You are on page 1of 153

ECE 320

Energy Conversion and Power Electronics


Dr. Tim Hogan

Chapter 1: Introduction and Three Phase Power

1.1 Review of Basic Circuit Analysis

Definitions:
Node - Electrical junction between two or more devices.
Loop - Closed path formed by tracing through an ordered sequence of nodes without passing through
any node more than once.

Element Constraints:
Ohms Law v = iR
dv
Capacitor Equation i=C
dt
di
Inductor Equation v=L
dt

Connection Constraints:
Kirchhoffs Current Law - The algebraic sum of currents entering a node is zero at every instant in
time.
ik = 0
Node
(1.1)

Kirchhoffs Voltage Law - The algebraic Sum of all voltages around a loop is zero at every instant in
time.
vk = 0
Loop
(1.2)

Passive Sign Convention:


Whenever the reference direction of current into a two terminal device is in the direction of the
reference voltage drop across the device, then the power absorbed (or dissipated) is positive.
+ v _
i

Figure 1. Circuit element and passive sign convention.

p (t ) = i (t ) v(t ) (1.3)

1- 1
When the above convention is used, p(t) > 0 for absorbed power, and p(t) < 0 for delivered power.

Time Varying Signals


Although a number of exceptions can be found throughout the world, the predominance of
electric power follows 60Hz or 50Hz frequencies. North America, part of Japan, and ships at sea use
60Hz while most of the rest of the world uses 50Hz. The historical reasons for these two frequencies
stem from the differences in lighting (filaments in vacuum or filaments in a gas atmosphere). The
lower frequencies caused an annoying flicker for lights having filaments in a gas atmosphere, and
thus a higher frequency was adopted in part of the world that initially used such lighting.

While not strictly adhered to within your textbook and these notes, an attempt to use the
following conventions has been made.

Scalar time varying signals (examples):


v = vmax sin (t ) (V)
v(t ) = vmax cos(t ) (V)
v = 185 sin (377 t ) (kV)
i (t ) = imax cos(100 t ) (A)

Spatial vectors (bold or arrow overhead or line overhead):


F or F or B

If necessary, unit vectors will be used, (for example): B = Bx x + By y + Bz z


These unit vectors should not be confused with phasors below

Phasors (phasor representation of a time varying signal):


A phasor can be represented as a complex number with real and imaginary components
such that a phasor of magnitude (or length) R that is at a phase angle of with respect to
the x-axis (the Real axis) can be written as R = x + jy = R e j = R cos( ) + jR sin ( )
where Eulers formula e j = cos( ) + j sin ( ) was used for the last representation. Note
that R is the magnitude and the phase of the phasor, and R = x 2 + y 2 and
y
= arctan . The phasor can be shown graphically in Figure 2.
x

1- 2
Imaginary

y
R

Real
t x

Figure 2. Phasor of magnitude R and phase .

For a sinusoidal function v(t ) = V cos(t + ) , the phasor representation is

V = Ve j = V /
If the phasor is rotating counterclockwise about the origin at a rate of radians per
second, then we multiply the phasor by e jt such that Ve jt = Ve j e jt = Ae j (t + ) and
using Eulers formula Ve jt = Ve j (t + ) = V cos(t + ) + jV sin (t + ) . Then we see that
v(t ) = V cos(t + ) is the real part of Ve jt , or

{ }
v(t ) = Re Ve jt = V cos(t + ) .
In the above description of the phasor, the peak value is used, however we will use the
RMS value of V as described in (1.6) instead of the peak value. This simplifies many of
the calculations, particularly those associated with power as shown below.

Phasors will be represented with a hat (or caret) above the variable. Impedance is understood to be a
complex quantity in general, and the hat (or caret ^) is left off the impedance variable
Z = R + jX where R is the resistance, and X is the reactance component respectively.

For a circuit element such as the one shown in Figure 1 representing a load in the circuit with i(t)
as the instantaneous value of current through the load and v(t) is the instantaneous value of the
voltage across the load.
In quasi-steady state conditions, the current and voltage are both sinusoidal, with corresponding
2
amplitudes of Vmax, Imax, and initial phases, v and i, and the same frequency = 2f =
T
v(t ) = Vmax cos(t + v ) (1.4)
i(t ) = I max cos(t + i ) (1.5)

The root-mean-squared (RMS) value of the voltage and current are then:

1 T
V = [Vmax cos(t + v )]2 dt = Vmax (1.6)
T o 2

1- 3
1 T
I= [I max cos(t + i )]2 dt = I max (1.7)
T o 2

Phasor representations for the above signals use the RMS values as V = V /v and I = I /i.

The instantaneous power is the product of voltage times current or:

p(t ) = v(t ) i (t ) = Vmax cos(t + v ) I max cos(t + i ) = Vmax I max cos(t + v ) cos(t + i )
= 2VI cos(t + v ) cos(t + i )
1
= 2VI [cos(v i ) + cos(2t + v + i )] (1.8)
2

The average value is found by integrating over a period of time and then dividing the result by
that same time interval. The first term in (1.8) is independent of time, while the second term varies
from -1 to +1 symmetrically about zero. Because it is symmetric about zero, integration over an
integer number of cycles (or periods) gives a value of zero for the second term in (1.8), and the
average power which we define as the real power with units of watts (W) is:
P = VI cos(v i ) (1.9)
This shows the power is not only proportional to the RMS values of voltage and current, but also
proportional to cos(v i ) . The cosine of this angle is defined as the displacement factor, DF. In
more general terms for periodic, but not necessarily sinusoidal signals, the power factor is defined as:
P
pf (1.10)
VI
For sinusoidal signals, the power factor equals the displacement factor, or
pf = cos(v i ) (1.11)
For comparison, the voltage, current, and power for various angles between voltage and current
are shown below:

For leading power factors:


v(t ) = 5 cos(t )
Voltage (V), Current (A), Power (W)

10
i (t ) = 2 cos(t )
5
Vmax I max
P = VI cos(v i ) = cos(v i ) = 5 (W)
0
2 2
Imaginary

-5

-10
I
0 0.5 1 1.5 2 2.5 3 Real
# of Periods V

1- 4
v(t ) = 5 cos(t )
Voltage (V), Current (A), Power (W)
10
i (t ) = 2 cos(t + 30 )
5
V I
P = VI cos(v i ) = max max cos(v i ) = 4.33 (W)
0
2 2
Imaginary

-5

-10 I
0 0.5 1 1.5 2 2.5 3 Real
# of Periods V

v(t ) = 5 cos(t )
Voltage (V), Current (A), Power (W)

10
i (t ) = 2 cos(t + 60 )
5
V I
P = VI cos(v i ) = max max cos(v i ) = 2.5 (W)
0
2 2
Imaginary

-5

I
-10
0 0.5 1 1.5 2 2.5 3 Real
# of Periods V

v(t ) = 5 cos(t )
Voltage (V), Current (A), Power (W)

10
i (t ) = 2 cos(t + 90 )
5
V I
P = VI cos(v i ) = max max cos(v i ) = 0 (W)
0
2 2
Imaginary

-5

-10 I
0 0.5 1 1.5 2 2.5 3 Real
# of Periods V

v(t ) = 5 cos(t )
Voltage (V), Current (A), Power (W)

10
i (t ) = 2 cos(t + 120 )
5
V I
P = VI cos(v i ) = max max cos(v i ) = 2.5 (W)
0
2 2
Imaginary

-5
I
-10
0 0.5 1 1.5 2 2.5 3 Real
# of Periods V

1- 5
v(t ) = 5 cos(t )
Voltage (V), Current (A), Power (W)
10
i (t ) = 2 cos(t + 150 )
5
V I
P = VI cos(v i ) = max max cos(v i ) = 4.33 (W)
0
2 2
Imaginary

-5

I
-10
0 0.5 1 1.5 2 2.5 3 Real
# of Periods V

v(t ) = 5 cos(t )
Voltage (V), Current (A), Power (W)

10
i (t ) = 2 cos(t + 180 )
5
V I
P = VI cos(v i ) = max max cos(v i ) = 5 (W)
0
2 2
Imaginary

-5

-10
I
0 0.5 1 1.5 2 2.5 3 Real
# of Periods V

For lagging power factors:


v(t ) = 5 cos(t )
Voltage (V), Current (A), Power (W)

10
i (t ) = 2 cos(t )
5
Vmax I max
P = VI cos(v i ) = cos(v i ) = 5 (W)
0
2 2
Imaginary

