate (i) number of stator poles produced and total number of slots on the
stator (ii) speed of the rotating stator flux (or magnetic field).
Example 3.4. A 4-pole, 3-phase induction motor operates from a supply whose frequency is 50
Hz. Calculate: (i) the speed at which the magnetic field of the stator is rotating. (ii) the speed of
the rotor when the slip is 0.04. (iii) the frequency of the rotor currents when the slip is 0.03. (iv)
the frequency of the rotor currents at standstill.
Soln. (i) Stator field revolves at synchronous speed, given by Ns = 120 f/P = 120 × 50/4 = 1500
r.p.m. (ii) rotor (or motor) speed, Nm = Ns (1 − s) = 1500(1 − 0.04) = 1440 r.p.m. (iii) frequency
of rotor current, fr = sfe = 0.03 × 50 = 1.5 r.p.s = 90 r.p.m (iv) Since at standstill, s = 1, fr = sfe= 1
× f = f = 50Hz
Example 3.5. A 3-ϕ induction motor is wound for 4 poles and is supplied from 50-Hz system.
Calculate (i) the synchronous speed (ii) the rotor speed, when slip is 4% and (iii) rotor frequency
when rotor runs at 600 rpm.
Soln. (i) Ns = 120 f/P = 120 × 50/4 = 1500 rpm (ii) rotor speed, N = Ns (1 − s) = 1500 (1 − 0.04)
= 1440 rpm
(iii) when rotor speed is 600 rpm, slip is s = (Ns − N)/Ns = (1500 − 600)/1500 = 0.6 rotor current
frequency, fr = sfe = 0.6 × 50 = 30 Hz
Example 3.6. A 12-pole, 3-phase alternator driven at a speed of 500 r.p.m. supplies power to an
8-pole, 3-phase induction motor. If the slip of the motor, at full-load is 3%, calculate the full-load
speed of the motor.
Soln. Let N = actual motor speed; Supply frequency, f = 12 × 500/120 = 50 Hz. Synchronous
𝑛𝑠𝑦𝑛𝑐 −𝑛𝑚 750−𝑛𝑚
speed Ns = 120 × 50/8 = 750 r.p.m., % slip 𝑠 = (× 100%); 3 = (× 100%);
𝑛𝑠𝑦𝑛𝑐 750
112
∴ Nm = 727.5 r.p.m. Note. Since slip is 3%, actual speed Nm is less than Ns by 3% of Ns i.e. by 3
× 750/100 = 22.5 r.p.m.
Problem 1. Find the % slip of an induction motor with 4 poles and 60 Hz supply when it is
running at 1400 rpm.
Problem 2. Find the speed of an induction motor with 4 poles and 60 Hz supply when it is
running at a % slip of 4.
Problem 3. Find the % slip of an induction motor with 4 poles and 50 Hz supply when it is
running at 1350 rpm and also when poles is reduced to 2.
Expected questions
It has been shown that in the case of a d.c. motor, the torque Ta is proportional to the product of
armature current and flux per pole i.e. Ta ∝ ϕIa. Similarly, in the case of an induction motor, the
torque is also proportional to the product of flux per stator pole and the rotor current. However,
there is one more factor that has to be taken into account i.e. the power factor of the rotor.
∴ T ∝ E2 I2 cosϕ2
or T = k1 E2 I2 cosϕ2 (3.14)
113
The effect of rotor power factor on rotor torque is illustrated in Fig. 3.9a and Fig. 3.9b for various
values of ϕ2. From the above expression for torque, it is clear that as ϕ2 increases (and hence, cos
ϕ2 decreases) the torque decreases and vice versa. In the discussion to follow, the stator flux
distribution is assumed sinusoidal. This revolving flux induces in each rotor conductor or bar an
e.m.f. whose value depends on the flux density, in which the conductor is lying at the instant
considered (𝑒𝑖𝑛𝑑 = (𝑣 × 𝐵) ∙ 𝑙 volt). Hence, the induced e.m.f. in the rotor is also sinusoidal.
(a) (b)
Figure 3.9: The effect of rotor power factor on rotor torque
It is seen that for a portion ‘ab’ of the pole pitch, the torque is negative i.e. reversed. Hence, the
total torque which is the difference of the forward and the backward torques, is considerably
reduced. If ϕ2= 90°, then the total torque is zero because in that case the backward and the forward
torques become equal and opposite.
114
3.5.1 Starting Torque
The torque developed by the motor at the instant of starting is called starting torque. In some cases,
it is greater than the normal running torque, whereas in some other cases it is somewhat less.
R2 = rotor resistance/phase
𝐸2 𝑅2 𝐸2 2 𝑅2
Or 𝑇𝑠𝑡 = 𝑘1 𝐸2 . × = 𝑘1 𝑅 2 +𝑋 2 (3.16)
√𝑅2 2 +𝑋2 2 √𝑅2 2 +𝑋2 2 2 2
If supply voltage V is constant, then the flux Φ and hence, E2 both are constant.
𝑅2 𝑅
∴ 𝑇𝑠𝑡 = 𝑘2 𝑅 2 +𝑋 2 = 𝑘2 𝑍 22 where 𝐾2 is some other constant.
2 2 2
3 3 𝐸2 2 𝑅2
Now , 𝑘1 = 2𝜋𝑁 , ∴ 𝑇𝑠𝑡 = 2𝜋𝑁 . 𝑅 2 +𝑋 2 (3.17)
𝑠 𝑠 2 2
The resistance of a squirrel-cage motor is fixed and small as compared to its reactance which is
very large especially at the start because at standstill, the frequency of the rotor currents equals the
supply frequency. Hence, the starting current I2 of the rotor, though very large in magnitude, lags
by a very large angle behind E2 , with the result that the starting torque per ampere is very poor. It
is roughly 1.5 times the full-load torque, although the starting current is 5 to 7 times the full-load
current. Hence, such motors are not useful where the motor has to start against heavy loads.
The starting torque of such a motor is increased by improving its power factor by adding external
resistance in the rotor circuit from the star-connected rheostat, the rheostat resistance being
progressively cut out as the motor gathers speed. Addition of external resistance, however,
115
increases the rotor impedance and so reduces the rotor current. At first, the effect of improved
power factor predominates the current-decreasing effect of impedance. Hence, starting torque is
increased. But after a certain point, the effect of increased impedance predominates the effect of
improved power factor and so the torque starts decreasing.
It can be proved that starting torque is maximum when rotor resistance equals rotor reactance.
𝑘2 𝑅2 𝑑𝑇𝑠𝑡 1 𝑅2 (2𝑅2 )
Now, 𝑇𝑠𝑡 = 𝑅 2 2 ∴ = 𝑘2 [𝑅 2 2 − (𝑅 2 +𝑋 2 )2 ]=0
2 +𝑋2 𝑑𝑅2 2 +𝑋2 2 2
Or 𝑅2 2 + 𝑋2 2 = 2𝑅2 2 ∴ 𝑅2 = 𝑋2 . (3.18)
𝑉 2 𝑅2 𝑉 2 𝑅2
∴ 𝑇𝑠𝑡 = 𝑘3 𝑅 2 2 = 𝑘3 where 𝑘3 is yet another constant. Hence 𝑇𝑠𝑡 ∝ 𝑉 2 .
2 +𝑋2 𝑍2 2
Clearly, the torque is very sensitive to any changes in the supply voltage. A change of 5 percent in
supply voltage, for example, will produce a change of approximately 10% in the rotor torque. This
fact is of importance in star-delta and auto transformer starters.
When rotor is stationary i.e. s = 1, the frequency of rotor e.m.f. is the same as that of the stator
supply frequency. The value of e.m.f. induced in the rotor at standstill is maximum because the
relative speed between the rotor and the revolving stator flux is maximum. In fact, the motor is
equivalent to a 3-phase transformer with a short-circuited rotating secondary.
When rotor starts running, the relative speed between it and the rotating stator flux is decreased.
Hence, the rotor induced e.m.f. which is directly proportional to this relative speed, is also
decreased (and may disappear altogether if rotor speed were to become equal to the speed of stator
flux). Hence, for a slip s, the rotor induced e.m.f. will be s times the induced e.m.f. at standstill.
116
The frequency of the induced e.m.f. will likewise become fr = sf2
Due to decrease in frequency of the rotor e.m.f., the rotor reactance will also decrease.
∴ 𝑋𝑟 = 𝑠𝑋2 (3.20)
where Er and Xr are rotor e.m.f. and reactance under running conditions.
Now Er = sE2
𝐸 𝑠𝐸2
∴ 𝐼𝑟 = 𝑍𝑟 = (3.21)
𝑟 √[𝑅2 2 +(𝑠𝑋2 )2 ]
𝑅2
From Fig.3.11, 𝑐𝑜𝑠𝜙2 =
√[𝑅2 2 +(𝑠𝑋2 )2 ]
𝑠𝜙𝐸2 𝑅2 𝑠𝜙𝐸2 𝑅2
∴ 𝑇 ∝ (𝑅 2 2
= 𝑘 (𝑅 2 +(𝑠𝑋 )2 ) Where, (∴ 𝐸2 ∝ 𝜙)
2 +(𝑠𝑋2 ) ) 2 2
𝑘 .𝑠𝐸 2 𝑅2
Also , 𝑇 = (𝑅 12+(𝑠𝑋
2
2
(3.22)
2 2) )
3 𝑠𝐸2 2 𝑅2 3 𝑠𝐸2 2 𝑅2
𝑇 = 2𝜋𝑁 . (𝑅 2 +(𝑠𝑋 = 2𝜋𝑁 . (3.23)
𝑠 2 2 )2 ) 𝑠 𝑍𝑟 2
𝑘 𝐸 2𝑅 3 𝐸2 2 𝑅2
𝑇 = 𝑅12+𝑋
2 2
2 (𝑜𝑟 = 2𝜋𝑁 . 𝑅 2 +𝑋 2 ) the same as Eqn (3.17).
2 2 𝑠 2 2
117
Figure 3.11 rotor impedance
𝑘𝑠𝜙𝐸2 𝑅2 𝑠𝐸2 2 𝑅2
𝑇 = (𝑅 2 = 𝑘1 (𝑅 2 (3.24)
2 +(𝑠𝑋2 )2 ) 2 +(𝑠𝑋2 )2 )
The condition for maximum torque may be obtained by differentiating the above expression with
1
respect to slip s and then putting it equal to zero. However, it is simpler to put 𝑌 = 𝑇 and then
differentiate it.
𝑑𝑌 −𝑅2 𝑋2 2
= + =0
𝑑𝑠 𝑘𝑠 2 𝜙𝐸2 𝑘𝜙𝐸2 𝑅2
𝑅2 𝑋 2
∴ = 𝑘𝜙𝐸2 𝑅 or 𝑅2 2 = 𝑠 2 𝑋2 2 or 𝑅2 = 𝑠𝑋2 (3.25)
𝑘𝑠2 𝜙𝐸2 2 2
Hence, torque under running condition is maximum at that value of the slip s which makes rotor
reactance per phase equal to rotor resistance per phase. This slip is sometimes written as sb and
the maximum torque as Tb.
Putting R2 = sX2 in the above equation (3.24) for the torque, we get
Substituting value of s = R2 /X2 in the other equation given in (3.24) above, we get
(R 2 /X2 ). 𝐸2 2 . 𝑅2 𝐸2 2
𝑇𝑚𝑎𝑥 = 𝑘1 = 𝑘1 .
(𝑅2 2 + ((R 2 /X2 ). 𝑋2 )2 ) 2𝑋2
3 3 𝐸 2
Since , 𝑘1 = 2𝜋𝑁 , we have 𝑇𝑚𝑎𝑥 = 2𝜋𝑁 . 2𝑋2 𝑁. 𝑚 (3.27)
𝑠 𝑠 2
118
resistance. As seen from above, torque becomes maximum when rotor reactance equals its
resistance. Hence, by varying rotor resistance (possible only with slip-ring motors) maximum
torque can be made to occur at any desired slip (or motor speed).
3 maximum torque varies inversely as standstill reactance. Hence, it should be kept as small as
possible.
4 maximum torque varies directly as the square of the applied voltage.
5 for obtaining maximum torque at starting (s =1), rotor resistance must be equal to rotor
reactance.
The rotor torque at any slip s can be expressed in terms of the maximum (or breakdown) torque
Tb by the following equation
2
𝑇 = 𝑇𝑏 [ 𝑠𝑏⁄ 𝑠 ] (3.28)
( 𝑠)+( ⁄𝑠𝑏 )
A family of torque/slip curves is shown in Fig. 3.12 for a range of s = 0 to s = 1 with R2 as the
parameter. We have seen above in the Eqn, (3.22) that
𝑘𝜙𝑠𝐸2 𝑅2
𝑇 = (𝑅 2 +(𝑠𝑋 )2 )
2 2
At normal speeds, close to synchronism, the term (s X2) is small and hence negligible w.r.t. R2.
𝑠
∴ 𝑇∝𝑅 (3.29)
2
Or 𝑇∝𝑠 if R2 is constant.
Hence, for low values of slip, the torque/slip curve is approximately a straight line. As slip
increases (for increasing load on the motor), the torque also increases and becomes maximum
when s = R2 /X2. This torque is known as ‘pull-out’ or ‘breakdown’ torque Tb or stalling torque.