-5

-10
I
0 0.5 1 1.5 2 2.5 3 Real
# of Periods V

1- 6
v(t ) = 5 cos(t )
Voltage (V), Current (A), Power (W)
10
i (t ) = 2 cos(t 30 )
5
V I
P = VI cos(v i ) = max max cos(v i ) = 4.33 (W)
0
2 2
Imaginary

-5

-10
0 0.5 1 1.5 2 2.5 3 Real
# of Periods V
I

v(t ) = 5 cos(t )
Voltage (V), Current (A), Power (W)

10
i (t ) = 2 cos(t 60 )
5
V I
P = VI cos(v i ) = max max cos(v i ) = 2.5 (W)
0
2 2
Imaginary

-5

-10
0 0.5 1 1.5 2 2.5 3 Real
# of Periods V
I
v(t ) = 5 cos(t )
Voltage (V), Current (A), Power (W)

10
i (t ) = 2 cos(t 90 )
5
V I
P = VI cos(v i ) = max max cos(v i ) = 0 (W)
0
2 2
Imaginary

-5

-10
0 0.5 1 1.5 2 2.5 3 Real
# of Periods V
I

v(t ) = 5 cos(t )
Voltage (V), Current (A), Power (W)

10
i (t ) = 2 cos(t 120 )
5
V I
P = VI cos(v i ) = max max cos(v i ) = 2.5 (W)
0
2 2
Imaginary

-5

-10
0 0.5 1 1.5 2 2.5 3 Real
# of Periods V
I

1- 7
v(t ) = 5 cos(t )
Voltage (V), Current (A), Power (W)
10
i (t ) = 2 cos(t 150 )
5
V I
P = VI cos(v i ) = max max cos(v i ) = 4.33 (W)
0
2 2
Imaginary

-5

-10
0 0.5 1 1.5 2 2.5 3 Real
# of Periods V
I

v(t ) = 5 cos(t )
Voltage (V), Current (A), Power (W)

10
i (t ) = 2 cos(t 180 )
5
V I
P = VI cos(v i ) = max max cos(v i ) = 5 (W)
0
2 2
Imaginary

-5

-10
0 0.5 1 1.5 2 2.5 3 Real
# of Periods I V

We call the power factor leading or lagging, depending on whether the current of the load leads or
lags the voltage across it. We know that current does not change instantly through an inductor, so it
is clear that for a resistive - inductive (RL) load, the power factor is lagging. Likewise the voltage
can not change instantly across a capacitor, therefore for a resistive capacitive (RC) load, the power
factor is leading (the current changes before the voltage can). Also, for a purely inductive or
capacitive load the power factor is 0, while for a purely resistive load it is 1.
The product of the RMS values of voltage and current at a load is the apparent power, S having
units of volt-amperes (VA):
S = VI (1.12)

The reactive power is Q with units of volt-amperes reactive (VA reactive, or VAr):
Q = VI sin (v i ) (1.13)

The reactive power represents the energy oscillating in and out of an inductor or capacitor. Since
the energy oscillation in an inductor is 180 out of phase with the energy oscillating in an capacitor,
the reactive power of these two have opposite signs with the convention that it is positive for the
inductor and negative for the a capacitor.
Using phasors, the complex apparent power, S is:
S = VI* = V/v I/-i (1.14)
or

1- 8
S = P + jQ (1.15)

As an example, consider the following voltage and current for a given load:


v(t ) = 120 2 sin 377t + (V) (1.16)
6


i (t ) = 5 2 sin 377t + (A) (1.17)
4

then S = VI = 120 5 = 600 (W), while the power factor is pf = cos = 0.966 leading. Also,
6 4
the complex apparent power is:

S = VI* = 120//6 5/-/4 = 600/- /12 = 579.6 (W) j155.3 (VAr) (1.18)

It should also be noted that when the angles are represented in radian, care must be taken to
assure your calculator is in radians mode. Often we represent the argument of the sine or cosine term
in mixed units. For example we might write cos(377t + 30). The first term (377t) has units of
radians, while the second (30) has units of degrees, and one of these must be converted before
making calculations with your calculator or computer either in radians mode or in degrees mode.
As a phasor the complex apparent power can be shown on a complex plane for the cases of
leading and lagging power factors as shown in Figure 3.

Imaginary Imaginary

Real V Real
V
I

Imaginary Imaginary

S
P Q
Real Real
Q P
S

(a) Leading power factor (b) Lagging power factor


Figure 3. Phasor of magnitude R and phase .
1- 9
Recall that the leading power factor corresponds to a resistive-capacitive load. We see from
Figure 3(a), this also corresponds to a negative value for Q. Similarly, lagging power factor has the
current lagging the voltage corresponding to an inductive-resistive load, and Figure 3(b) shows this
corresponds to a positive value for Q.
To summarize, we can use the following tables:

Equation Units
Real Power P = VI cos(v i ) Watts
= VI pf
= S pf
Reactive Power Q = VI sin (v i ) Volt-Amperes-Reactive
= S 1 pf 2 (for lagging)
= S 1 pf 2 (for leading)
Apparent Power S = VI Volt-Amperes
S = VI* = V/v I/-i
S = P + jQ
S 2 = P2 + Q2

Type of Load Reactive Power Power Factor


Inductive Q>0 lagging
Capacitive Q<0 leading
Resistive Q=0 1

From the interdependence of the four quantities, S, P, Q, pf, if we know any two of these
quantities, the other two can be determined. For example, if S = 100 (kVA), and pf = 0.8 leading,
then:
P = S pf = 800 (kW)
Q = S 1 pf 2 = 60 (kVAR), or
sin (v i ) = sin[arccos(0.8)] then
Q = S sin (v i )

It is important to notice that Q < 0, such that sin (v i ) is a negative quantity. This can be seen
when it is understood that there are two possible answers for the arccos(0.8), that is cos(36.87) = 0.8,
and cos(-36.87) = 0.8 so to obtain a Q < 0 we use (v i) = 36.87.
Generally, in systems that contain more than one load (or source), the real and reactive power can
be found by adding individual contributions, but this is not the case with the apparent power. That is

Ptotal = i Pi
Qtotal = i Qi (1.19)
Stotal i Si

1- 10
For the above example, if the load voltage is VL = 2000 (V), then the load current would be
IL = S/VL = [100103 (VA)]/[2103 (V)] = 50 (A). If we use the load voltage as the reference, then:
V = 2000 /0 (V)
I = 50 /i = 50 /+36.87 (A)
S = VI* = [2000 /0][50 /36.87] = P + jQ = 80103 (W) j60103 (VAr)

1.2 Three Phase Balanced Systems

Three-phase systems offer significant advantages over single phase systems: for the same power
and voltage there is less copper in the windings, and the total power absorbed remains constant rather
than oscillating about an average value.
For a three phase system consisting of three current sources having the same amplitude and
frequency, but with phases differing by 120 as:

i1 (t ) = 2 I sin (t + )
2
i2 (t ) = 2 I sin t + (1.20)
3
2
i3 (t ) = 2 I sin t + +
3

If these are connected as shown in Figure 4, then at node n or n, the current adds to zero, and the
neutral line n-n (dashed) is not needed.

2 2
i1 (t ) + i2 (t ) + i3 (t ) = 2 I sin (t + ) + sin t + + sin t + + = 0
3 3

i1

n' n

i3 i2

Figure 4. Balanced three phase Y-connected system with zero neutral current.

If instead we had three voltage sources Y-connected as in Figure 5 with the following values

1- 11
v1 (t ) = 2V sin (t + )
2
v2 (t ) = 2V sin t + (1.21)
3
2
v3 (t ) = 2V sin t + +
3
then, the current through each of the three loads (assuming the loads are equal), would have equal
magnitudes, but each current would have a phase that is shifted by an equal amount with respect to
the voltages, v1(t), v2(t), v3(t).
ia

+ v1

n' n
ib
+

v3 v2
ic

Figure 5. Balanced voltage fed three phase Y-connected system with zero neutral current.

With equal impedances for the loads, then

2V
ia (t ) = sin (t + + )
Z
2V 2
ib (t ) = sin t + + (1.22)
Z 3
2V 2
ic (t ) = sin t + + +
Z 3

and again the currents sum to zero at nodes n or n, and for this balanced three phase system, the
neutral wire (dashed) is not required.
In comparison to a single phase system, where two wires are required per phase, the three phase
system delivers three times the power, and requires only three transmission wires total. This is a
significant advantage considering the hundreds of miles of wire needed for power transmission.