As the slip further increases (i.e. motor speed falls) with further increase in motor load, then R2
becomes negligible as compared to (sX2.). Therefore, for large values of slip
119
𝑠 1
𝑇 ∝ (𝑠𝑋 ∝𝑠 (3.30)
2 )2
Hence, the torque/slip curve is a rectangular hyperbola. So, we see that beyond the point of
maximum torque, any further increase in motor load results in decrease of torque developed by the
motor. The result is that the motor slows down and eventually stops. The circuit-breakers will be
tripped open if the circuit has been so protected. In fact, the stable operation of the motor lies
between the values of s = 0 and that corresponding to maximum torque. The operating range is
shown shaded in Fig. 3.12.
It is seen that although maximum torque does not depend on R2 , yet the exact location of Tmax is
dependent on it. Greater the R2 , greater is the value of slip at which the maximum torque occurs.
Obviously, torque at any speed is proportional to the square of the applied voltage. If stator voltage
decreases by 10%, the torque decreases by 20%. Changes in supply voltage not only affect the
starting torque Tst but torque under running conditions also. If V decreases, then T also decreases.
Hence, for maintaining the same torque, slip increases i.e. speed falls.
2
Let V change to V′, s to s′ and T to T′; then 𝑇⁄𝑇 ′ = 𝑠𝑉 ⁄ ′ ′ 2 (3.31)
𝑠𝑉
120
Hardly any important changes in frequency take place on a large distribution system except during
a major disturbance. However, large frequency changes often take place on isolated, low power
systems in which electric energy is generated by means of diesel engines or gas turbines. Examples
of such systems are: emergency supply in a hospital and the electrical system on a ship etc.
The major effect of change in supply frequency is on motor speed. If frequency drops by 10%,
then motor speed also drops by 10%. Machine tools and other motor-driven equipment meant for
50 Hz causes problem when connected to 60-Hz supply. Everything runs (60 − 50) × 100/50 =
20% faster than normal and this may not be acceptable in all applications. In that case, we have to
use either gears to reduce motor speed or an expensive 50-Hz source.
A 50-Hz motor operates well on a 60-Hz line provided its terminal voltage is raised to 60/50 = 6/5
(i.e. 120%) of the name-plate rating. In that case, the new breakdown torque becomes equal to the
original breakdown torque and the starting torque is only slightly reduced. However, power factor,
efficiency and temperature rise remain satisfactory.
Similarly, a 60-Hz motor can operate satisfactorily on 50-Hz supply provided its terminal voltage
is reduced to 5/6 (i.e. 80%) of its name-plate rating.
𝑇𝑓 2𝑠𝑓 𝑅2 𝑋2
∴ = (𝑅 2 +(𝑠 𝑋 )2 ) (3.33)
𝑇𝑚𝑎𝑥 2 𝑓 2
In fact a = sm —slip corresponding to maximum torque. In that case, the relation becomes
𝑇𝑓 2𝑠𝑚 𝑠𝑓 2
=𝑠 2 +𝑠 2
(3.35)
𝑇𝑚𝑎𝑥 𝑚 𝑓
121
𝑜𝑝𝑒𝑟𝑎𝑡𝑖𝑛𝑔 𝑡𝑜𝑟𝑞𝑢𝑒 𝑎𝑡 𝑎𝑛𝑦 𝑠𝑙𝑖𝑝 𝑠 2𝑎𝑠
In general, = 𝑎2 +𝑠 2
(3.36)
𝑚𝑎𝑥𝑖𝑚𝑢𝑚 𝑡𝑜𝑟𝑞𝑢𝑒 𝑓
1
𝑇𝑚𝑎𝑥 ∝ 2𝑋
2
𝑇𝑠𝑡 2𝑅2 𝑋2 2𝑅 /𝑋 2𝑎
∴ =𝑅 2 2 = 1+(𝑅2 /𝑋2 )2 = 1+𝑎2 (3.38)
𝑇𝑚𝑎𝑥 2 +𝑋2 2 2
𝑟𝑜𝑡𝑜𝑟 𝑟𝑒𝑠𝑖𝑠𝑡𝑎𝑛𝑐𝑒
Where, 𝑎 = 𝑅2 /𝑋2 = 𝑠𝑡𝑎𝑛𝑑𝑠𝑡𝑖𝑙𝑙 𝑟𝑒𝑎𝑐𝑡𝑎𝑛𝑐𝑒 per phase
The equivalent circuit of any machine shows the various parameter of the machine such as its
Ohmic losses and also other losses. The losses are modeled just by inductor and resistor. The
copper losses are occurred in the windings so the winding resistance is taken into account. Also,
the winding has inductance for which there is a voltage drop due to inductive reactance and also a
term called power factor comes into the picture. There are two types of equivalent circuits in case
of a three-phase induction motor-
122
Exact Equivalent Circuit
Here, R1 is the winding resistance of the stator. X1 is the inductance of the stator winding. Rc is
the core loss component. XM is the magnetizing reactance of the winding. R2/s is the power of
the rotor, which includes output mechanical power and copper loss of rotor. If we draw
the circuit with referred to the stator then the circuit will look like-
to stator winding. X2’ is the rotor winding inductance with referred to stator winding. R2(1 - s) / s
is the resistance which shows the power which is converted to mechanical power output or useful
power. The power dissipated in that resistor is the useful power output or shaft power.
The approximate equivalent circuit is drawn just to simplify our calculation by deleting one node.
The shunt branch is shifted towards the primary side. This has been done as the voltage drop
between the stator resistance and inductance is less and there is not much difference between the
supply voltage and the induced voltage. However, this is not appropriate due to following reasons-
1 The magnetic circuit of induction motor has an air gap so exciting current is larger compared to
transformer so exact equivalent circuit should be used.
This model can be used if approximate analysis has to be done for large motors. For smaller
motors, we cannot use this.
123
Actual rotor resistance
Resistance equivalent to
Losses and the Power-Flow Diagram mechanical load
An induction motor can be basically described as a rotating transformer. Its input is a three-phase
system of voltages and currents. For an ordinary transformer, the output is electric power from the
secondary windings. The secondary windings in an induction motor (the rotor) are shorted out, so
no electrical output exists from normal induction motors. Instead, the output is mechanical. The
relationship between the input electric power and the output mechanical power of this motor is
shown in the power-t1ow diagram in Figure 6-13.
The input power to an induction motor An is in the form of three-phase electric voltages and
cunents. The first losses encountered in the machine are [2R losses in the stator windings (the
stator copper loss PSCL ). Then some amount of power is lost as hysteresis and eddy currents in
the stator (Pcore ). The power remaining at this point is transfened to the rotor of the machine
across the air gap between the stator and rotor. This power is called the air-gap power PAG of the
( machine. After the power is transferred to the rotor, some of it is lost as [ 2R losses (the rotor
copper loss PRCL), and the rest is converted from electrical to mechanical form (PconJ. Finally,
friction and windage losses PF&W and stray losses Pmisc are subtracted. The remaining power is
the output of the motor Pout
124
Figure 3- 18 shows the per-phase equivalent circuit of an induction motor. If the equivalent circuit
is examined closely, it can be used to derive the power and torque equations governing the
operation of the motor.
The input current to a phase of the motor can be found by dividing the input voltage by the total
equivalent impedance:
Therefore, the stator copper losses, the core losses, and the rotor copper losses can be found. The
stator copper losses in the three phases are given by
Therefore, the stator copper losses, the core losses, and the rotor copper losses can be found. The
stator copper losses in the three phases are given by
𝑃𝑆𝐶𝐿 = 3 𝐼12 𝑅1
𝑃𝑅𝐶𝐿 = 3𝐼22 𝑅2
𝑃𝑐𝑜𝑛𝑣
𝜏𝑖𝑛𝑑 =
𝜔𝑚
Induction motors do not present the types of starting problems that synchronous motors do. In
many cases, induction motors can be staI1ed by simply connecting them to the power line.
However, there are sometimes good reasons for not doing this. For example, the starting current
required may cause such a dip in the power system voltage that across-the-line starting is not
acceptable.
125
For wound-rotor induction motors, starting can be achieved at relatively low currents by inserting
extra resistance in the rotor circuit during starting. This extra resistance not only increases the
starting torque but also reduces the starting current.
For cage induction motors, the starting current can vary widely depending primarily on the motor's
rated power and on the effective rotor resistance at starting conditions. To estimate the rotor current
at starting conditions, all cage motors now have a starting code letter (not to be confused with their
design class letter) on their nameplates. The code letter sets limits on the amount of current the
motor can draw at starting conditions.
The equivalent-circuit parameters needed for computing the performance of a polyphase induction
motor under load can be obtained from the results of a no-load test, a blocked-rotor test, and
measurements of the dc resistances of the stator windings. Stray-load losses, which must be taken
into account when accurate values of efficiency are to be calculated, can also be measured by tests
which do not require loading the motor. The stray-load-loss tests are not described here, however.
No-Load Test
Like the open-circuit test on a transformer, the no-load test on an induction motor gives
information with respect to exciting current and no-load losses. This test is ordinarily performed
at rated frequency and with balanced polyphase voltages applied to the stator terminals. Readings
are taken at rated voltage, after the motor has been running long enough for the bearings to be
properly lubricated. We will assume that the no-load test is made with the motor operating at its
rated electrical frequency fr and that the following measurements are available from the no-load
test:
In polyphase machines it is most common to measure line-to-line voltage, and thus the phase-to-
neutral voltage must be then calculated (dividing by ~/3 in the case of a three-phase machine).
126
At no load, the rotor current is only the very small value needed to produce sufficient torque to
overcome the friction and windage losses associated with rotation. The no-load rotor IZR loss is,
therefore, negligibly small. Unlike the continuous magnetic core in a transformer, the magnetizing
path in an induction motor includes an air gap which significantly increases the required exciting
current. Thus, in contrast to the case of a transformer, whose no-load primary 12 R loss is
negligible, the no-load stator I 2 R loss of an induction motor may be appreciable because of this
larger exciting current.
Blocked.Rotor Test
Like the short-circuit test on a transformer, the blocked-rotor test on an induction motor gives
information with respect to the leakage impedances. The rotor is blocked so that it cannot rotate
(hence the slip is equal to unity), and balanced polyphase voltages are applied to the stator
terminals. We will assume that the following measurements are available from the blocked-rotor
test:
Vl,bl = The line-to-neutral voltage [V] II,bl -- The line current [V] Pbl -- The total polyphase
electrical input power [W] fbl = The frequency of the blocked-rotor test [Hz] In some cases, the
blocked-rotor torque also is measured. The equivalent circuit for blocked-rotor conditions is
identical to that of a shortcircuited transformer. An induction motor is more complicated than a
transformer, however, because its leakage impedance may be affected by magnetic saturation of
the leakage-flux paths and by rotor frequency. The blocked-rotor impedance may also be affected
by rotor position, although this effect generally is small with squirrel-cage rotors. The guiding
principle is that the blocked-rotor test should be performed under conditions for which the current
and rotor frequency are approximately the same as those in the machine at the operating condition
for which the performance is later to be calculated. For example, if one is interested in the
characteristics at slips near unity, as in starting, the blocked-rotor test should be taken at normal
frequency and with currents near the values encountered in starting. If, however, one is interested
in normal running characteristics, the blocked-rotor test should be taken at a reduced voltage which
results in approximately rated current; the frequency also should be reduced, since the values of
rotor effective resistance and leakage inductance at the low rotor frequencies corresponding to
small slips may differ appreciably from their values at normal frequency, particularly with double-
cage or deep-bar rotors, as discussed in Section 6.7.2.
127
OBJECTIVE TESTS – 3
1. Regarding skewing of motor bars in a squirrel- cage induction motor, (SCIM) which statement is
false?
(a) it prevents cogging (d) it reduces motor ‘hum’ during its
(b) it increases starting torque operation
128
8. In a 3-ϕ induction motor, the rotor field rotates at synchronous speed with respect
to
(a) stator (c) stator flux
10. In the case of a 3-ϕ induction motor having Ns = 1500 rpm and running with
s= 0.04
(a) revolving speed of the stator flux is space is….rpm
(b) rotor speed is….rpm
(c) speed of rotor flux relative to the rotor is….rpm
(d) speed of the rotor flux with respect to the stator is….rpm
11. The number of stator poles produced in the rotating magnetic field of a 3-ϕ
induction motor having 3 slots per pole per phase is
(a) 3 (b) 6 (c) 2 (d) 12
12. The power factor of a squirrel-cage induction motor is
(a) low at light loads only (c) low at light and heavy loads both
(b) low at heavy loads only (d) low at rated load only
13. Which of the following rotor quantity in a SCIM does NOT depend on its slip?
(a) reactance (c) induced emf
(b) speed (d) frequency
14. A 6-pole, 50-Hz, 3-j induction motor is running at 950 rpm and has rotor Cu
loss of 5 kW. Its rotor input is… . kW.
(a) 100 (b) 10 (c) 95 (d) 5.3
15. The efficiency of a 3-phase induction motor is approximately proportional to
(a) (1- s) (b) s (c) N (d) Ns
129
16. A 6-pole, 50-Hz, 3-ϕ induction motor has a full- load speed of 950 rpm. At half-
load, its speed would be… . rpm.
(a) 475 (b) 500 (c) 975 (d) 1000
17. If rotor input of a SCIM running with a slip of 10% is 100 kW, gross power
developed by its rotor is… . kW.
(a) 10 (b) 90 (c) 99 (d) 80
18. Pull-out torque of a SCIM occurs at that value of slip where rotor power factor
equals
(a) unity (b) 0.707 (c) 0.866 (d) 0.5
19. Fill in the blanks.
When load is placed on a 3-phase induction motor, its
130
three factors:
(a) speed, frequency, number of poles frequency
(b) voltage, current and stator impedance (d) rotor emf, rotor current and rotor p.f
131
Chapter Four
DC machines
Contents of this chapter includes:
4.1 Background
DC machines are generators that convert mechanical energy to dc electric energy and motors that
convert dc electric energy to mechanical energy. Most dc machines are like ac machines in that
they have ac voltages and currents within them. DC machines have a dc output only because a
mechanism exists that converts the internal ac voltages to dc voltages at their terminals. Since this
mechanism is called a commutator, dc machinery is also known as commutating machinery.cham
The fundamental principles involved in the operation of dc machines are very simple.
Unfortunately, they are usually somewhat obscured by the complicated construction of real
machines.
4.2 Construction
Working principles
Imagine the coil to be rotating in clock-wise direction [Fig 4.2 (a)]. As the coil assumes successive
positions in the field, the flux linked with it changes. Hence, an e.m.f. is induced in it which is
𝑁𝑑𝜙
proportional to the rate of change of flux linkages (𝑒 = ). When the plane of the coil is at right
𝑑𝑡
angles to lines of flux i.e. when it is in position, 1, then flux linked with the coil is maximum but
rate of change of flux linkages is minimum.
It is so because in this position, the coil sides AB and CD do not cut or shear the flux, rather they
slide along them i.e. they move parallel to them. Hence, there is no induced e.m.f. in the coil. Let
us take this no-e.m.f. or vertical position of the coil as the starting position. The angle of rotation
or time will be measured from this position.
(a) (b)
Figure:4.2 (a) a rotating coil in magnetic field (b) induced emf output
As the coil continues rotating further, the rate of change of flux linkages (and hence induced e.m.f.
in it) increases, till position 3 is reached where θ = 90º. Here, the coil plane is horizontal i.e. parallel
to the lines of flux. As seen, the flux linked with the coil is minimum but rate of change of flux
linkages is maximum. Hence, maximum e.m.f. is induced in the coil when in this position [Fig.
4.2(b)].
133
In the next quarter revolution i.e. from 90º to 180º, the flux linked with the coil gradually increases
but the rate of change of flux linkages decreases. Hence, the induced e.m.f. decreases gradually
till in position 5 of the coil, it is reduced to zero value.
So, we find that in the first half revolution of the coil, no (or minimum) e.m.f. is induced in it when
in position 1, maximum when in position 3 and no e.m.f. when in position 5. The direction of this
induced e.m.f. can be found by applying Fleming’s Right-hand rule which gives its direction from
A to B and C to D. Hence, the direction of current flow is ABMLCD (Fig. 4.1). The current through
the load resistance R flows from M to L during the first half revolution of the coil.
In the next half revolution i.e. from 180º to 360º, the variations in the magnitude of e.m.f. are
similar to those in the first half revolution. Its value is maximum when coil is in position 7 and
minimum when in position 1. But it will be found that the direction of the induced current is from
D to C and B to A as shown in [Fig. 4.1 (b)]. Hence, the path of current flow is along DCLMBA
which is just the reverse of the previous direction of flow.
Therefore, we find that the current which we obtain from such a simple generator reverses its
direction after every half revolution. Such a current undergoing periodic reversals is known as
alternating current. It is, obviously, different from a direct current which continuously flows in one
and the same direction. It should be noted that alternating current not only reverses its direction, it
does not even keep its magnitude constant while flowing in any one direction. The two half-cycles
may be called positive and negative half-cycles respectively [Fig. 4.2(b)].
For making the flow of current unidirectional in the external circuit, the slip-rings are replaced by
split-rings [Fig. 4.3(a)]. The split-rings are made out of a conducting cylinder which is cut into two
halves or segments insulated from each other by a thin sheet of mica or some other insulating mate-
rial [Fig. 4.3(b)].
As before, the coil ends are joined to these segments on which rest the carbon or copper brushes.
It is seen [Fig. 4.4 (a)] that in the first half revolution current flows along (ABMNLCD) i.e. the
brush No. 1 in contact with segment ‘a’ acts as the positive end of the supply and ‘b’ as the negative
end. In the next half revolution [Fig. 4.4 (b)], the direction of the induced current in the coil has
reversed. But at the same time, the positions of segments ‘a’ and ‘b’ have also reversed with the
134
(b) rectangular loop with slip-ring (b) slip-ring structure
Figure:4.3 Rotating loop with slip rings
result that brush No. 1 comes in touch with the segment which is positive i.e. segment ‘b’ in this
case. Hence, current in the load resistance again flows from M to L. The waveform of the current
through the external circuit is as shown in Fig. 4.4. This current is unidirectional but not
continuous like pure direct current.
It should be noted that the position of brushes is so arranged that the changeover of segments ‘a’
and ‘b’ from one brush to the other takes place when the plane of the rotating coil is at right angles
to the plane of the lines of flux. It is so because in that position, the induced e.m.f. in the coil is
zero.
Another important point worth remembering is that even now the current induced in the coil sides
is alternating as before. It is only due to the rectifying action of the split-rings (also called
commutator) that it becomes unidirectional in the external circuit. Hence, it should be clearly
understood that even in the armature of a dc generator, the induced voltage is alternating.
135
4.3 Practical Generator
The simple loop generator has been considered in detail merely to bring out the basic principle
underlying construction and working of an actual generator illustrated in Fig. 4.5 which consists
of the following essential parts:
i. It provides mechanical support for the poles and acts as a protecting cover for the whole machine
and
ii. It carries the magnetic flux produced by the poles. In small generators where cheapness rather
than weight is the main consideration, yokes are made of cast iron. But for large machines usually
cast steel or rolled steel is employed. The modern process of forming the yoke consists of rolling
a steel slab round a cylindrical mandrel and then welding it at the bottom. The feet and the
136
terminal box etc. are welded to the frame afterwards. Such yokes possess sufficient mechanical
strength and have high permeability.
The field magnets consist of pole cores and pole shoes. The pole shoes serve two purposes:
i. They spread out the flux in the air gap and also, being of larger cross-section, reduce the
reluctance of the magnetic path
ii. They support the exciting coils (or field coils
a) The pole core itself may be a solid piece made out of either cast iron or cast steel but the
pole shoe is laminated and is fastened to the pole face by means of counter sunk screws as
shown in [Fig. 4.6(b)].
b) In modern design, the complete pole cores and pole shoes are built of thin laminations of
annealed steel which are rivetted together under hydraulic pressure [Fig. 4.6(c)]. The
thickness of laminations varies from 1 mm to 0.25 mm. The laminated poles may be
secured to the yoke in any of the following two ways:
i. Either the pole is secured to the yoke by means of screws bolted through the yoke and
into the pole body or
ii. The holding screws are bolted into a steel bar which passes through the pole across the
plane of laminations [Fig. 4.6(d)].
(c) complete pole shoes (d) pole shoes with steel rod
Figure 4.6: components of dc generator structure
137
Pole Coils
The field coils or pole coils, which consist of copper wire or strip, are former-wound for the correct
dimension. Then, the former is removed and wound coil is put into place over the core.
When current is passed through these coils, they electro-magnetize the poles which produce the
necessary flux that is cut by revolving armature conductors.
Armature Core
It houses the armature conductors or coils and causes them to rotate and hence cut the magnetic
flux of the field magnets. In addition to this, its most important function is to provide a path of
very low reluctance to the flux through the armature from a N-pole to a S-pole.
It is cylindrical or drum-shaped and is built up of usually circular sheet steel discs or laminations
approximately 0.5 mm thick [Fig. 4.7(a)]. It is keyed to the shaft.
The slots are either die-cut or punched on the outer periphery of the disc and the keyway is located
on the inner diameter as shown. In small machines, the armature stampings are keyed directly to
the shaft. Usually, these laminations are perforated for air ducts which permits axial flow of air
through the armature for cooling purposes. Such ventilating channels are clearly visible in the
laminations shown in [Fig. 4.7(b)]
Up to armature diameters of about one meter, the circular stampings are cut out in one piece as
shown in [Fig. 4.7(b)]. But above this size, these circles, especially of such thin sections, are
difficult to handle because they tend to distort and become wavy when assembled together. Hence,
the circular laminations, instead of being cut out in one piece, are cut in a number of suitable
sections or segments which form part of a complete ring.
138
A complete circular lamination is made up of four or six or even eight segmental laminations.
Usually, two keyways are notched in each segment and are dove-tailed or wedge-shaped to make
the laminations self-locking in position.
The purpose of using laminations is to reduce the loss due to eddy currents. Thinner the
laminations, greater is the resistance offered to the induced e.m.f., smaller the current and hence
lesser the I2R loss in the core.
Armature Windings
The armature windings are usually former-wound. These are first wound in the form of flat
rectangular coils and are then pulled into their proper shape in a coil puller. Various conductors of
the coils are insulated from each other. The conductors are placed in the armature slots which are
lined with tough insulating material. This slot insulation is folded over above the armature
conductors placed in the slot and is secured in place by special hard wooden or fiber wedges.
Commutator
The function of the commutator is to facilitate collection of current from the armature conductors.
It rectified i.e. converts the alternating current induced in the armature conductors into
unidirectional current in the external load circuit. It is of cylindrical structure and is built up of
wedge-shaped segments of high-conductivity hard-drawn or drop forged copper. These segments
are insulated from each other by thin layers of mica. The number of segments is equal to the
number of armature coils. Each commutator segment is connected to the armature conductor by
means of a copper lug or strip (or riser). To prevent them from flying out under the action of
centrifugal forces, the segments have V-grooves, these grooves being insulated by conical micanite
rings. A sectional view of commutator is shown in Fig. 4.8
139
Brushes and Bearings
The brushes whose function is to collect current from commutator, are usually made of carbon or
graphite and are in the shape of a rectangular block. These brushes are housed in brush-holders
usually of the box-type variety. As shown in Fig. 4.9, the brush-holder is mounted on a spindle
and the brushes can slide in the rectangular box open at both ends. The brushes are made to bear
down on the commutator by a spring whose tension can be adjusted by changing the position of
lever in the notches. A flexible copper pigtail mounted at the top of the brush conveys current from
the brushes to the holder. The number of brushes per spindle depends on the magnitude of the
current to be collected from the commutator.
Because of their reliability, ball-bearings are frequently employed, though for heavy duties, roller
bearings are preferable. The ball and rollers are generally packed in hard oil for quieter operation
and for reduced bearing wear, sleeve bearings are used which are lubricated by ring oilers fed from
oil reservoir in the bearing bracket.
In real dc machines, there are several ways in which the loops on the rotor (also called the armature)
can be connected to its commutator segments. These different connections affect the number of
parallel current paths within the rotor, the output voltage of the rotor, and the number and position
of the brushes riding on the commutator segments. We wil1 now examine the construction of the
coils on a real dc rotor and then look at how they are connected to the commutator to produce a dc
voltage.
Regardless of the way in which the windings are connected to the commutator segments, most of
the rotor windings themselves consist of diamond-shaped preformed coils which are inserted into
140
the armature slots as a unit (see Fig.4.10). Each coil consists of a number of turns (loops) of wire,
each turn taped and insulated from the other turns and from the rotor slot. Each side of a turn is
ca11ed a conductor. The number of conductors on a machine's armature is given by
𝑍 = 2𝐶Nc (4.1)
Normally, a coil spans 180 electrical degrees. This means that when one side is under the center
of a given magnetic pole, the other side is under the center of a pole of opposite polarity. The
physical poles may not be located 180 mechanical degrees apart, but the magnetic field has
completely reversed its polarity in traveling from under one pole to the next. The relationship
between the electrical angle and mechanical angle in a given machine is given by
𝑃
𝜃𝑒 = 2 𝜃𝑚 (4.2)
If a coil spans 180 electrical degrees, the voltages in the conductors on either side of the coil will
be exactly the same in magnitude and opposite in direction at all times. Such a coil is called a full-
pitch coil.
141
4.4 Types of Generators
Generators are usually classified according to the way in which their fields are excited. Generators
may be divided into: (a) separately-excited generators and (b) self-excited generators.
a) Separately-excited generators are those whose field magnets are energized from an
independent external source of dc current. It is shown diagrammatically in Fig. 4.11.
b) Self-excited generators are those whose field magnets are energized by the current produced
by the generators themselves. Due to residual magnetism, there is always present some flux
in the poles. When the armature is rotated, some e.m.f. and hence some induced current is
produced which is partly or fully passed through the field coils thereby strengthening the
residual pole flux.
There are three types of self-excited generators named according to the manner in which their field
coils (or windings) are connected to the armature:
i. Shunt wound The field windings are connected across or in parallel with the armature
conductors and have the full voltage of the generator applied across them Fig. 4.11.
ii. Series Wound In this case, the field windings are joined in series with the armature
conductors Fig. 4.11. As they carry full load current, they consist of relatively few turns
of thick wire or strips. Such generators are rarely used except for special purposes i.e. as
boosters etc.
i. Compound Wound It is a combination of a few series and a few shunt windings and can
be either short-shunt or long-shunt as shown in Fig. 4.11. In a compound generator, the
shunt field is stronger than the series field. When series field aids the shunt field,
generator is said to be commutatively-compounded. On the other hand, if series field
opposes the shunt field, the generator is said to be differentially compounded.