Y and Connections

The loads in the previous two figures, as well as in Figure 6 are connected in a Y or star
configuration. If the load of Figure 6 is for a balanced Y system, then the voltages between each
phase and the neutral are: V1n = V / , V2 n = V / 2/3 , and V3n = V / + 2/3 .
Kirchhoffs voltage law (KVL) states that the sum of voltages around a closed loop equals zero.
This is also the case here however the voltages are complex numbers or phasors, and as such must be
1- 12
added as vectors. The phase can be any value, but the relative position of the phase to neutral
phasors must be 120 with respect to each other as shown in Figure 7.

1 V3n +
+ +
V1n
n
V12
V2n
+
2

Figure 6. Y-connected loads with voltages relative to neutral identified.

By KVL: V12 + V1n V2 n = 0 , or V12 = V1n V2 n as shown in Figure 7.

-V1n
V31
V3n
V2n V12
V1n
-V3n -V2n
V23

Figure 7. Voltage phasors of the Y-connected loads shown in Figure 6.

We could also use the phasor representation V1n = V / , V2 n = V / 2/3 ,


and V = V / + 2/3 to determine the line-to-line voltages as
3n

V12 = V1n V2 n = 3V / + /6 (1.23)

This shows the RMS value of the line-to-line voltage, Vl-l , at a Y load is 3 times the line-to-
neutral or phase voltage, Vln. In the Y connection, the phase current is equal to the line current, and
the power supplied to the system is three times the power supplied to each phase, since the voltage
and current amplitudes and phase differences between them are the same in all three phases. If the
power factor in one phase is pf = cos(v i ) , then the total power to the system is:

1- 13
S3 = P3 + jQ3
= 3V1n I1*n (1.24)
= 3Vl l I l cos(v i ) + j 3Vl l I l sin (v i )

Similarly, for a connection of the loads in the configuration (as in Figure 8), the phase voltage
is equal to the line voltage however; the phase currents are not equal to the line currents for the
configuration. If the phase currents are I12 = I / , I23 = I / 2/3 , and I31 = I / + 2/3 then
using Kirchhoffs current law (KCL) the current of line 1, as shown in Figure 8 is:

I1 = I12 I31 = 3I / - /6 (1.25)

Thus for the configuration, the line current is 3 times the I current.

I1
1
-I23
I12 I2 I31
-I12 I3
I31 I23
I12
I2
I1
2
I23 -I31
3
I3

Figure 8. connected load, line, and phase currents.

To calculate the power in the three-phase connected load:

S3 = P3 + jQ3
= 3V12 I12* (1.26)
= 3Vl l I l cos(v i ) + j 3Vl l I l sin (v i )

which is the same value as for the Y connected load.


For a balanced system, the loads of the three phases are equal. Also, a configured load, can be
replaced with a Y configured load (and visa versa) if:

1
ZY = Z (1.27)
3
1- 14
Under these conditions, the two loads are indistinguishable by the power transmission lines.
You might recall the to Y transformation for resistor circuits can be remembered by overlaying
the and Y configurations such as in Figure 9.
A

RB R1 RC
R3
R2
C B
RA

Figure 9. to Y or Y to resistor network transformation.

R B RC
R1 =
R A + R B + RC
(1.28)
R1 R 2 + R 2 R 3 + R 3 R1
RA =
R1

Equation (1.28) gives a similar result to that of (1.27) when RA = RB = RC and R1 = R2 = R3.
B

1.3 Calculations in Three-Phase Systems

Calculations of quantities like currents, voltages, and power in three-phase systems can be
simplified by the following procedure:

1. transform the circuits to Y,


2. connect a neutral conductor,
3. solve one of the three 1-phase systems
4. convert the results back to the systems

1.3.1 Example
For the 3-phase system in Figure 10 calculate the line-to-line voltage, real power, and power
factor at the load.
To solve this by the procedure outlined above, first consider only one phase as shown in Figure
11.

1- 15
j1 ()

+ v1 = 120 (V) 7 + j5 ()

n' n

+
v3 v2

Figure 10. Three phase system with Y connected load, and line impedance.

j1 () I

+ v1 = 120 (V) 7 + j5 ()

n' n
Figure 11. One phase of the three phase system shown in Figure 10.

For the one-phase in Figure 11,

120
I = = 13.02 /-40.6 (A)
j1 + (7 + j 5)
V1n = IZ L = 13.02 /-40.6 (7+j5) = 112/-5.1 (V)
S = V I* = (112/-5.1)(13.02/40.6) = 1458.3/35.5 = 1.187103 + j0.848103
L ,1 L

PL ,1 = 1.187 (kW), QL ,1 = 0.848 (kVAr)


pf = cos(-5.1 - (-40.6)) = 0.814 lagging

For the three-phase system of Figure 10 the load voltage (line-to-line), the real, and reactive
power are:

VL ,l l = 3 112 = 194 (V)


PL ,3 = 3.56 (kW)
QL ,3 = 2.544 (kVAr)

1- 16
1.3.2 Example
For the Y to three-phase system in Figure 12, calculate the power factor and the real power at
the load, as well as the phase voltage and current. The source voltage is 400 (V) line-to-line.

j1 ()

+ v1
18 + j6 ()

n'
+

+
v3 v2

Figure 12. connected load.

First convert the load to an equivalent Y connected load, then work with one phase of the system.
400
The line to neutral voltage of the source is Vln = = 231 (V).
3

j1 ()

+ 231 (V) 6 + j2 ()

n' n
+

Figure 13. Equivalent Y connected load.

IL
j1 ()

+ v1 = 231 (V) 6 + j2 ()

n' n

Figure 14. One phase of the Y connected load.

1- 17
231
IL = = 34.44 /-26.6 (A)
j1 + (6 + j 2)

VL = IL (6 + j 2) = 217.8 /-8.1 (V)

The power factor at the load is:

pf = cos(v i ) = cos( 8.1 +26.6 ) = 0.948 lagging

Converting back to a connected load gives:

I L 34.44
I = = = 19.88 (A)
3 3

Vl l = 217.8 3 = 377.22 (V)

At the load the power is:

PL ,3 = 3Vl l I L pf = 3 377.22 34.44 0.948 = 21.34 (kW)

1.3.3 Example
Two three-phase loads are connected as shown in . Load 1 draws from the system PL1 = 500
(kW) at 0.8 pf lagging, while the total load is ST = 1000 (kVA) at 0.95 pf lagging. What is the power
factor of load 2?

Load 1
Power System

Load 2

Figure 15. Two three-phase loads connected to the same power source.

1- 18
For the total load we can add the real and reactive power for each of the two loads (we can not
add the apparent power).

PT = PL1 + PL 2
QT = QL1 + QL 2
ST S L1 + S L 2

From the information we have for the total load we can write the following:

PT = ST pfT = 950 (kW)


[ ]
QT = ST sin cos 1 (0.95) = 312.25 (kVAr)

The reactive power, QT, is positive since the power factor is lagging.
For the load L1, PL1 = 500 (kW), pf1 = 0.8 lagging, thus:

500 103
S L1 = = 625 (kVA)
0.8
QL1 = S L21 PL21 = 375 (kVA)

Again, QL1 is positive since the power factor is lagging. This leads to:

PL 2 = PT PL1 = 450 (kW)


QL 2 = QT QL1 = -62.75 (kVAr)

and
PL 2 450
pf L 2 = = = 0.99 leading.
SL2 4502 + 62.752

Chapter Notes
A sinusoidal signal can be described uniquely by:
1. Time dependent form as for example: v(t ) = 5 sin (2ft + v ) (V)
2. by a time dependent graph of the signal
3. as a phasor along with the associated frequency of the phasor
one of these descriptions is enough to produce the other two.
It is the phase difference that is important in power calculations, not the phase. The phase is
arbitrary depending on the defined time (t = 0). We need the phase to solve circuit problems
after we take one quantity (some voltage or current) as a reference. For that reference quantity
we assign an arbitrary phase (often zero).
In both three-phase and one-phase systems the total real power is the sum of the real power
from the individual loads. Likewise the total reactive power is the sum of the reactive power of
the individual loads. This is not the case for the apparent power or the power factor.

1- 19
Of the four quantities: real power, reactive power, apparent power, and power factor, any two
describe a load adequately. The other two quantities can be calculated from the two given.
To calculate real, reactive, and apparent power when using equations (1.9), (1.12), and (1.13)
we must use absolute values, not complex values for the currents and voltages. To calculate the
complex power using equation (1.14) we do use complex currents and voltages and find
directly both the real and reactive power (as the real and imaginary components respectively).
When solving a circuit to calculate currents and voltages, use complex impedances, currents
and voltages.

1- 20
ECE 320
Energy Conversion and Power Electronics
Dr. Tim Hogan

Chapter 2: Magnetic Circuits and Materials

Chapter Objectives
In this chapter you will learn the following:
How Maxwells equations can be simplified to solve simple practical magnetic problems
The concepts of saturation and hysteresis of magnetic materials
The characteristics of permanent magnets and how they can be used to solve simple problems
How Faradays law can be used in simple windings and magnetic circuits
Power loss mechanisms in magnetic materials
How force and torque is developed in magnetic fields

2.1 Amperes Law and Magnetic Quantities


Amperes experiment is illustrated in Figure 1 where there is a force on a small current element
I2l when it is placed a distance, r, from a very long conductor carrying current I1 and that force is
quantified as:
I
F = 1 I 2l (N) (2.1)
2 r

I1
I2
r F

Conductor Current Element


1 of Length l

Figure 1. Amperes experiment of forces between current carrying wires.

The magnetic flux density, B, is defined as the first portion of equation (2.1) such that:

F = BI 2l (N) (2.2)

1- 1
From (2.1) and (2.2) we see the magnetic flux density around conductor 1 is proportional to the
current through conductor 1, I1, and inversely proportional to the distance from conductor 1. Looking
N
at the units and constants handout given in class, or from (2.2) the units of, B, are seen as ,
Am
N
thus , called permeability, has units of 2 . More commonly, the relative permeability of a given
A
N
material is given where = r 0 and 0 = 400 109 2 . Since a Newton-meter is a Joule, and
A
N N m J V A s V s
a Joule is a Watt-second: = 2
= 2
= 2
= 2 . This shows B is a
Am Am Am Am m
per meter squared quantity, and the (Vs) units represents the magnetic flux and is given units of
Webers (Wb). This flux can be found by integrating the normal component of B over the area of a
given surface:


= B nds
S
(2.3)

The magnetic field intensity is related to the magnetic flux density by the permeability of the
media in which the magnetic flux exists.
B
H (2.4)

B I1 A
For the system in Figure 1, H = = and have units of . If there were multiple
2r m
conductors in place of conductor 1, for example in a coil, then the units would be ampere-turns per
meter. A line integration of H over a closed circular path gives the current enclosed by that path, or
for the system in Figure 1:
2r
I1
C
H = H dl =
0
2r
dl = I1 (2.5)
again, if the system contained multiple conductors within the enclosed path, the result would give
ampere-turns. Equation (2.5) is Amperes circuital law.

An alternative approach as described in your textbook is to begin with Maxwells equations


which include Amperes circuital law in a more general form as shown in Table I below:

Table I. Maxwells equations.


Name Point Form Integral Form
B d
Faradays Law E =
t C
E dl =
dt B n ds
S

D D
Amperes Law Modified by Maxwell H = J +
t H dl = J + t n ds
C S

Gausss Law D = D n ds = dv
S V

Gausss Law for Magnetism B = 0 B n ds = 0


S

1- 2
where
C
represents the integral over a closed path,
S
represents the integral over the surface of a

closed volume of space, n is a surface normal vector, E is the spatial vector of electric field, B is the
spatial vector of magnetic flux density, J is the spatial vector of the electric current density, D is
the displacement charge vector, is the electric charge density,
V
is a volume integral is the
curl and the divergence of the vector being acted upon.

Some assumptions commonly used in electromechanical energy conversion include a low enough
D
frequency, that the displacement current, , can be neglected, and the assumption of homogeneous
t
and isotropic media used in the magnetic circuit. Under these assumptions, Amperes circuital law is
modified to remove the displacement current component such that

H dl =
C S
J n ds (2.6)

which for the system of Figure 1 reduces to equation (2.5).

2.2 Magnetic Circuits


From Amperes circuital law,
H dl =
C S
J n ds , we see the magnetic field intensity around a
closed contour is a result of the total electric current density passing through any surface linking that
contour. Gausss Law for magnetism,
B n ds = 0 states that there are no magnetic monopoles
S
that is to say there for a closed surface there is as much magnetic field density leaving that closed
surface as there is entering the closed surface. If the integration is for an area, but not a closed
surface area, then we obtain the flux or =
B nds .
S
The permeability of free space is 1, while the permeability of magnetic steel is a few hundred
thousand. Magnetic flux can be confined to the structures or paths formed by high permeability
materials. In this way, magnetic circuits can be formed such as the one shown in Figure 2.

Figure 2. Simple magnetic circuit [1].

1- 3
The driving force for the magnetic field is the magnetomotive force (mmf), F, which equals the
ampere-turn product
F = N i (2.7)

The analysis of a magnetic circuit is similar to the analysis of an electric circuit, and an analogy
can be made for the individual variables as shown in Table II.

Table II. Comparison of electric and magnetic circuits.


Electric Circuit Magnetic Circuit
Electrically Conductive High Permeability
I Material Material
b b

a a

V r V r
I I

c c

Driving Force
applied battery voltage = V applied ampere-turns = F

Response
driving force driving force
current = flux =
electric resistance magnetic reluctance
or or
V F
I= =
R R

Impedance
Impedance is used to indicate the impediment to the driving force in establishing a response.
l l
resistance = R = reluctance = R =
A A

where l = 2r, = electrical conductivity, A = where l = 2r, = permeability, A = cross-


cross-sectional area sectional area

Equivalent Circuit
I

V R F R

V = IR F = R

1- 4
Electric Circuit Magnetic Circuit

Fields
Electric Field Intensity Magnetic Field Intensity
V V F F
E = (V/m) H = (A-t/m)
l 2r l 2r
or or

E dl = V H dl = F
Potential
Electric Potential Difference Magnetic Potential Difference
b V b IR I l b F R l
a
Vab = E dl =
l a
dl = lab =
l l A
lab = IRab
a
Fab = H dl = lab =
l l
lab =
l A
lab = Rab

Flow Densities
Current Density Flux Density
I V El F Hl
= E = H
J =
A AR A l
=
A( ) B = =
A AR A l
A

An example magnetic circuit is shown in Figure 3, below.

Figure 3. Simple magnetic circuit with an air gap [1].

For this circuit, we will assume:


the magnetic flux density is uniform throughout the magnetic cores cross-sectional area and is
perpendicular to the cross-sectional area
the magnetic flux remains within the core and the air gap defined by the cross-sectional area of the
core and the length of the gap (no leaking of field, no fringe fields at the gap).

Then


= B dA = Bc Ac = Bg Ag
Ac
(2.8)

with Ac = Ag
1- 5
Bc = Bg
(2.9)
c H = g H

Since, F = Hl
F = H c lc + H g g
Bc Bg (2.10)
= lc + lg
c g
or

l g
F = c +
A A (2.11)
c c g g
= (Rc + Rg )

Thus the magnetic circuit shown in Figure 3 can be represented as

Rc
F N i
= =
F Rc + Rg Rc + Rg

Rg

Figure 4. Equivalent circuit for the magnetic circuit in Figure 3.

This concept is helpful for more complex configurations such as shown in Figure 5.

1- 6
Area A2
Area
A1
Area
I1 A1 I2
Area
A1 Area
A3

g1 g2

Area
A1
Area A2

Figure 5. Magnetic circuit with various cross-sectional areas and two coils.

Using the lengths defined in Figure 6, and paying attention to the direction of the magnetomotive
force from each of the coils by use of the right hand rule, the equivalent circuit can be drawn as seen
in Figure 7.

Area A1 l1 Area A2

I1 I2

Area
A3
Area A1
g1 g2

1
2
l 3

1
2
l 2

Area A1

Figure 6. Lengths for each section of the magnetic circuit of Figure 5.

Then,
F1 = N1 I1 = Rl 1 (1 + 2 ) + Rg 11 (2.12)
and
(
F2 = N 2 I 2 = 1Rg1 2 Rl3 + Rg2 + Rl2 ) (2.13)

For a system with the dimensions, number of turns, current, and permeability known, then
equations (2.12) and (2.13) give two equations with two unknowns such that 1 and 2 can be found.

1- 7
2
Rl F2
1 1
Rl
3

F1 Rg
1
Rg
2
Rl
2

Figure 7. Equivalent circuit for the configuration of Figure 5.