142
Brush Contact drop
It is the voltage drop over the brush contact resistance when current passes from commutator
segments to brushes and finally to the external load. Its value depends on the amount of current
and the value of contact resistance. This drop is usually small and includes brushes of both
polarities. However, in practice, the brush contact drop is assumed to have following constant
values for all loads.
Generated e.m.f. Eg = e.m.f. generated in any one of the parallel paths i.e. E.
𝑑𝜙
Average e.m.f. generated/conductor 𝜃𝑒 = volt ( ∵ n = 1)
𝑑𝑡
143
𝜙𝑃𝑁 𝑍 𝜙𝑃𝑁𝑍
∴ E.M.F. generated/path = ×2= volt (4.4)
60 120
𝜙𝑁𝑍 𝑃
In general, generated e.m.f. Eg = × (𝐴) volt
60
For a given dc machine, Z, P and A are constant. Hence, putting Ka = ZP/A, we get
Eg = Ka Φ N volts––where N is in r.p.s.
How can the torque on the rotor of a real machine be determined? The torque on the armature of a
real machine is equal to the number of conductors Z times the torque on each conductor. The
torque in any single conductor under the pole faces was previously shown to be
If there are a current path in the machine, then the total armature current IA is split among the a
current paths, so the current in a single conductor is given by
𝐼𝐴
𝐼𝑐𝑜𝑛𝑑 = (4.8)
𝑎
144
𝑟𝐼𝐴 𝑙𝐵
𝐼𝑐𝑜𝑛𝑑 = (4.9)
𝑎
Since there are Z conductors, the total induced torque in a de machine rotor is
𝑍𝑟𝐼𝐴 𝑙𝐵
𝜏𝑖𝑛𝑑 = (4.10)
𝑎
Finally,
Both the internal generated voltage and the induced torque equations just given are only
approximations, because not all the conductors in the machine are under the pole faces at any given
1
time and also because the surfaces of each pole do not cover an entire 𝑃 of the rotor's surface. To
achieve greater accuracy. the number of conductors under the pole faces could be used instead of
the total number of conductors on the rotor.
DC generators take in mechanical power and produce electric power, while dc motors take in
electric power and produce mechanical power. In either case, not all the power input to the machine
appears in useful form at the other end-there is always some loss associated with the process.
The difference between the input power and the output power of a machine is the losses that occur
inside it. Therefore,
𝑃𝑜𝑢𝑡 −𝑃𝑙𝑜𝑠𝑠
𝜂= x100% (4.16)
𝑃𝑖𝑛
145
4.7.1 The Losses in DC Machines
The losses that occur in dc machines can be divided into five basic categories:
Electrical or Copper Losses: Copper losses are the losses that occur in the armature and field
windings of the machine. The copper losses for the armature and field windings are given by
𝐼𝐴 = armature current
𝐼𝐹 = Field current
𝑅𝐴 = armature resistance
𝑅𝐹 = Field resistance
The resistance used in these calculations is usually the winding resistance at normal operating
temperature.
Brush Losses: the brush drop loss is the power lost across the contact potential at the brushes of
the machine. It is given by the equation
𝐼𝐴 = armature current
146
The reason that the brush losses are calculated in this manner is that the voltage drop across a set
of brushes is approximately constant over a large range of armature currents. Unless otherwise
specified, the brush voltage drop is usually assumed to be about 2V.
Core Losses. The core losses are the hysteresis losses and eddy current losses occurring in the
metal of the motor. These losses are described in Chapter 1. these losses vary as the square of the
flux density (𝐵2) and, for the rotor, as the 1.5th power of the speed of rotation (n1.5).
Mechanical Losses. the mechanical losses in a dc machine are the losses associated with
mechanical effects. There are two basic types of mechanical losses: friction and windage. Friction
losses are losses caused by the friction of the bearings in the machine, while windage losses are
caused by the friction between the moving parts of the machine and the air inside the motor's
casing. These losses vary as the cube of the speed of rotation of the machine.
Stray Losses (or Miscellaneous Losses). Stray losses are losses that cannot be placed in one of
the previous categories. No matter how carefully losses are accounted for, some always escape
inclusion in one of the above categories. All such losses are lumped into stray losses. For most
machines, stray losses are taken by convention to be 1 percent of full load.
One of the most convenient techniques for accounting for power losses in a machine is the power-
flow diagram. A power-flow diagram for a dc generator is shown in [Fig. 4.12(a)]. In this figure,
mechanical power is input into the machine,
and then the stray losses, mechanical losses, and core losses are subtracted. After they have been
subtracted, the remaining power is ideally converted from mechanical to electrical form at the
point labeled 𝑃𝑐𝑜𝑛𝑣 . The mechanical power that is converted is given by
147
𝑃𝑐𝑜𝑛𝑣 = 𝜏𝑖𝑛𝑑 𝜔𝑚 (4.20)
𝑃𝑐𝑜𝑛𝑣 = 𝐸𝐴 𝐼𝐴 (4.21)
However, this is not the power that appears at the machine's terminals. Before the terminals are
reached, the electrical 𝐼 2 𝑅 losses and the brush losses must be subtracted.
In the case of dc motors, this power-now diagram is simply reversed. The power-now diagram for
a motor is shown in [Fig. 4.12(b)].
The earliest power systems in the United States were dc systems, but by the 1890s ac power
systems were clearly winning out over dc systems. Despite this fact, dc motors continued to be a
significant fraction of the machinery purchased each year through the 1960s (that fraction has
declined in the last 40 years). Why were dc motors so common, when dc power systems themselves
were fairly rare?
There were several reasons for the continued popularity of dc motors. One was that dc power
systems are still common in cars, trucks, and aircraft. When a vehicle has a dc power system, it
makes sense to consider using dc motors. Another application for dc motors was a situation in
which wide variations in speed are needed. Before the widespread use of power electronic rectifier-
inverters, dc motors were unexcelled in speed control applications. Even if no dc power source
were available, solid-state rectifier and chopper circuits were used to create the necessary dc
power, and dc motors were used to provide the desired speed control.
(Today, induction motors with solid-state drive packages are the preferred choice over dc motors
for most speed control applications. However, there are still some applications where dc motors
are preferred.)
DC motors are often compared by their speed regulations. The speed regulation (SR) of a motor is
defined by
𝜔𝑛𝑙 −𝜔𝑓𝑙
𝑆𝑅 = x100% (4.22)
𝜔𝑓𝑙
148
𝑛𝑛𝑙 −𝑛𝑓𝑙
𝑆𝑅 = x100% (4.23)
𝑛𝑓𝑙
DC motors are, of course, driven from a dc power supply. Unless otherwise specified, the input
voltage to a dc motor is assumed to be constant, because that assumption simplifies the analysis of
motors and the comparison between different types of motors.
The equivalent circuit of a dc motor is shown in Fig. 4.13. In this figure, the armature circuit is
represented by an ideal voltage source 𝐸𝐴 , and a resistor 𝑅𝐴 . This representation is really the
Thevenin equivalent of the entire rotor structure, including rotor coils, interpoles, and
compensating windings, if present. The brush voltage drop is represented by a small battery 𝑉𝑏𝑟𝑢𝑠ℎ
opposing the direction of current flow in the machine. The field coils, which produce the magnetic
flux in the generator, are represented by inductor 𝐿𝐹 and resistor 𝑅𝐹 . The separate resistor 𝑅𝑎𝑑𝑗
represents an external variable resistor used to control the amount of current in the field circuit.
There are a few variations and simplifications of this basic equivalent circuit. The brush drop
voltage is often only a very tiny fraction of the generated voltage in a machine. Therefore, in cases
where it is not too critical, the brush drop voltage may be left out or approximately included in the
value of 𝑅𝐴 .Also, the internal resistance of the field coils is sometimes lumped together with the
variable resistor, and the total is called 𝑅𝐹 [see Figure 4.13(b)]. A third variation is that some
generators have more than one field coil, all of which will appear on the equivalent circuit.
149
𝐸𝐴 = 𝐾𝜙𝜔 (4.24)
Figure 4.13: (a) The equivalent circuit of a dc motor. (b) A simplified equivalent circuit
eliminating the brush voltage drop and combining 𝑅𝑎𝑑𝑗 with the field resistance.
These two equations, the Kirchhoff's voltage law equation of the armature circuit and the machine's
magnetization curve, are all the tools necessary to analyze the behavior and performance of a dc
motor.
𝐸𝐴 = 𝐾𝜙𝜔
Therefore, 𝑬𝑨 is directly proportional to the nux in the machine and the speed of rotation of the
machine. How is the internal generated voltage related to the field current in the machine?
Figure 4.14: The magnetization curve of a dc machine expressed as a plot of EA versus IF. for a
fixed ω0.
150
The field current in a dc machine produces a field magnetomotive force given by ℱm = NFIF. This
magnetomotive force produces a flux in the machine in accordance with its magnetization curve
(Figure 4.13). Since the field current is directly proportional to the magnetomotive force and since
EA is directly proportional to the flux, it is customary to present the magnetization curve as a plot
of EA versus field current for a given speed ω0.
It is worth noting here that, to get the maximum possible power per pound of weight out of a
machine, most motors and generators are designed to operate near the saturation point on the
magnetization curve (at the knee of the curve). This implies that a fairly large increase in field
current is often necessary to get a small increase in EA when operation is near full load.
(a)
(b)
Figure 4.15:(a) The equivalent circuit of a separately excited dc motor. (b) The equivalent
circuit of a shunt dc motor.
151
constant-voltage power supply, while a shunt dc motor is a motor whose field circuit gets its power
directly across the armature terminals of the motor. When the supply voltage to a motor is assumed
constant, there is no practical difference in behavior between these two machines. Unless otherwise
specified, whenever the behavior of a shunt motor is described, the separately excited motor is
included, too.
The Kirchhoff's voltage law (KVL) equation for the armature circuit of these motors is
𝑉𝑇 = 𝐸𝐴 + 𝐼𝐴 𝑅𝐴 (4.26)
A terminal characteristic of a machine is a plot of the machine's output quantities versus each other.
For a motor, the output quantities are shaft torque and speed, so the terminal characteristic of a
motor is a plot of its output torque versus speed.
How does a shunt dc motor respond to a load? Suppose that the load on the shaft of a shunt motor
is increased. Then the load torque 𝜏𝑙𝑜𝑎𝑑 will exceed the induced torque 𝜏𝑖𝑛𝑑 in the machine, and
the motor will start to slow down. When the motor slows down, its internal generated voltage drops
(𝐸𝐴 = 𝐾𝜙𝜔 ↓), so the armature current in the motor IA = (VT – EA↓)/RA increases. As the armature
current rises, the induced torque in the motor increases (𝜏𝑖𝑛𝑑 = 𝐾𝜙𝐼𝐴 ↑), and finally the induced
torque will equal the load torque at a lower mechanical speed of rotation ω.
The output characteristic of a shunt dc motor can be derived from the induced voltage and torque
equations of the motor plus Kirchhoff's voltage law. (KVL) The KVL equation for a shunt motor
is
𝑉𝑇 = 𝐸𝐴 + 𝐼𝐴 𝑅𝐴 (4.27)
𝑉𝑇 = 𝐾𝜙𝜔 + 𝐼𝐴 𝑅𝐴 (4.28)
152
Finally, solving for the motor's speed yields
𝑇𝑉 𝐴 𝑅
𝜔 = 𝐾𝜙 − (𝐾𝜙) 𝜏
2 𝑖𝑛𝑑
(4.31)
This equation is just a straight line with a negative slope. The resulting torque- speed characteristic
of a shunt dc motor is shown in Figure 4.16a.
It is important to realize that, in order for the speed of the motor to vary linearly with torque, the
other terms in this expression must be constant as the load changes. The terminal voltage supplied
by the dc power source is assumed to be constant if it is not constant, then the voltage variations
will affect the shape of the torque- speed curve.
Another effect internal to the motor that can also affect the shape of the torque-speed curve is
armature reaction. If a motor has armature reaction, then as its load increases, the flux-weakening
effects reduce its flux. As Equation (4.31) shows, the effect of a reduction in flux is to increase the
motor's speed at any given load over the speed it would run at without armature reaction. The
torque-speed characteristic of a shunt motor with armature reaction is shown in Figure 4.16b. If a
motor has compensating windings, of course there will be no flux-weakening problems in the
machine, and the flux in the machine will be constant.
If a shunt dc motor has compensating windings so that its flux is constant regardless of load, and
the motor's speed and armature current are known at any one value of load, then it is possible to
calculate its speed at any other value of load, as long as the armature current at that load is known
or can be determined.
153
Speed Control of Shunt DC Motor
How can the speed of a shunt dc motor be controlled? There are two common methods and one
less common method in use. The two common ways in which the speed of a shunt dc machine can
be controlled are by
To understand what happens when the field resistor of a dc motor is changed, assume that the field
resistor increases and observe the response. If the field resistance increases, then the field current
𝑉
decreases (𝐼𝐹 = 𝑅𝑇 ), and as the field current decreases, the flux ϕ decreases with it. A decrease in
𝐹
flux causes an instantaneous decrease in the internal generated voltage EA (=K ϕ↓ω), which causes
a large increase in the machine's armature current, since
𝑉𝑇 −𝐸𝐴 ↓
𝐼𝐴 ↑= 𝑅𝐴
The increase in current predominates over the decrease in flu x, and the induced torque rises:
𝜏𝑖𝑛𝑑 = 𝐾𝜙 ↓ 𝐼𝐴 ↑↑
However, as the motor speeds up, the internal generated voltage EA rises, causing IA to fall. As IA
falls, the induced torque 𝜏𝑖𝑛𝑑 falls too, and finally 𝜏𝑖𝑛𝑑 again equals 𝜏𝑙𝑜𝑎𝑑 at a higher steady-state
speed than originally.