Assuming the gaps are air gaps, the value of each reluctance can be found as:

l1 g1 l2 l3 g2
Rl 1 = Rg 1 = Rl2 = Rl3 = Rg2 =
A1 0 A1 A2 A3 0 A2

Then we can find the magnetic flux density for each gap as:

1
Bg1 =
A1
(2.14)
2
Bg 2 =
A2

Note that the permeability of the core material can be much larger than the permeability of the gaps
such that the total reluctance is dominated by the reluctances of the gaps.

2.3 Inductance
From circuit theory we recall that the voltage across an inductor is proportional to the time rate of
change of the current through the inductor.
di (t )
v L (t ) = L L (2.15)
dt
and while the power of an inductor can be positive or negative, the energy is always positive as
1
wL (t ) = L iL2 (2.16)
2
d
In Table I, Faradays law is
E dl = dt B n ds or the electric field intensity around a
C S
closed contour C is equal to the time rate of change of the magnetic flux linking that contour.
Integrating over the closed contour of the coil itself gives us the negative of the voltage at the
terminals of the coil. On the right side of Faradays law we then must integrate over the surface of
the full coil, thus including the N turns of the coil. Then Faradays law gives
1- 8
d (t )
vL (t ) = N (2.17)
dt

Comparing equation (2.15) and (2.17) gives


d
L=N (2.18)
di

For linear inductors, the flux is directly proportional to current, i, for all values such that
N
L= (2.19)
i
with units of (Weber-turns per ampere), or Henrys (H). The flux linkage, , is defined as = N
F Ni
and combining this with the relationship between flux and total circuit reluctance = =
Rtot Rtot
along with (2.19) gives
N2
L= (2.20)
Rtot
Thus inductance can be increased by increasing the number of turns, using a metal core with a higher
permeability, reducing the length of the metal core, and by increasing the cross-sectional area of the
metal core. This is information that is not readily seen by either the circuit laws for inductors, or
through Faradays equation alone.

Mutual Inductance
For a magnetic circuit containing two coils and an air gap such as in Figure 8, with each coil
wound such that the flux is additive, then the total magnetomotive force is given by the sum of
contributions from the two coils as
F = N1i1 + N 2i2 (2.21)

Area Ac lc

i1 i2
g

1 2

Figure 8. Mutual inductance magnetic circuit.

The equivalent circuit is shown in Figure 9.


1- 9

Rg

F1 F2

R lc

Figure 9. Equivalent circuit for Figure 8.

If the permeability of the core is large such that Rlc << Rg and the cross-sectional area of the gap
is assumed equal to the cross-sectional area of the core (Ag = Ac), then

0 Ac
(N1i1 + N 2i2 ) (2.22)
g
The flux linkage, 1, in Figure 8 is 1 = N1 , or

0 Ac A
1 = N1 = N12 i1 + N1N 2 0 c i2
g g (2.23)
= L11i1 + L12i2

where
A
L11 = N12 0 c (2.24)
g
is the self-inductance of coil 1 and L11i1 is the flux linkage of coil 1 due to its own current i1. The
mutual inductance between coils 1 and 2 is
A
L12 = N1N 2 0 c (2.25)
g
and L12i2 is the flux linkage of coil 1 due to current i2 in the other coil. Similarly, for coil 2
0 Ac A
2 = N 2 = N1N 2 i1 + N 22 0 c i2
g g (2.26)
= L21i1 + L22i2
where L21 = L12 is the mutual inductance and L22 is the self-inductance of coil 2.
A
L22 = N 22 0 c (2.27)
g

1- 10
2.4 Magnetic Material Properties
A simplification we have used is that the permeability of a given material is constant for different
applied magnetic fields. This is true for air, but not for magnetic materials. Materials that have a
relatively large permeability are ferromagnetic materials in which the magnetic moments of the atoms
can align in the same direction within domains of the material when and external field is applied. As
more of these domains align, saturation is reached when there is no further increase in flux density of
that of free space for further increases in the magnetizing force. This leads to a changing
permeability of the material and a nonlinear B vs. H relationship as shown in Figure 10.

Figure 10. Nonlinear B vs H normal magnetization curve.

When the field intensity is increased to some value and is then decreased, it does not follow the
curve shown in Figure 10, but exhibits hysteresis as shown by the abcdea loop in Figure 11. The
deviation from the normal magnetization curve is caused by some of the domains remaining oriented
in the direction of the originally applied field. The value of B that remains after the field intensity H
is removed is called residual flux density. Its value varies with the extent to which the material is
magnetized. The maximum possible value of the residual flux density is called retentivity and results
whenever values of H are used that cause complete saturation. When the applied magnetic field is
cyclically applied so as to form the hysteresis loop such as abcdea in Figure 11, the field intensity
required to reduce the residual flux density to zero is called the coercive force. The maximum value
of the coercive force is called the coercivity.
The delayed reorientation of the domains leads to the hysteresis loops. The units of BH is

amperes newtons N J
units of HB = = 2 = 3 (2.28)
meter ampere - meter m m
1- 11
or an energy density.

B (Wb/m2)

Coercivity
f
a

b
Retentivity
Residual
flux
c H (At/m)
O

Coercive force

Figure 11. Hysteresis loops. The normal magnetization curve is in bold.

a
f

c
O H

Figure 12. Energy relationship for hysteresis loop per half-cycle.

1- 12
The full shaded area in Figure 12 outlined by eafe represents the energy stored in the magnetic field
during the positive half cycle of H. The hatched area eabe represents the hysteresis loss per half
cycle. This energy is what is required to move around the magnetic dipoles and is dissipated as heat.
The energy released by the magnetic field during the positive half cycle of they hysteresis loop is
given by the cross-hatched area outlined by bafb and is energy that is returned to the source.
The power loss due to hysteresis is given by the area of the hysteresis loop times the volume of
the ferromagnetic material times the frequency of variation of H. This power loss is empirically
given as
n
Ph = khfBmax (2.29)

where n lies in the range 1.5 n 2.5 depending on the material used, is the volume of the
ferromagnetic material, and the value of the constant, kh, also depends on the material used. Some
typical values for kh are: cast steel 0.025, silicon sheet steel 0.001, and permalloy 0.0001.
In addition to the hysteresis power loss, eddy current losses also exist for time-varying magnetic
fluxes. Circulating currents within the ferromagnetic material follow from the induced voltages
described by Faradays law. To reduce these eddy current losses, thin laminations (typically 14-25
mils thick) are commonly used where the magnetic material is composed of stacked layers with an
insulating varnish or oxide between the thin layers. An empirical equation for the eddy-current loss
is
Pe = ke 2 f 2 Bmax
2
(2.30)
where ke = constant dependent on the material
f = frequency of the variation of flux
Bmax = maximum flux density
B

= lamination thickness
= total volume of the material

The total magnetic core loss is the sum of the hysteresis and eddy current losses.
If the value of H, when increasing towards some maximum, Hmax, does not increase continuously,
but at some point, H1, decreases to H = 0 then increases again to it maximum value of Hmax, then a
minor hysteresis loop is created as shown in Figure 13. The energy loss in one cycle includes these
additional minor loop surfaces.

1- 13
B

Hmax

O H1 H

Figure 13. Minor hysteresis loops.

2.5 Permanent Magnets


For a ring of iron with a uniform cross-section and hysteresis curve shown in Figure 15, the
magnetic field is zero when there is a nonzero flux density, Br called the remnant flux density. To
B

achieve a zero flux density, we could wind a coil around a section of the iron, and send current
through the coil to reach a field intensity of Hc (the coercive field). In practice a permanent magnet
operates on a minor loop as shown in Figure 13 that can be approximated as a straight line, recoil
line, such that
B
Bm = r H m + Br (2.31)
Hc
Magnetization curves for some important permanent-magnet materials is shown in figure 1.19
from your textbook and shown below in Figure 14.

Example
In the magnetic circuit shown with the length of the magnet,
lm = 1cm, the length of the air gap is g = 1mm and the length of the li
iron is li = 20cm. For the magnet Br = 1.1 (T), Hc = 750 (kA/m), what
B

is the flux density in the air gap if the iron is assumed to have an
infinite permeability and the cross-section is uniform? lm g
Since the cross-section is uniform, and there is no current:
Hi[0.2(m)] + Hg[g] + Hm[lm]=0
With the iron assumed to have infinite permeability, Hi = 0 and li

1- 14
Bg H B
H g g + H m lm = 0 = g + c m H c lm = 0
0 Br
g H
Bg + (Bm 1.1) c lm = 0
0 Br
or B795.77 + (B-1.1)6818=0
B=0.985 (T)

Figure 14. Magnetization curves for common permanent-magnet materials [1].