2. Decreasing IF decreases ϕ.
3. Decreasing ϕ lowers EA (=𝐾𝜙 ↓ 𝜔).
𝑉𝑇 −𝐸𝐴 ↓
4. Decreasing EA increases IA (= )
𝑅𝐴
154
5. Increasing IA increases 𝜏𝑖𝑛𝑑 (= 𝐾𝜙 ↓ 𝐼𝐴 ↑↑ ), with the change in IA dominant over the
change in flux).
6. Increasing 𝜏𝑖𝑛𝑑 makes 𝜏𝑖𝑛𝑑 > 𝜏𝑙𝑜𝑎𝑑 , and the speed w increases.
7. Increasing to increases EA =𝐾𝜙𝜔 ↑ again.
8. Increasing EA decreases IA
(a) (b)
Figure 4.17: The effect of field resistance speed control on a shunt motor's torque-speed
characteristic: (a) over the motor's normal operating range: (b) over the entire range from no-load
to stall conditions.
The effect of increasing the field resistance on the output characteristic of a shunt motor is shown
in Fig. 4.17a. Notice that as the flux in the machine decreases, the no-load speed of the motor
increases, while the slope of the torque-speed curve becomes steeper. Naturally, decreasing RF
would reverse the whole process, and the speed of the motor would drop.
Fig. 4.17a shows the terminal characteristic of the motor over the range from no-load to full-load
conditions. Over this range, an increase in field resistance increases the motor's speed, as described
above in this section. For motors operating between no-load and full-load conditions, an increase
in RF may reliably be expected to increase operating speed.
Now examine Fig. 4.17b. This figure shows the terminal characteristic of the motor over the full
range from no-load to stall conditions. It is apparent from the figure that at very slow speeds an
increase in field resistance will actually decrease the speed of the motor. This effect occurs
because, at very low speeds, the increase in armature current caused by the decrease in E A. is no
longer large enough to compensate for the decrease in flux in the induced torque equation. With
the flux decrease actually larger than the armature current increase, the induced torque decreases,
and the motor slows down
155
Changing the Armature Voltage.
The second form of speed control involves changing the voltage applied to the armature of the
motor without changing the voltage applied to the field.
If the voltage VA is increased, then the armature current in the motor must rise [IA = (VA↑ - EA)/RA.
As IA increases, the induced torque 𝜏𝑖𝑛𝑑 = 𝐾𝜙𝐼𝐴 ↑ increases, making 𝜏𝑖𝑛𝑑 > 𝜏𝑙𝑜𝑎𝑑 , and the speed
ω of the motor increases.
Figure 4.18: The effect of armature voltage speed control on a shunt motor's torque-speed
characteristic.
But as the speed ω increases, the internal generated voltage EA (= 𝐾𝜙𝜔 ↑) increases, causing the
armature current to decrease. This decrease in IA decreases the induced torque, causing 𝜏𝑖𝑛𝑑 to
equal 𝜏𝑙𝑜𝑎𝑑 at a higher rotational speed ω.
If a resistor is inserted in series with the armature circuit, the effect is to drastically increase the
slope of the motor's torque-speed characteristic, making it operate more slowly if loaded (Fig.
156
4.19). This fact can easily be seen from Eq. (4.31). The insertion of a resistor is a very wasteful
method of speed control, since the losses in the inserted resistor are very large. For this reason, it
is rarely used. It will be found only in applications in which the motor spends almost all its time
operating at full speed or in applications too inexpensive to justify a better form of speed control.
The two most common methods of shunt motor speed control-field resistance variation and
armature voltage variation- have different safe ranges of operation.
Figure 4.19: The effect of armature resistance speed control on a shunt motor's torque-speed
characteristic.
In field resistance control, the lower the field current in a shunt (or separately excited) dc motor,
the faster it turns: and the higher the fi eld current, the slower it turns. Since an increase in field
current causes a decrease in speed, there is always a minimum achievable speed by field circuit
control. This minimum speed occurs when the motor's field circuit has the maximum permissible
current flowing through it.
If a motor is operating at its rated terminal voltage, power, and fi eld current, then it will be running
at rated speed, also known as base speed. Field resistance control can control the speed of the
motor for speeds above base speed but not for speeds below base speed. To achieve a speed slower
than base speed by field circuit control would require excessive field current, possibly burning up
the field windings.
In armature voltage control, the lower the armature voltage on a separately excited dc motor, the
slower it turns; and the higher the armature voltage, the faster it turns. Since an increase in armature
voltage causes an increase in speed, there is always a maximum achievable speed by armature
voltage control. This maximum speed occurs when the motor's armature voltage reaches its
maximum permissible level.
If the motor is operating at its rated voltage, field current, and power, it will be turning at base
speed. Armature voltage control can control the speed of the motor for speeds below base speed
157
but not for speeds above base speed. To achieve a speed faster than base speed by armature voltage
control would require excessive armature voltage, possibly damaging the armature circuit.
These two techniques of speed control are obviously complementary. Armature voltage control
works well for speeds below base speed, and field resistance or field current control works well
for speeds above base speed. By combining the two speed-control techniques in the same motor,
it is possible to get a range of speed variations of up to 40 to I or more. Shunt and separately excited
dc motors have excellent speed control characteristics.
Figure 4.20: Power and torque limits as a function of speed for a shunt motor under armature volt
and field resistance control.
There is a significant difference in the torque and power limits on the machine under these two
types of speed control. The limiting factor in either case is the heating of the armature conductors,
which places an upper limit on the magnitude of the armature current IA.
For armature voltage control, the flux in the motor is constant, so the maximum torque in the motor
is
This maximum torque is constant regardless of the speed of the rotation of the motor. Since the
power out of the motor is given by P = 𝜏𝜔, the maximum power of the motor at any speed under
armature voltage control is
Thus, the maximum power out of the motor is directly proportional to its operating speed under
armature voltage control
On the other hand, when field resistance control is used, the flux does change. In this form of
control, a speed increase is caused by a decrease in the machine's flux. In order for the armature
158
current limit not to be exceeded, the induced torque limit must decrease as the speed of the motor
increases. Since the power out of the motor is given by 𝑃 = 𝜏𝜔, and the torque limit decreases as
the speed of the motor increases, the maximum power out of a dc motor under field current control
is constant, while the maximum torque varies as the reciprocal of the motor's speed.
These shunt dc motor power and torque limitations for safe operation as a function of speed are
shown in Fig.4.20.
PMDC motor will have a lower induced torque 𝜏𝑖𝑛𝑑 per ampere of armature current IA than a shunt
motor of the same size and construction.
A series dc motor is a dc motor whose field windings consist of a relatively few turns connected
in series with the armature circuit. The equivalent circuit of a series dc motor is shown in Fig. 4.21.
In a series motor, the armature current, field current, and line current are all the same. The
Kirchhoff's voltage law equation for this motor is
𝑉𝑇 = 𝐸𝐴 + 𝐼𝐴 ( 𝑅𝐴 + 𝑅𝑆 ) (4.34)
The terminal characteristic of a series dc motor is very different from that of the shunt motor
previously studied. The basic behavior of a series dc motor is due to the fact that the flux is directly
proportional to the armature current, at least until saturation is reached. As the load on the motor
increases, its flux increases too. As seen earlier, an increase in flu x in the motor causes a decrease
in its speed. The result is that a series motor has a sharply drooping torque-speed characteristic.
The flux in this machine is directly proportional to its armature current (at least until the metal
saturates). Therefore, the flux in the machine can be given by
𝜙 = 𝑐𝐼𝐴 (4.36)
where c is a constant of proportionality. The induced torque in this machine is thus given by
In other words, the torque in the motor is proportional to the square of its armature current. As a
result of this relationship, it is easy to see that a series motor gives more torque per ampere than
any other dc motor. It is therefore used in applications requiring very high torques. Examples of
such applications are the starter motors in cars, elevator motors, and tractor motors in locomotives.
To determine the terminal characteristic of a series dc motor, an analysis will be based on the
assumption of a linear magnetization curve, and then the effects of saturation will be considered
in a graphical analysis.
The assumption of a linear magnetization curve implies that the flux in the motor will be given by
Eq. (4.36):
𝜙 = 𝑐𝐼𝐴
This equation will be used to derive the torque-speed characteristic curve for the series motor.
The derivation of a series motor's torque-speed characteristic starts with Kirchhoff's voltage law:
𝑉𝑇 = 𝐸𝐴 + 𝐼𝐴 ( 𝑅𝐴 + 𝑅𝑆 ) (4.38)
160
𝑖𝑛𝑑 𝜏
𝐼𝐴 = √ 𝐾𝑐
𝑖𝑛𝑑 𝜏
𝑉𝑇 = 𝐾𝜙𝜔 + √ 𝐾𝑐 ( 𝑅𝐴 + 𝑅𝑆 ) (4.39)
If the flux can be eliminated from this expression, it will directly relate the torque of a motor to its
speed. To eliminate the flux from the expression, notice that
𝜙
𝐼𝐴 =
𝑐
𝑐
𝜙 = √𝐾 √𝜏𝑖𝑛𝑑 (4.40)
Substituting Eq. (4.40) into Eq. (4.39) and solving for speed yields
𝑐 𝑖𝑛𝑑 𝜏
𝑉𝑇 = 𝐾√𝐾 √𝜏𝑖𝑛𝑑 𝜔 + √ 𝐾𝑐 ( 𝑅𝐴 + 𝑅𝑆 )
( 𝑅𝐴 +𝑅𝑆 )
√𝐾𝑐√𝜏𝑖𝑛𝑑 𝜔 = 𝑉𝑇 − √𝜏𝑖𝑛𝑑
√𝐾𝑐
𝑉𝑇 ( 𝑅𝐴 +𝑅𝑆 )
𝜔= −
√𝜏𝑖𝑛𝑑 √𝐾𝑐 𝐾𝑐
Notice that for an unsaturated series motor the speed of the motor varies as the reciprocal of the
square root of the torque. That is quite an unusual relationship! This ideal torque-speed
characteristic is plotted in Fig. 4.22.
One disadvantage of series motors can be seen immediately from this equation. When the torque
on this motor goes to zero, its speed goes to infinity. In practice, the torque can never go entirely
161
to zero because of the mechanical, core, and stray losses that must be overcome. However, if no
other load is connected to the motor, it can turn fast enough to seriously damage itself. Never
completely unload a series motor, and never connect one to a load by a belt or other mechanism
that could break. If that were to happen and the motor were to become unloaded while running,
the results could be serious.
Unlike with the shunt dc motor, there is only one efficient way to change the speed of a series dc
motor. That method is to change the terminal voltage of the motor. If the terminal voltage is
increased, the first term in Eq. (4.41) is increased, resulting in a higher speed for any given torque.
The speed of series dc motors can also be controlled by the insertion of a series resistor into the
motor circuit, but this technique is very wasteful of power and is used only for intermittent periods
during the start-up of some motors.
Until the last 40 years or so, there was no convenient way to change VT, so the only method of
speed control available was the wasteful series resistance method. That has all changed today with
the introduction of solid-state control circuits.
162
The Kirchhoff's voltage law equation for a compounded dc motor is
𝑉𝑇 = 𝐸𝐴 + 𝐼𝐴 ( 𝑅𝐴 + 𝑅𝑆 ) (4.42)
𝐼𝐴 = 𝐼𝐿 − 𝐼𝐹 (4.43)
𝑉
𝐼𝐹 = 𝑅𝑇 (4.44)
𝐹
The net magnetomotive force and the effective shunt field current in the compounded motor are
given by
And
𝑁𝑆𝐸 ℱ𝐴𝑅
𝐼𝐹 ∗ = 𝐼𝐹 ± 𝐼𝐴 − (4.46)
𝑁𝐹 𝑆𝐸 𝑁𝐹
Figure 4.23: The equivalent circuit of compounded dc motors: (a.) long-shunt connection:
(b) short-shunt connection.
where the positive sign in the equations is associated with a cumulatively compounded motor and
the negative sign is associated with a differentially compounded motor
In the cumulatively compounded dc motor, there is a component of flux which is constant and
another component which is proportional to its armature current (and thus to its load). Therefore,
the cumulatively compounded motor has a higher starting torque than a shunt motor (whose flux
is constant) but a lower starting torque than a series motor (whose entire flux is proportional to
armature current).
163
In a sense, the cumulatively compounded dc motor combines the best features of both the shunt
and the series motors. Like a series motor, it has extra torque for starting; like a shunt motor, it
does not overspeed at no load.
At light loads, the series field has a very small effect, so the motor behaves approximately as a
shunt dc motor. As the load gets very large, the series flux becomes quite important and the torque-
speed curve begins to look like a series motor's characteristic. A comparison of the torque-speed
characteristics of each of these types of machines is shown in Figures 4.24&4.25.
164
with armature reaction. It is so bad that a differentially compounded motor is unsuitable for any
application.
To make matters worse, it is impossible to start such a motor. At starting conditions, the armature
current and the series field current are very high. Since the series flux subtracts from the shunt
flux, the series field can actually reverse the magnetic polarity of the machine's poles. The motor
will typically remain still or turn slowly in the wrong direction while burning up, because of tile
excessive armature current. When this type of motor is to be started, its series fi eld must be short
circuited, so that it behaves as an ordinary shunt motor during the starting period.
Because of the stability problems of the differentially compounded de motor, it is almost never
intentionally used. However, a differentially compounded motor can result if the direction of
power flow reverses in a cumulatively compounded generator. For that reason, if cumulatively
compounded dc generators are used to supply power to a system, they will have a reverse-power
trip circuit to disconnect them from the line if the power flow reverses. No motor- generator set in
which power is expected to flow in both directions can use a differentially compounded motor,
and therefore it cannot use a cumulatively compounded generator.
The techniques available for the control of speed in a cumulatively compounded dc motor are the
same as those available for a shunt motor:
165
The arguments describing the effects of changing RF or VA are very similar to the arguments given
earlier for the shunt motor.
DC MOTOR STARTERS
In order for a dc motor to function properly on the job, it must have some special control and
protection equipment associated with it. The purposes of this equipment are
1. To protect the motor against damage due to short circuits in the equipment
2. To protect the motor against damage from long-term overloads
3. To protect the motor against damage from excessive starting currents
4. To provide a convenient manner in which to control the operating speed of the motor
Summary of Applications
166
4.9 Introduction to DC Generators
DC generators are dc machines used as generators. As previously pointed out, there is no real
difference between a generator and a motor except for the direction of power flow. There are five
major types of dc generators, classified according to the manner in which their field flux is
produced:
1. Separately excited generator: In a separately excited generator, the field flux is derived
from a separate power source independent of the generator itself.
2. Shunt generator: In a shunt generator, the field flux is derived by connecting the fie ld
circuit directly across the terminals of the generator.
3. Series generator: In a series generator, the field flux is produced by connecting the fie ld
circuit in series with the armature of the generator.
4. Cumulatively compounded generator: In a cumulatively compounded generator, both a
shunt and a series field are present, and their effects are additive.
5. Differentially compounded generator: In a differentially compounded generator, both a
shunt and a series fie ld are present, but their effects are subtractive.
These various types of dc generators differ in their terminal (voltage-current) characteristics, and
therefore in the applications to which they are suited.
DC generators are compared by their voltages, power ratings, efficiencies, and voltage regulations.
Voltage regulation (VR) is defined by the equation
𝑉𝑛𝑙 −𝑉𝑓𝑙
𝑆𝑅 = x100% (4.47)
𝑉𝑓𝑙
where 𝑉𝑛𝑙 , is the no-load terminal voltage of the generator and 𝑉𝑓𝑙 is the full-load terminal voltage
of the generator. It is a rough measure of the shape of the generator's voltage-current characteristic-
a positive voltage regulation means a drooping characteristic, and a negative voltage regulation
means a rising characteristic.
All generators are driven by a source of mechanical power, which is usually called the prime mover
of the generator. A prime mover for a dc generator may be a steam turbine, a diesel engine, or even
an electric motor. Since the speed of the prime mover affects the output voltage of a generator, and
since prime movers can vary widely in their speed characteristics, it is customary to compare the
167
voltage regulation and output characteristics of different generators, assuming constant speed
prime movers. Throughout this chapter, a generator's speed will be assumed to be constant unless
a specific statement is made to the contrary.
DC generators are quite rare in modern power systems. Even dc power systems such as those in
automobiles now use ac generators plus rectifiers to produce dc power.
The equivalent circuit of a dc generator is shown in Fig.4.27, and a simplified version of the
equivalent circuit is shown in Fig. 4.28. They look similar to the equivalent circuits of a dc motor,
except that the direction of current flow and the brush loss are reversed.
Figure 4.28: A simplified equivale circuit of a dc generator. with RF combining the resistances of
the field coils and the variable control resistor
𝐼𝐴 = 𝐼𝐿 (4.48)
The terminal characteristic of a device is a plot of the output quantities of the device versus each
other. For a dc generator, the output quantities are its terminal voltage and line current. The
terminal characteristic of a separately excited generator is
168
Figure 4.29: A separately excited de generator.
thus, a plot of VT versus IL for a constant speed ω. By Kirchhoff's voltage law, the terminal voltage
is
𝑉𝑇 = 𝐸𝐴 − 𝐼𝐴 𝑅𝐴 (4.49)
Since the internal generated voltage is independent of IA, the terminal characteristic of the
separately excited generator is a straight line, as shown in Fig. 4.30a.
What happens in a generator of this sort when the load is increased? When the load supplied by
the generator is increased, IL (and therefore IA) increases. As the armature current increases, the
IARA drop increases, so the terminal voltage of the generator falls.
This terminal characteristic is not always entirely accurate. In generators without compensating
windings, an increase in IA causes an increase in armature reaction, and armature reaction causes
flux weakening. This flux weakening causes a decrease in 𝐸𝐴 = 𝐾𝜙 ↓ 𝜔 which further decreases
the terminal voltage of the generator. The resulting terminal characteristic is shown in Fig.4.30b.
In all future plots, the generators will be assumed to have compensating windings unless stated
otherwise. However, it is important to realize that armature reaction can modify the characteristics
if compensating windings are not present.
The terminal voltage of a separately excited dc generator can be controlled by changing the internal
generated voltage EA of the machine. By Kirchhoff's voltage law 𝑉𝑇 = 𝐸𝐴 − 𝐼𝐴 𝑅𝐴 , so if EA
increases, VT will increase, and if EA decreases, VT will decrease. Since the internal generated
voltage EA is given by the equation 𝐸𝐴 = 𝐾𝜙𝜔 there are two possible ways to control the voltage
of this generator:
169
𝑉
2. Change the field current. If RF is decreased. then the field current increases (𝐼𝐹 = 𝑅 𝐹↓) Therefore,
𝐹
the flux in the machine increases. As the flux rises, 𝐸𝐴 = 𝐾𝜙 ↑ 𝜔 must rise too, so 𝑉𝑇 = 𝐸𝐴 ↑
−𝐼𝐴 𝑅𝐴 increases.
Figure 4.30: The terminal characteristic of a separately excited dc generator (a) with and (b)
without compensating windings
In many applications, the speed range of the prime mover is quite limited, so the terminal voltage
is most commonly controlled by changing the field current. A separately excited generator
driving a resistive load is shown in Fig.4.31a. Fig 4.31b shows the effect of a decrease in field
resistance on the terminal voltage of the generator when it is operating under a load.
(a)
(b)
Figure: 4.31(a) A separately excited de generator with a resistive load. (b) The effect of a
decrease in field resistance on the output voltage of the generator.
As with de motors, it is customary to define an equivalent field current that would produce the
same output voltage as the combination of all the magnetomotive forces in the machine. The
170
resulting voltage EA0 can then be determined by locating that equivalent field current on the
magnetization curve. The equivalent field current of a separately excited dc generator is given by
ℱ𝐴𝑅
𝐼𝐹 ∗ = 𝐼𝐹 − (4.51)
𝑁𝐹
Also, the difference between the speed of the magnetization curve and the real speed of the
generator must be taken int o account using Eq. (4.52):
𝐸𝐴 𝑛
=𝑛 (4.52)
𝐸𝐴0 0
𝐼𝐴 = 𝐼𝐹 + 𝐼𝐿 (4.53)
The Kirchhoff's voltage law equation for the armature circuit of this machine is
𝑉𝑇 = 𝐸𝐴 − 𝐼𝐴 𝑅𝐴 (4.54)
This type of generator has a distinct advantage over the separately excited dc generator in that no
external power supply is required for the field circuit. But that leaves an important question
unanswered: If the gene rat or supplies its own field current, how does it get the initial field flux
to start when it is first turned on?
171
Voltage Build up in a Shunt DC Generator
Assume that the generator in Fig.4.32 has no load connected to it and that the prime mover starts
to turn the shaft of the generator. How does an initial voltage appear at the terminals of the
machine?
The voltage buildup in a dc generator depends on the presence of a residual flux in the poles of the
generator. When a generator first starts to turn, an internal voltage will be generated which is given
by
𝐸𝐴 = 𝐾𝜙𝑟𝑒𝑠 𝜔
This voltage appears at the terminals of the generator (it may only be a volt or two). But when that
𝑉𝑇 ↑
voltage appears at the terminals, it causes a current to flow in the generator's field coil (𝐼𝐹 = ).
𝑅𝐹
This fi eld current produces a magnetomotive force in the poles, which increases the flux in them.
1lle increase in flux causes an increase in 𝐸𝐴 = 𝐾𝜙 ↑ 𝜔, which increases the terminal voltage VT,
When VT rises, IF increases further, increasing the flux ϕ more, which increases EA, etc.
This voltage buildup behavior is shown in Fig 4.33. Notice that it is the effect of magnetic
saturation in the pole faces which eventually limits the terminal voltage of the generator.
172
1. There may be no residual magnetic flux in the generator to start the process going, If the
residual flux 𝜙𝑟𝑒𝑠 = 0, then EA = 0, and the voltage never builds up. If this problem occurs,
disconnect the field from the armature circuit and connect it directly to an external dc source
such as a battery. The current flow from this external dc source will leave a residual flux in
the poles, which will then allow normal starting. This procedure is known as "flashing the
field,"
2. The direction of rotation of the generator may have been reversed, or the connections of the
field may have been reversed. In either case, the residual flux produces an internal generated voltage
EA. The voltage EA produces a field current which produces a flux opposing the residual flux, instead
of adding to it. Under these circumstances, the flux actually decreases below 𝜙𝑟𝑒𝑠 and no voltage can
ever build up.
If this problem occurs, it can be fixed by reversing the direction of rotation, by reversing the field
connections, or by flashing the field with the opposite magnetic polarity.
Figure 4.34: The effect of shunt field resistance on no-load terminal voltage in a dc generator. If
RF > R2 (the critical resistance). then the generator's voltage will never build up.
3. The field resistance may be adjusted to a value greater than the critical resistance. To
understand this problem, refer to Fig.4.34. Normally, the shunt generator will build up to the
point where the magnetization curve intersects the field resistance line. If the field resistance
has the value shown at R2 in the figure, its line is nearly parallel to the magnetization curve.
At that point, the voltage of the generator can fluctuate very widely with only tiny changes
in RF or IA. This value of the resistance is called the critical resistance. If RF exceeds the
critical resistance (as at R3 in the figure), then the steady-state operating voltage is essentially
at the residual level, and it never builds up. The solution to this problem is to reduce RF.
173
Since the voltage of the magnetization curve varies as a function of shaft speed, the critical
resistance also varies with speed. In general, the lower the shaft speed, the lower the critical
resistance.
The Terminal Characteristic of a Shunt DC Generator
The terminal characteristic of a shunt dc generator differs from that of a separately excited dc
generator, because the amount of field current in the machine depends on its terminal voltage. To
understand the terminal characteristic of a shunt generator, start with the machine unloaded and
add loads, observing what happens.
As the load on the generator is increased, IL- increases and so IA = IF + IL↑ also increases. An
increase in IA increases the armature resistance voltage drop IARA, causing 𝑉𝑇 = 𝐸𝐴 − 𝐼𝐴 ↑ 𝑅𝐴 to
decrease. This is precisely the same behavior observed in a separately excited generator. However,
when Vr decreases, the field current in the machine decreases with it. This causes the flux in the
machine to
174
Changing the field resistor is the principal method used to control terminal voltage in real shunt
𝑉
generators. If the field resistor RF is decreased, then the field current 𝐼𝐹 = 𝑅 𝑇 increases. When IF
𝐹↓
increases, the machine's flux ϕ increases, causing the internal generated voltage EA to increase.
The increase in EA causes the terminal voltage of the generator to increase as well.
𝑉𝑇 = 𝐸𝐴 − 𝐼𝐴 ( 𝑅𝐴 + 𝑅𝑆 ) (4.55)
The magnetization curve of a series dc generator looks very much like the magnetization curve of
any other generator. At no load, however, there is no field current, so VT is reduced to a small level
given by the residual flux in the machine. As the load increases, the field current rises, so EA. rises
rapidly. The 𝐼𝐴 ( 𝑅𝐴 + 𝑅𝑆 ) drop goes up too, but at first the increase in EA. goes up more rapidly
than the 𝐼𝐴 ( 𝑅𝐴 + 𝑅𝑆 ) drop rises, so VT increases. After a while, the machine approaches
saturation, and EA, becomes almost constant. At that point, the resistive drop is the predominant
effect, and VT starts to fall.
175
Figure 4.37: Derivation of the terminal characteristic for a series dc generator.
Figure 4.38: A series generator terminal characteristic with large armature reaction effects.
suitable for electric welders.
This type of characteristic is shown in Fig.4.37. It is obvious that this machine would make a bad
constant-voltage source. In fact, its voltage regulation is a large negative number.
Series generators are used only in a few specialized applications, where the steep voltage
characteristic of the device can be exploited. One such application is arc welding. Series generators
used in arc welding are deliberately designed to have a large armature reaction, which gives them
a terminal characteristic like the one shown in Fig.4.38. Notice that when the welding electrodes
make contact with each other before welding commences, a very large current flow. As the
operator separates the welding electrodes, there is a very steep rise in the generator's voltage, while
the current remains high. This voltage ensures that a welding arc is maintained through the air
between the electrodes.