1- 15
B (Wb/m2)

Br

(Hm, Bm)

-Hc
H (At/m)
O

Figure 15. Remnant flux and coercive field for a piece of iron.

2.6 Torque and Force


In Figure 12 we found the energy density in the field for the positive half-cycle is the total area or

Ba
wf =
B e
H dB (2.32)

If this is simply approximated as a triangular area, then wf = BH (J/m3). The total energy, Wf,
would be found by multiplying this by the volume of the core

1
Wf = (BH )( Al ) = 1 (BA)Hl = 1 F
2 2 2
(2.33)
1
= 2R
2

Mechanical energy is done by reducing the reluctance (as the armatures of a relay are brought
together, for example), thus
1
dWm = 2dR (2.34)
2

Since force is given as Fdx = dWm , the magnetic force is given by

1- 16
1 dR
F = 2 (2.35)
2 dx

and has units of newtons.

In a mechanical system with a force F acting on a body and moving it at velocity v, the power
Pmech is
Pmech = F v (2.36)

For a rotating system with torque T, rotating a body with angular velocity mech:
Pmech = T mech (2.37)
On the other hand, an electrical source e, supplying current, i, to a load provides electrical power
Pelec
Pelec = ei (2.38)
Since power has to balance, if there is no change in the field energy,
Pelec = Pmech = ei = T mech (2.39)

Notes
It is more reasonable to solve magnetic circuits starting from the integral form of Maxwells
equations than finding equivalent resistance, voltage and current. This also makes it easier to use
saturation curves and permanent magnets.
Permanent magnets do not have flux density equal to Br. Equation (2.31) defines the relation
B

between the variables, flux density Bm and field intensity Hm in a permanent magnet.
B

There are two types of iron losses: eddy current losses that are proportional to the square of the
frequency and the square of the flux density, and hysteresis losses that are proportional to the
frequency and to some power n of the flux density.

1 A. E. Fitzgerald, C. Kingsley, Jr., S. D. Umans, Electric Machinery, 6th edition, McGraw-Hill, New York, 2003.

1- 17
ECE 320
Energy Conversion and Power Electronics
Dr. Tim Hogan

Chapter 3: Transformers (Textbook Chapter 2)

Chapter Objectives
In this chapter you will be able to:
Choose the correct rating and characteristics of a transformer for a specific application
Calculate the losses, efficiency, and voltage regulation of a transformer under specific operating
conditions.
Experimentally determine the transformer parameters given its ratings.
Utilize the per unit system.

3.1 Introduction
Transformers do not have moving parts, nor are they energy conversion devices, however their
ability to modify the current-voltage characteristics of a given load or source, make them invaluable
components in energy conversion systems. They are utilized for power applications and in low
power signal processing systems. One application in power transmission is the use of transformers
on a transmission line utility pole commonly seen as a cylinder with a few wires sticking out. These
wires enter the transformer through bushings that provide isolation between the wires and the tank.
Inside the tank there is an iron core commonly made of silicon-steel laminations that are 14 mils
(0.014) thick. The insulation often used is paper with the whole coil system immersed in insulating
oil. The oil increases the dielectric strength of the paper and helps to transfer heat from the core/coil
assembly. An drawing of one such distribution transformer is shown in Figure 2.2 in your textbook.
Connection of the transformer to the transmission lines can take several electrical configurations. A
relatively simple connection to a 2.4 kV three phase transmission line is shown in Figure 1.
A 2400 volt
2400
2400 B three-phase
2400 three-wire
C primary system
H1 H2

X3 X1
X2
120 volt
one-phase
two-wire service

Figure 1. Example configuration of a distribution pole transformer connection to


three phase power lines to provide 120 (V) service to your home.

1- 1
A 2400 volt
2400
2400 B three-phase
2400 three-wire
C primary system
H1 H2

X3 X1
X2

120 (V)
240 (V)
120 (V)

Figure 2. Example configuration of a distribution pole transformer connection to


three phase power lines and provides 120 (V) and 240 (V) service.

If a neutral line is also part of the three phase transmission line (perhaps between the substation
and your home), then the connection could be made as shown in Figure 3.

A
4160 2400 4160Y/2400 volt
4160 B three-phase
4160 2400
C four-wire Y
2400 with neutral
N N
H1 H2

X3 X1
X2

120 (V)
240 (V)
120 (V)

Figure 3. Distribution transformer connection to provide 120 (V) and 240 (V)
service from a 4160Y/2400 (V) four-wire transmission line.

1- 2
For a three phase line at the service end, a system could be connected to a four wire three phase
transmission line source as shown in Figure 4 below.
A 4160Y/2400 volt
4160 2400
4160 B three-phase
4160 2400 four-wire Y
C with neutral
2400 primary service
N N
H1 H2 H1 H2 H1 H2

X3 X1 X3 X1 X3 X1
X2 X2 X2
N
120 (V) three-phase
A
120 (V) 208 (V) four-wire Y
B 208 (V) grounded neutral
120 (V) 208 (V) secondary service
C

Figure 4. Three phase to three phase distribution transformer connection providing a


four-wire three phase distribution of 120 (V) and 208 (V) service from a
4160Y/2400 (V) four-wire transmission line.

Transformers are very common place in society, and have had significant impact at many power
levels. They also continue to be improved on today, with such studies as amorphous metals to further
decrease core losses. For more understanding, we begin with the ideal transformer.

3.2 Ideal Transformer


Under ideal conditions, the iron core has infinite permeability and the coils have zero electrical
resistance. Coil 1 has N1 turns, and coil 2 has N2 turns, and all of the magnetic flux is maintained in
the iron (no flux leakage).

i1 i2

e1 e2 F1 F2

Figure 5. Transformer and equivalent magnetic circuit.


1- 3
The electromotive force, or emf, is represented with the symbol e. For an ideal transformer, this
is the voltage at the terminals of a given coil. The flux linkage in each coil is 1 = N1 and
2 = N 2 . The electromotive force induced in each coil is then the time derivative of the flux
linkage or
d d
e1 = 1 = N1 (V) (3.1)
dt dt

d2 d
e2 = = N2 (V) (3.2)
dt dt

The ratio of the voltage at the terminals of coil 1 to the voltage at coil 2 is then
e1 N1
= (3.3)
e2 N 2
Using an equivalent circuit shown in Figure 5 with the magnetomotive force F1 for coil 1 and F2
for coil 2 we would write:
F1 F2 = N1i1 N 2i2 = 0 (3.4)
or
i1 N 2
= (3.5)
i2 N1

Transformers are often used with a voltage source connected to one coil and a load connected to
the other coil. The dots above the coils help to indicate the voltage reference marks for the coils such
d
that for an ideal transformer a positive voltage, v1 = e1 = N1 in coil 1 (primary coil) results in a
dt
d
positive voltage, v2 = e2 = N 2 in the coil 2 (secondary coil) as shown in Figure 6.
dt

i1 i2

N1
v1 v2 Load
N2

Figure 6. Circuit utilizing a transformer.

1- 4
The circuit symbol for the transformer is shown in Figure 7. Since the voltages across the coils
transform in the direct ratio of the turns and the current transforms in the inverse ratio of the turns,
then the impedance can also be transformed through the transformer.

N1 N2
i1 i2

Figure 7. Transformer circuit symbol.

For a voltage applied to the primary coil, and a load impedance, ZL, connected to the secondary as
shown in Figure 8.
N1 N2

+ I1 I2 +

V1 V2 Z L

Figure 8. Transformer with a load connected to the secondary.

The voltages and currents from secondary to primary are related by (3.3) and (3.5), or
N
V1 = 1 V2 (3.6)
N2
N
I1 = 2 I2 (3.7)
N1
Then the impedance measured at the terminals of the primary is
2
V1 N 2 V2
= Z1 = (3.8)
I1 N1 I2
Thus the following circuits are equivalent

1- 5
( )
2
Z L N1
N2
N1 N2 N1 N2
I1 I2 + + I2 I1
+ +
I1
( )
2
V1 V2 ZL V1 V1 Z L N1
N2

(a) (b) (c)

Figure 9. Three equivalent circuits assuming an ideal transformer.

3.3 Non-Ideal Transformer


There are several non-ideal properties that we can take into account by modifying the equivalent
circuit shown in Figure 7. These include a finite core permeability ( c < ), leakage flux, core
losses, and coil resistance. The contribution from each of these is discussed below.