176
ℱ𝑛𝑒𝑡 = ℱ𝐹 + ℱ𝑆𝐸 − ℱ𝐴𝑅 (4.56)
where ℱF is the shunt field magnetomotive force, ℱSE is the series field magnetomotive force, and
ℱAR is the armature reaction magnetomotive force. The equivalent effective shunt field current for
this machine is given by
The other voltage and current relationships for this generator are
𝐼𝐴 = 𝐼𝐹 + 𝐼𝐿 (4.58)
𝑉𝑇 = 𝐸𝐴 − 𝐼𝐴 ( 𝑅𝐴 + 𝑅𝑆 ) (4.59)
𝑉
𝐼𝐹 = 𝑁𝑇 (4.60)
𝐹
Figure 4.39: The equivalent circuit of a cumulatively compounded dc generator with a long-
shunt connection.
177
Suppose that the load on the generator is increased. Then as the load increases, the load current IL
increases. Since IA = IF + IL↑ the armature current IA increases too. At this point two effects occur
in the generator:
1. As IA increases, the IA (RA + RS) voltage drop increases as well. This tends to cause a decrease
in the terminal voltage 𝑉𝑇 = 𝐸𝐴 − 𝐼𝐴 ↑ ( 𝑅𝐴 + 𝑅𝑆 ).
2. As IA increases, the series field magnetomotive force ℱ𝑆𝐸 = 𝑁𝑆𝐸 ℱ𝑆𝐸 increases too. This
increases the total magnetomotive force ℱ𝑡𝑜𝑡 = 𝑁𝐹 ℱ𝐹 + 𝑁𝑆𝐸 𝐼𝐴 ↑ which increases the flux in
the generator. The increased flux in the generator increases EA, which in turn tends to make
𝑉𝑇 = 𝐸𝐴 ↑ −𝐼𝐴 ( 𝑅𝐴 + 𝑅𝑆 ) rise.
These two effects oppose each other, with one tending to increase VT and the other tending to
decrease VT. Which effect predominates in a given machine? It all depends on just how many
series turns were placed on the poles of the machine. The question can be answered by taking
several individual cases:
1. Few series turn (𝑁𝑆𝐸 small). If there are only a few series turns, the resistive voltage drop
effect wins hands down. The voltage falls off just as in a shunt generator, but not quite as
steeply (Fig.4.40). This type of construction, where the full-load terminal voltage is less than
the no-load terminal voltage, is called under compounded.
2. More series turns (𝑁𝑆𝐸 larger). If there are a few more series turns of wire on the poles, then
at first the flux-strengthening effect wins, and the terminal voltage rises with the load.
However, as the load continues to increase, magnetic saturation sets in, and the resistive drop
becomes stronger than the flux increase effect. In such a machine, the terminal voltage first
rises and then falls as the load increases. If VT at no load is equal to VT at full load, the
generator is called flat-compounded.
3. Even more series turns are added (𝑁𝑆𝐸 large). If even more series turns are added to the
generator, the flux-strengthening effect predominates for a longer time before the resistive
drop takes over. The result is a characteristic with the full-load terminal voltage actually
higher than the no-load terminal voltage. If VT at a full load exceeds VT at no load, the
generator is called over compounded.
178
Figure 4.40: Terminal characteristics of cumulatively compounded dc generators.
The techniques available for controlling the terminal volt age of a cumulatively compounded dc
generator are exactly the same as the techniques for controlling the voltage of a shunt dc generator:
total magnetomotive force in the generator. ℱtot, increases, the flux ϕ in the machine
increases, and 𝐸𝐴 = 𝐾𝜙 ↑ 𝜔 increases. Finally, an increase in EA raises VT.
5. The Differentially Compounded DC Generator
A differentially compounded dc generator is a generator with both shunt and series fields, but this
time their magnetomotive forces subtract from each other. The equivalent circuit of a differentially
compounded dc generator is shown in Fig.4.41. Notice that the armature current is now flowing
out of a dotted coil end, while the shunt field current is flowing into a dotted coil end. In this
machine, the net magnetomotive force is
and the equivalent shunt field current due to the series field and armature reaction is given by
𝑁𝑆𝐸 ℱ𝐴𝑅
𝐼𝑒𝑞 = − 𝐼 − (4.63)
𝑁𝐹 𝐴 𝑁𝐹
𝐼𝐹 ∗ = 𝐼𝐹 + 𝐼𝑒𝑞 (4.64)
179
Or
𝑁𝑆𝐸 ℱ𝐴𝑅
𝐼𝐹 ∗ = 𝐼𝐹 − 𝐼 − (4.65)
𝑁𝐹 𝐴 𝑁𝐹
Figure 4.41: The equivalent circuit of a differentially compounded dc generator with a long-shunt
connection.
Like the cumulatively compounded generator, the differentially compounded generator can be
connected in either long-shunt or short-shunt fashion.
In the differentially compounded dc generator, the same two effects occur that were present in the
cumulatively compounded dc generator. This time, though, the effects both act in the same
direction. They are
1. As IA increases, the 𝐼𝐴 ( 𝑅𝐴 + 𝑅𝑆 ) voltage drop increases as well. This increase tends to cause
the terminal voltage to decrease 𝑉𝑇 = 𝐸𝐴 − 𝐼𝐴 ↑ ( 𝑅𝐴 + 𝑅𝑆 ).
2. As IA increases, the series field magnetomotive force ℱ𝑆𝐸 = 𝑁𝑆𝐸 𝐼𝐴 increases too. This
increase in series field magnetomotive force reduces the net magnetomotive force on the
generator (ℱ𝑡𝑜𝑡 = 𝑁𝐹 𝐼𝐹 − 𝑁𝑆𝐸 𝐼𝐴 ↑), which in turn reduces the net flux in the generator. A
decrease in flux decreases EA, which in turn decreases VT
Since both these effects tend to decrease VT, the voltage drops drastically as the load is increased
on the generator. A typical terminal characteristic for a differentially compounded dc generator is
shown in Fig.4.42.
180
Voltage Control of Differentially Compounded DC Generators
Even though the voltage drop characteristics of a differentially compounded dc generator are
quite bad, it is still possible to adjust the terminal voltage at any given load setting. The
techniques available for adjusting terminal voltage are exactly the same as those for shunt and
cumulatively compounded dc generators:
SUMMARY
There are several types of dc motors, differing in the manner in which their field fluxes are derived.
These types of motors are separately excited, shunt, permanent-magnet, series, and compounded.
The manner in which the flux is derived affects the way it varies with the load, which in turn affects
the motor's overall torque-speed characteristic.
A shunt or separately excited dc motor has a torque- speed characteristic whose speed drops
linearly with increasing torque. Its speed can be controlled by changing its fi e ld current, its
armature voltage, or its armature resistance.
A permanent-magnet dc motor is the same basic machine except that its flux is derived from
permanent magnets. Its speed can be controlled by any of the above methods except varying the
field current.
A series motor has the highest starting torque of any dc motor but tends to overspeed at no load. It
is used for very high-torque applications where speed regulation is not important, such as a car
starter.
A cumulatively compounded dc motor is a compromise between the series and the shunt motor,
having some of the best characteristics of each. On the other hand, a differentially compounded dc
motor is a complete disaster. It is unstable and tends to overspeed as load is added to it.
DC generators are dc machines used as generators. There are several different types of dc
generators, differing in the manner in which their field fluxes are derived. These methods affect
the output characteristics of the different types of generators. The common dc generator types are
separately excited, shunt, series, cumulatively compounded, and differentially compounded.
181
The shunt and compounded dc generators depend on the nonlinearity of their magnetization curves
for stable output voltages. If the magnetization curve of a dc machine were a straight line, then the
magnetization curve and the terminal voltage line of the generator would never intersect. There
would thus be no stable no-load voltage for the generator. Since nonlinear effects are at the heart
of the generator's operation, the output voltages of dc generators can only be determined
graphically or numerically by using a computer.
Today, dc generators have been replaced in many applications by ac power sources and solid-state
electronic components. This is true even in the automobile, which is one of the most common users
of dc power.
Objective test-4
is to
(a) obtain a coil span of 180º (electrical) (c) distribute the winding uniformly
(b) ensure the addition of e.m.fs. of under different poles
consecutive turns (d) obtain a full-pitch winding.
6. In a 4-pole, 35 slot dc armature, 180 electrical-degree coil span will be obtained when coils
occupy ............ slots.
(a) 1 and 10 (b) 1 and 9 (c) 2 and 11 (d) 3 and 12
182
7. The armature of a dc generator has a 2-layer lap-winding housed in 72 slots with six
conductors/slot. What is the minimum number of commutator bars required for the armature?
(a) 72 (b) 432 (c) 216 (d) 36
8. The sole purpose of a commutator in a dc Generator is to
(a) increase output voltage (c) provide smoother output
(b) reduce sparking at brushes (d) convert the induced ac into dc
9. For a 4-pole, 2-layer, dc, lap-winding with 20 slots and one conductor per layer, the number
of commutator bars is
(a) 80 (b) 20 (c) 40 (d) 160
10. A 4-pole, 12-slot lap-wound dc armature has two coil-sides/slot. Assuming single turn coils
and progressive winding, the back pitch would be
(a) 5 (b) 7 (c) 3 (d) 6
11. If in the case of a certain dc armature, the number of commutator segments is found either one
less or more than the number of slots, the armature must be having a simplex ............ winding.
(a) wave (c) frog leg
(b) lap (d) multielement
12. Lap winding is suitable for ............ current, ............ voltage dc generators.
(a) high, low (c) low, low
(b) low, high (d) high, high
13. The series field of a short-shunt dc generator is excited by ............ currents.
(a) shunt (b) armature (c) load (d) external
14. In a dc generator, the generated e.m.f. is directly proportional to the
(a) field current (c) number of armature parallel paths
(b) pole flux (d) number of dummy coils
15. In a 12-pole triplex lap-wound dc armature, each conductor can carry a current of 100 A.The
rated current of this armature is ............ampere.
(a) 600 (b) 1200 (c) 2400 (d) 3600
16. The commercial efficiency of a shunt generator is maximum when its variable loss equals
............ loss.
(a) constant (c) iron
(b) stray (d) friction and windage
17. In small dc machines, armature slots are sometimes not made axial but are skewed. Though
skewing makes winding a little more difficult, yet it results in
183
(a) quieter operation (c) saving of copper
(b) slight decrease in losses (d) both (a) and (b)
18. The critical resistance of the dc generator is the resistance of
(a) armature (b) field (c) load (d) brushes
184
Chapter Five
Synchronous Machines
Descriptions: Construction, equivalent circuit, parameter testing, and characteristics of
synchronous machines as an alternator, motor operation of synchronous machine.
Background
A.C. generators or alternators (as they are usually called) operate on the same fundamental
principles of electromagnetic induction as dc generators. They also consist of an armature winding
and a magnetic field. But there is one important difference between the two. Whereas in dc
generators, the armature rotates and the field system is stationary, the arrangement in alternators
is just the reverse of it.
The stator consists of a cast-iron frame, which supports the armature core, having slots on its inner
periphery for housing the armature conductors. The rotor is like a flywheel having alternate N and
S poles fixed to its outer rim. The magnetic poles are excited (or magnetized) from direct current
supplied by a dc source at 125 to 600 volts. In most cases, necessary exciting (or magnetizing)
current is obtained from a small dc shunt generator which is belted or mounted on the shaft of the
alternator itself. Because the field magnets are rotating, this current is supplied through two
sliprings. As the exciting voltage is relatively small, the slip-rings and brush gear are of light
185
construction. Recently, brushless excitation systems have been developed in which a 3-phase ac
exciter and a group of rectifiers supply dc to the alternator. Hence, brushes, slip-rings and
commutator are eliminated.
When the rotor rotates, the stator conductors (being stationary) are cut by the magnetic flux, hence
they have induced e.m.f. produced in them. Because the magnetic poles are alternately N and S,
they induce an e.m.f. and hence current in armature conductors, which first flows in one direction
and then in the other. Hence, an alternating e.m.f. is produced in the stator conductors (i) whose
frequency depends on the number of N and S poles moving past a conductor in one second and (ii)
whose direction is given by Fleming's Right-hand rule.
Stationary Armature
Advantages of having stationary armature (and a rotating field system) are : 1. The output current
can be led directly from fixed terminals on the stator (or armature windings) to the load circuit,
without having to pass it through brush-contacts.
The equivalent circuit of a salient-pole synchronous generator is shown in Fig. 37.71 (a). The
component currents I d and I q provide component voltage drops jId Xd and j I q X q as shown in
Fig. 37.71(b) for a lagging load power factor The armature current I a has been resolved into its
rectangular components with respect to the axis for excitation voltage E0 .The angle ψ between
E0 and I a is known as the internal power factor angle. The vector for the armature resistance drop
I a Ra is drawn parallel to I a . Vector for the drop I d Xd is drawn perpendicular to I d whereas
that for I q × X q is drawn perpendicular to I q . The angle δ between E0 and V is called the power
angle. Following phasor relationships are obvious from Fig. 37.71 (b) E0 = V + IaRa + jId Xd +
jIq Xq and I a = Id + Iq If Ra is neglected the phasor diagram becomes as shown in Fig. 5.72 (a).
In this case, E0 = V + jId Xd + jI q X q
Alternator on Load
As the load on an alternator is varied, its terminal voltage is also found to vary as in dc generators.