Finite Core Permeability


For the core of the transformer to have a finite permeability, then the circuit in Figure 5 is
modified to include the reluctance of the core.
R

F1 F2

Figure 10. Transformer shown in Figure 5, but with a core of finite permeability.

Then we can write


F1 F2 = N1i1 N 2i2 = R (3.9)

Defining a magnetomotive force that is equal to the drop across the reluctance of the core as:

Fc = N1i1,ex = R (3.10)
Solving for the flux gives
N1i1,ex
= (3.11)
R
The rate of change in the flux is proportional to the induced voltage as

1- 6
N1i1,ex
d
d R
e1 = N1 = N1
dt dt (3.12)
N 2 di1,ex
= 1
R dt

This induced voltage is proportional to the time derivative of the current i1,ex which can be
N12
represented by an inductor in our equivalent circuit with a value of Lm = . The equivalent circuit
R
for the ideal transformer is then modified to account for the finite permeability of the core by placing
an additional inductor across the primary coil as shown in Figure 11.

i1 N1 N2
i 1,ex i'1 i2
+ +

e1 Lm e2

Ideal transformer

Figure 11. Equivalent circuit for a transformer with finite core permeability.

We can also see from Figure 11 that i1 i1,ex = i1 which is in agreement with equations (3.9) and
(3.10).

Leakage Flux
With finite core permeability, not all of the flux will be confined to the metal core, but some will
leak outside the core in the surrounding air. The influence of this leakage flux can also be included
in the equivalent circuit, by considering an additional reluctance associated with the leakage flux, l1,
for coil 1, and l2 for coil 2 such that
Ni
l1 = 1 1
Rl1
(3.13)
N 2i2
l 2 =
Rl 2

These reduce the magnetizing flux, m, of the core, and modify the induced voltages for the primary
(coil 1) and secondary (coil 2) to be

d1 d d N 2 di
v1 = = N1 m + N1 l1 = e1 + 1 1
dt dt dt Rl1 dt

(3.14)
d d d N 2 di
v2 = 2 = N 2 m + N 2 l 2 = e2 + 2 2
dt dt dt Rl 2 dt

1- 7
N2
The last term for v1 can be represented by an inductor, Ll1 = 1 and the last term for v2 can be
Rl1

N 2
represented by Ll 2 = 2 . These can be incorporated into the equivalent circuit as shown in
Rl 2

Figure 12.
i1 N1 N2
Ll1 + i 1,ex i'1 i2 + Ll2

e1 Lm e2

Ideal transformer

Figure 12. Equivalent circuit for a transformer with finite core permeability and
leakage flux.

Core Losses
The dissipation of heat in the core due to the flux through it can be represented by a resistor in the
equivalent circuit. It is placed in parallel to Lm since the power loss in the core is proportional to the
flux through the core squared, and thus is proportional to e12 .

Resistance of the Coils


The resistance of copper is approximately 16.810-9 (m), however with hundreds to thousands
of turns for the primary and secondary coils it can lead to an appreciable resistance. Losses to these
resistances are proportional to the current through the coils squared.

Adding the core losses and resistance of the coils to the equivalent circuit we obtain our
completed model of the transformer as show in Figure 13.

i1 i'1 i2
N1 N2
+ i 1,ex + + +
R1 Ll1 Ll2 R2

V1 Rc Lm e1 e2 V2

Ideal transformer

Figure 13. Equivalent circuit for a non-ideal transformer.

1- 8
Example
Consider a transformer with a turns ratio of 4000/120, with a primary coil resistance R1 = 1.6 (),
a secondary coil resistance of R2 = 1.44 (m), leakage flux that corresponds to Ll1 = 21 (mH), and
Ll2 = 19 (H), and a realistic core characterized by Rc = 160 (k) and Lm = 450 (H). The low voltage
side of the transformer is at 60 (Hz), and V2 = 120 (V), and the power there is P2 = 20 (kW) at
pf = 0.85 lagging. Calculate the voltage at the high voltage side and the efficiency of the transformer.

X m = Lm = 2 60 450 = 169.7 (k)


X1 = Ll1 = 2 60 0.021 = 7.92 ()
X 2 = Ll 2 = 2 60 19 10 6 = 7.16 (m)
P2
I2 = /-31.8 = 196.08/-31.8 (A)
(V2 pf )
E = V + I (R + jX ) = 120.98 + j1.045 (V)
2 2 2 2 2
N
E1 = 1 E 2 = 4032.7 + j 34.83 = 4032.8/0.49 (V)
N2
N
I1 = 2 I2 = 5.001 j 3.1017 (A)
N1
1 1
I1,ex = E1 + = 0.0254 j 0.0236 (A)
Rc jX m
I1 = I1,ex + I1 = 5.0255 j3.125 = 5.918/-31.87 (A)
V = E + I (R + jX ) = 4065.5 + j 69.2 (V) = 4066/0.9 (V)
1 1 1 1 1

The power losses in the windings and core are:

PR 2 = I 22 R2 = 196.082 0.0144 = 55.39 (W)


PR1 = I12 R1 = 5.9182 1.6 = 56.04 (W)
E 2 4032.82
Pc = 1 = = 101.64 (W)
Rc 160 103
Ploss = PR1 + PR 2 + Pc = 213.08 (W)
Pout P2 20 103
= = = = 0.9895
Pin (P2 + Ploss ) 20 103 + 213.08

3.4 Losses and Ratings


The impedances shown in Figure 13 can be reflected into either the primary side, or the secondary
side of the transformer as shown in Figure 14.

1- 9
( )
2
R2 N1
N2 N1 N2
+ + +
R1 Ll1
( )
2
L l2 N1
N2
V1 Rc Lm V'2 V2

( )
2
L l1 N2
N1 N2 N1 Ll2 R2
+ + +
( )
2
R1 N2
N1
V1 V'1
( )
N 2
Rc 2
N1
Lm ( )
N2 2
N1
V2

Figure 14. Reflection of the impedances to the primary or secondary side of the transformer.

For a given frequency, the power losses in the core (iron losses) increase with the voltage e1 (or
e2). If these losses exceed a particular limit, a hot spot in the transformer will reach a temperature
that dramatically reduces the life of the insulation. Limits are therefore put on E1 and E2 which are
the voltage limits for the transformer. The current limits for the transformer limit I1 and I2 to avoid
excessive Joule heating due to winding resistances.
Typically a transformer is described by its rated voltages, E1N and E2N, that give both the limits
I
and turns ratio. The ratio of the rated currents 1N , is the inverse of the ratio of the voltages when
I2N
the magnetizing current is neglected. Instead of listing these rated current limits, a transformer is
described by its rated apparent power as:
S N = E1N I1N = E2 N I 2 N (3.15)
When a transformer is operated at its rated conditions, that is at its maximum current and
maximum voltage, the magnetizing current, I1,ex, typically does not exceed 1% of the current in the
transformer. Its effect on the voltage drop on the leakage inductance and on the winding resistance is
therefore negligible.
Under maximum (rated) current, the total voltage drops on the winding resistances, and leakage
inductances typically do not exceed 6% of the rated voltage. Their effect on E1 and E2 is small and
their effect on the magnetizing current can be neglected.
Because these effects are small, we can modify the equivalent circuits shown in Figure 14 to a
slightly inaccurate, but much more useful on shown in Figure 15.

1- 10
R1 Ll1 N1 N2
+
( ) + +
2
R2 N1 ( )
2

N2 L l2 N1
N2
V1 Rc Lm V'2 V2

R2 Ll2
N1 N2
+ + +
( ) ( )
2 2
R1 N2 L l1 N2
N1 N1
V1 V'1
( )
N 2
Rc 2
N1
Lm ( )
N2 2
N1 V2

Figure 15. Slightly inaccurate, but highly simplified equivalent circuits for the transformer.

When we utilize this simplification and work with the reflected voltages, the transformer
equivalent circuits can be shown as:

i1 Req = R1 + R'2

+ i 1,ex L eq = Ll1+ L'l2 +


i'1
V1 Rc Lm V'2

L eq = L'l1+ L l2
+ Req = R'1 + R 2 +

V'1 R' c L'm V2

Figure 16. Working with V2 or V1 the above approximate circuits for the
transformers can simplify the analysis.