186
3. voltage drop due to armature reaction
Armature Resistance The armature resistance/phase Ra causes a voltage drop/phase of IRa which
is in phase with the armature current I. However, this voltage drop is practically negligible. (b)
Armature Leakage Reactance When current flows through the armature conductors, fluxes are set
up which do not cross the air-gap, but take different paths. Such fluxes are known as leakage
fluxes. Various types of leakage fluxes are shown in Fig. 5.22.
The leakage flux is practically independent of saturation, but is dependent on I and its phase angle
with terminal voltage V. This leakage flux sets up an e.m.f. of self-inductance which is known as
reactance e.m.f. and which is ahead of I by 90°. Hence, armature winding is assumed to possess
leakage reactance XL (also known as Potier rectance XP ) such that voltage drop due to this equals
IXL . A part of the generated e.m.f. is used up in overcoming this reactance e.m.f.
E = V + I (R + jXL )
Armature Reaction
As in dc generators, armature reaction is the effect of armature flux on the main field flux. In the
case of alternators, the power factor of the load has a considerable effect on the armature
reaction. We will consider three cases : (i) when load of p.f. is unity (ii) when p.f. is zero lagging
and (iii) when p.f. is zero leading.
Synchronous Reactance
From the above discussion, it is clear that for the same field excitation, terminal voltage is
decreased from its no-load value E0 to V (for a lagging power factor). This is because of
187
may be accounted for by assumiung the presence of a fictitious reactance Xa in the armature
winding. The value of Xa is such that IXa represents the voltage drop due to armature reaction.The
leakage reactance XL (or XP) and the armature reactance Xa may be combined to give
under load is = IRa + jIXS = I(Ra + jXS) = IZS where ZS is known as synchronous impedance
of the armature, the word ‘synchronous’ being used merely as an indication that it refers to
the working conditions.
Hence, we learn that the vector difference between no-load voltage E0 and terminal voltage
V is equal to IZS, as shown in Fig. 37.26.
Voltage Regulation
It is clear that with change in load, there is a change in terminal voltage of an alternator. The
magnitude of this change depends not only on the load but also on the load power factor.
The voltage regulation of an alternator is defined as “the rise in voltage when full-load is removed
(field excitation and speed remaining the same) divided by the rated terminal voltage.”
Note. (i) E0 − V is the arithmetical difference and not the vectorial one. (ii) In the case of leading
load p.f., terminal voltage will fall on removing the full-load. Hence, regulation is negative in that
case. (iii) The rise in voltage when full-load is thrown off is not the same as the fall in voltage
when full-load is applied. Voltage characteristics of an alternator are shown in Fig. 37.29.
In the case of small machines, the regulation may be found by direct loading. The procedure is as
follows : The alternator is driven at synchronous speed and the terminal voltage is adjusted to its
rated value V. The load is varied until the wattmeter and ammeter (connected for the purpose)
indicate the rated values at desired p.f. Then the entire load is thrown off while the speed and field
188
excitation are kept constant. The open-circuit or no-load voltage E0 is read. Hence, regulation can
be found from
E
0 −V
% regn = V × 100
In the case of large machines, the cost of finding the regulation by direct loading becomes
prohibitive. Hence, other indirect methods are used as discussed below. It will be found that
all these methods differ chiefly in the way the no-load voltage E0 is found in each case.
2. Open-circuit/No-load characteristic.
Armature resistance Ra per phase can be measured directly by voltmeter and ammeter method
or by using Wheatstone bridge. However, under working conditions, the effective value of
Ra is increased due to ‘skin effect’. The value of Ra so obtained is increased by 60% or so to
allow for this effect. Generally, a value 1.6 times the d.c. value is taken.
O.C. Characteristic
As in dc machines, this is plotted by running the machine on no-load and by noting the values
of induced voltage and field excitation current. It is just like the B-H curve.
S.C. Characteristic
189
Operation of a Salient Pole Synchronous Machine
A multipolar machine with cylindrical rotor has a uniform air-gap, because of which its reactance
remains the same, irrespective of the spatial position of the rotor. However, a synchronous machine
with salient or projecting poles has non-uniform air-gap due to which its reactance varies with the
rotor position. Consequently, a cylindrical rotor machine possesses one axis of symmetry (pole
axis or direct axis) whereas salient-pole machine possesses two axes of geometric symmetry (i)
field poles axis, called direct axis or d-axis and (ii)axis passing through the centre of the interpolar
space, called the quadrature axis or qaxis, as shown in Fig. 37.70.
The operation of connecting an alternator in parallel with another alternator or with common bus-
bars is known as synchronizing. Generally, alternators are used in a power system where they are
in parallel with many other alternators. It means that the alternator is connected to a live system of
constant voltage and constant frequency. Often the electrical system to which the alternator is
connected, has already so many alternators and loads connected to it that no matter what power is
delivered by the incoming alternator, the voltage and frequency of the system remain the same. In
that case, the alternator is said to be connected to infinite bus-bars. It is never advisable to connect
a stationary alternator to live bus-bars, because, stator induced e.m.f. being zero, a short-circuit
will result. For proper synchronization of alternators, the following three conditions must be
satisfied : 1. The terminal voltage (effective) of the incoming alternator must be the same as bus-
bar voltage. 2. The speed of the incoming machine must be such that its frequency (= PN/120)
equals bus-bar frequency. 3. The phase of the alternator voltage must be identical with the phase
of the bus-bar voltage. It means that the switch must be closed at (or very near) the instant the two
voltages have correct phase relationship. Condition (1) is indicated by a voltmeter, conditions (2)
and (3) are indicated by synchronizing lamps or a synchronoscope.
Synchronizing of Alternators
Single-phase Alternators Suppose machine 2 is to be synchronized with or ‘put on’ the bus-bars to
which machine 1 is already connected. This is done with the help of two lamps L1 and L2 (known
as synchronizing lamps) connected as shown in Fig. 37.74. It should be noted that E1 and E2 are
in-phase relative to the external circuit but are in direct phase opposition in the local circuit (shown
dotted).
190
If the speed of the incoming machine 2 is not brought up to that of machine 1, then its frequency
will also be different, hence there will be a phase-difference between their voltages (even when
they are equal in magnitude, which is determined by field excitation). This phase-difference will
be continuously changing with the changes in their frequencies. The result is that their resultant
voltage will undergo changes similar to the frequency changes of beats produced, when two sound
sources of nearly equal frequency are sounded together, as shown in Fig. 37.75.
Sometimes the resultant voltage is maximum and some other times minimum. Hence, the current
is alternatingly maximum and minimum. Due to this changing current through the lamps, a flicker
will be produced, the frequency of flicker being (f 2 ∼ f 1 ). Lamps will dark out and glow up
alternately. Darkness indicates that the two voltages E1 and E2 are in exact phase opposition
relative to the local circuit and hence there is no resultant current through the lamps. Synchronizing
is done at the middle of the dark period. That is why, sometimes, it is known as ‘lamps dark’
synchronizing. Some engineers prefer ‘lamps bright’ synchronization because of the fact the lamps
are much more sensitive to changes in voltage at their maximum brightness than when they are
dark. Hence, a sharper and more accurate synchronization is obtained. In that case, the lamps are
connected as shown in Fig. 37.76. Now, the lamps will glow brightest when the two voltages are
in phase with the bus-bar voltage because then voltage across them is twice the voltage of each
machine.
Three-phase Alternators
In 3-φ alternators, it is necessary to synchronize one phase only, the other two phases will then be
synchronized automatically. However, first it is necessary that the incoming alternator is correctly
‘phased out’ i.e. the phases are connected in the proper order of R, Y, B and not R, B, Y etc.
In this case, three lamps are used. But they are deliberately connected asymmetrically, as shown
in Fig. 37.77 and 37.78. This transposition of two lamps, suggested by Siemens and Halske, helps
to indicate whether the incoming machine is running too slow. If lamps were connected
symmetrically, they would dark out or glow up simultaneously (if the phase rotation is the same
as that of the bus-bars). Lamp L1 is connected between R and R′, L2 between Y and B′ (not Y and
Y′) and L3 between B and Y′ (and not B and B′), as shown in Fig. 37.78. Voltage stars of two
machines are shown superimposed on each other in Fig. 37.79.
191
OBJECTIVE TESTS – 5
1. The frequency of voltage generated by an alternator having 4-poles and rotating at 1800 r.p.m.
is .......hertz.
(a) 60 (b) 7200 (c) 120 (d) 450
2. A 50-Hz alternator will run at the greatest possible speed if it is wound for ....... poles.
(a) 8 (b) 6 (c) 4 (d) 2
3. The main disadvantage of using short-pitch winding in alternators is that it
(a) reduces harmonics in the generated (c) produces asymmetry in the three
voltage phase windings
(b) reduces the total voltage around the (d) increases Cu of end connections.
armature coils
4. Three-phase alternators are invariably Y-connected because
(a) magnetic losses are minimized (c) smaller conductors can be used
(b) less turns of wire are required (d) higher terminal voltage is obtained
5. The winding of a 4-pole alternator having 36 slots and a coil span of 1 to 8 is short-pitched by
....... degrees.
(a) 140 (b) 80 (c) 20 (d) 40
6. If an alternator winding has a fractional pitch of 5/6, the coil span is ....... degrees.
(a) 300 (b) 150 (c) 30 (d) 60
7. The harmonic which would be totally eliminated from the alternator e.m.f. using a fractional
pitch of 4/5 is
(a) 3rd (b) 7th
192
(a) 5α (b) α/5 (c) 25α (d) α/25
(e)
10. Regarding distribution factor of an armature winding of an alternator which statement is false?
(a) it decreases as the distribution of coils (c) it is not affected by the type of
(slots/pole) increases winding either lap, or wave
(b) higher its value, higher the induced (d) it is not affected by the number of
e.m.f. per phase turns per coil.
11. When speed of an alternator is changed from 3600 r.p.m. to 1800 r.p.m., the generated
e.m.f./phases will become
(a) one-half (b)twice (c) four times (d) one-fourth.
12. The magnitude of the three voltage drops in an alternator due to armature resistance, leakage
reactance and armature reaction is solely determined by
(a) load current, Ia (c) whether it is a lagging or leading p.f.
(b) p.f. of the load load
(d) field construction of the alternator.
13. Armature reaction in an alternator primarily affects
(a) rotor speed (c) frequency of armature current
(b) terminal voltage per phase (d) generated voltage per phase.
14. Under no-load condition, power drawn by the prime mover of an alternator goes to
(a) produce induced e.m.f. in armature (c) produce power in the armature
winding (d) meet Cu losses both in armature and
(b) meet no-load losses rotor windings.
15. As load p.f. of an alternator becomes more leading, the value of generated voltage required to
give rated terminal voltage
(a) increases (c) decreases
(b) remains unchanged (d) varies with rotor speed.
16. With a load p.f. of unity, the effect of armature reaction on the main-field flux of an alternator
is
(a) distortional (b) magnetizing
193
(c) demagnetizing (d) nominal
17. At lagging loads, armature reaction in an alternator is
(a) cross-magnetizing (c) non-effective
(b) demagnetizing (d) magnetizing
18. At leading p.f., the armature flux in an alternator ....... the rotor flux.
(a) opposes (c) distorts
(b) aids (d) does not affect
19. The voltage regulation of an alternator having 0.75 leading p.f. load, no-load induced e.m.f.
of 2400V and rated terminal voltage of 3000V is ............... percent.
(a) 20 (b) − 20 (c) 150 (d) − 26.7
20. If, in a 3-φ alternator, a field current of 50A produces a full-load armature current of 200 A on
short-circuit and 1730 V on open circuit, then its synchronous impedance is ....... ohm.
(a) 8.66 (b) 4 (c) 5 (d) 34.6
21. The power factor of an alternator is determined by its
(a) speed (c) excitation
(b) load (d) prime mover
22. For proper parallel operation, ac polyphase alternators must have the same
(a) speed (c) kVA rating
(b) voltage rating (d) excitation.
23. Of the following conditions, the one which does not have to be met by alternators working in
parallel is
(a) terminal voltage of each machine must (c) the machines must operate at the same
be the same frequency
(b) the machines must have the same (d) the machines must have equal ratings.
phase rotation
24. After wiring up two 3-φ alternators, you checked their frequency and voltage and found them
to be equal. Before connecting them in parallel, you would
(a) check turbine speed (c) lubricate everything
(b) check phase rotation (d) check steam pressure
194
25. Zero power factor method of an alternator is used to find its
(a) efficiency (c) armature resistance
(b) voltage regulation (d) synchronous impedance
26. Some engineers prefer `lamps bright' synchronization to ‘lamps dark’ synchronization because
(a) brightness of lamps can be judged (c) flicker is more pronounced
easily (d) it can be performed quickly
(b) it gives sharper and more accurate
synchronization
27. It is never advisable to connect a stationary alternator to live bus-bars because it
(a) is likely to run as synchronous motor (d) will disturb generated e.m.fs. of other
(b) will get short-circuited alternators connected in parallel
respect to the other machine (d) its power factor would be decreased.
30. The load sharing between two steam-driven alternators operating in parallel may be adjusted
by varying the
(a) field strengths of the alternators (c) steam supply to their prime movers
31. Squirrel-cage bars placed in the rotor pole faces of an alternator help reduce hunting
(a) above synchronous speed only (b) below synchronous speed only
195
(c) above and below synchronous speeds (d) none of the above
both
196