1- 11
Example (repeat with simplified equivalent circuits)
Consider a transformer with a turns ratio of 4000/120, with a primary coil resistance R1 = 1.6 (),
a secondary coil resistance of R2 = 1.44 (m), leakage flux that corresponds to Ll1 = 21 (mH), and
Ll2 = 19 (H), and a realistic core characterized by Rc = 160 (k) and Lm = 450 (H). The low voltage
side of the transformer is at 60 (Hz), and V2 = 120 (V), and the power there is P2 = 20 (kW) at
pf = 0.85 lagging. Calculate the voltage at the high voltage side and the efficiency of the transformer.

Using the simplified equivalent circuits of Figure 16, we can first find Req, and Xeq

2
N
Req = R1 + 1 R2 = 3.2 ()
N2
N1
2

X eq = 2 60 Ll1 + Ll 2 = 15.876 ()

N2

then

P2
I2 = /-31.8 = 196.08/-31.8 (A)
(V2 pf )
E 2 = V2
N
I1 = I2 2 = 5 j3.102 = 5.884/-31.8 (A)
N1
N
E1 = E 2 1 = 4000 (V)
N2
1 1
I1,ex = E1 + = 0.0258 j 0.0235 (A)
Rc jX m
I1 = I1,ex + I1 = 5.0259 j3.125 = 5.918/-31.87 (A)
1 1 1 (
V = E + I R + jX eq eq )
= 4065 + j 69.79 (V) = 4066/0.98 (V)

The power losses in the windings and core are:

PReq = I12 Req = 5.8842 3.2 = 110.79 (W)

V2 40652
Pc = 1 = = 103.28 (W)
Rc 160 103
Ploss = PRe q + Pc = 214.07 (W)
Pout P2 20 103
= = = = 0.9894
Pin (P2 + Ploss ) 20 103 + 214.07

These values agree well with the previous analysis using the more accurate model.

1- 12
3.4 Per Unit System
A simplification in the analysis can come from expressing each of the values as a fraction of a
defined base system of quantities. When this is done, simple problems can be made more complex,
however more complex problems can be made easier to solve. As an example consider a simple
problem of a load impedance of 10 + j5 () that has a voltage of 100 (V) connected to it. Calculate
the power at the load.
The traditional solution is found as:
V 100
IL = L = = 8.94 /-26.57 (A)
Z L 10 + j 5

and the power is

PL = VL I L pf = 100 8.94 cos(26.57 ) = 800 (W)

Using the per unit system to find the solution:

1. First define a consistent system of values for the base. If we choose Vb = 50 (V), Ib = 10 (A),
then Zb = Vb/Ib = 5 (), and Pb = VbIb = 500 (W), Qb = 500 (VAr), and Sb = 500 (VA).
2. Convert all units to pu: VL,pu = VL/Vb = 2pu, ZL,pu = (10 + j5)/5 = 2 + j1 (pu)
3. Solve the problem using the pu system:
VL, pu 2
IL, pu = = = 0.894 /-26.57 (pu)
Z L, pu 2 + j1
PL, pu = VL, pu I L, pu pf = 2 0.894 cos(26.57 ) = 1.6 (pu)
4. Convert back to the SI system
I L = I L, pu I b = 0.894 10 = 8.94 (A)
PL = PL, pu Pb = 1.6 500 = 800 (W)

For the more complicated example of a transformer we choose the bases for each side of the
transformer such that
V1b N
= 1 (3.16)
V2b N 2

I1b N 2
= (3.17)
I 2b N1

This leads to the two base apparent powers being equal

S1b = V1b I1b = V2b I 2b = S 2b (3.18)

The bases are often chosen to be the rated quantities of the transformer on each side. This is
convenient since most of the time transformers operate at rated voltage (making the pu voltage unity),
and the currents and power are seldom above rated (seldom above 1 pu).
The base impedances are related by:
1- 13
V
Z1b = 1b (3.19)
I1b
2
V N V
Z 2b = 2b = 2 1b (3.20)
I 2b N1 I1b
or
2
N
Z 2b = 2 Z1b (3.21)
N1

To move impedances from one side of the transformer to the other, they get multiplied or divided
2
N
by the square of the turns ratio, 2 , but so does the base impedance, hence the pu value of an
N1
impedance stays the same regardless of which side of the transformer it is on.
Through our choice of the bases in (3.16) and (3.17), we also see that for an ideal transformer

E1, pu = E2, pu
I1, pu = I 2, pu

Thus when using the per unit system, an ideal transformer has voltages and currents on one side that
are identical to the voltages and currents on the other side, and the ideal transformer can be
eliminated.

Example (repeat and solved using pu system)


Consider a 30 (kVA) rated transformer with a turns ratio of 4000/120, with a primary coil
resistance R1 = 1.6 (), a secondary coil resistance of R2 = 1.44 (m), leakage flux that corresponds
to Ll1 = 21 (mH), and Ll2 = 19 (H), and a realistic core characterized by Rc = 160 (k) and
Lm = 450 (H). The low voltage side of the transformer is at 60 (Hz), and V2 = 120 (V), and the power
there is P2 = 20 (kW) at pf = 0.85 lagging. Calculate the voltage at the high voltage side and the
efficiency of the transformer.

1. First calculate the impedances of the equivalent circuit.

V1b = 4000 (V)


S1b = 30 (kVA)
30 103
I1b = = 7.5 (A)
4 103
V2
Z1b = 1b = 533 ()
S1b

V2b = 120 (V)


S 2b = S1b = 30 (kVA)
S
I 2b = 2b = 250 (A)
V2b
1- 14
V
Z 2b = 2b = 0.48 ()
I 2b

2. Convert everything to per unit.

R1
R1, pu = = 0.003 (pu)
Z1b
R
R2, pu = 2 = 0.003 (pu)
Z 2b
R
Rc, pu = c = 300 (pu)
Z1b
2 60 Ll1
X l1, pu = = 0.0149 (pu)
Z1b
2 60 Ll 2
X l 2, pu = = 0.0149 (pu)
Z 2b
2 60 Lm
X m, pu = = 318 (pu)
Z1b

V2
V2, pu = = 1 (pu)
V2b
P
P2, pu = 2 = 0.6667 (pu)
S 2b

3. Solve in the per unit system.

P2, pu
I2, pu = /arccos(pf) = 0.666 j0.413(pu)
(V2, pu pf )
( )
V1, pu = V2, pu + I2, pu Req + jX eq = 1.0163 + j 0.01736 (pu)
V1, pu V1, pu
Im, pu = + = 0.0034 j 0.0031 (pu)
Rc, pu jX m, pu
I1, pu = Im, pu + I2, pu = 0.6701 j 0.4163 (pu)

PReq = I 22, pu Req, pu = 0.00369 (pu)

V12,pu
Pc, pu = = 0.00344 (pu)
Rc, pu

P2, pu 0.6667
= = = 0.9894
( P2, pu + Ploss, pu ) 0.6667 + (0.0037 + 0.0034 )

1- 15
4. Convert back to SI units. The efficiency is unitless and thus stays the same. Converting V1
gives
V1 = V1b V1, pu = 4000(1.0163 + j 0.01736 ) = 4065.8 + j 69.46 = 4065.8/0.979 (V)

3.5 Testing Transformers


Purchased transformers often give information on the frequency, winding ratio, power, and
voltage ratings, but not the impedances. These impedances are important in calculating the voltage
regulation, efficiency, etc. Use of open circuit and short circuit tests we will determine Req, Leq, Rc,
and Lm.

Open Circuit Test


Leaving one side of the transformer open circuited, while the other has the rated voltage
{Vin-oc = 1(pu)} applied to it, we measure the current and power. The current that flows into the
transformer is mostly determined by the impedances Xm and Rc, and it is much lower than the rated
current for the transformer. It is often the case that the rated voltage for the low voltage side (low
tension side) of the transform since for the above example it would be easier applying 120 (V) instead
of 4000 (V). Since the units we use indicate if we are using the per unit system, the following
calculations will drop the subscript pu. Using the following equivalent circuit with the primary open
circuited, and V2 = Voc = 120 (V) applied to the secondary.

L eq = L'l1+ L l2
+ Req = R'1 + R 2 +

V'1 R' c L'm V2

Figure 17. Open circuit testing of a 4000/120 rated transformer. With 120 (V)
applied to the low voltage side {V2 = 120 (V)}, the primary voltage for
this open circuit test is 4000 (V).

For this open circuit test, the following can be compared to the measured values of current and
power as:
V2 1
Poc = oc = (pu) (3.22)
Rc Rc

V V
Ioc = oc + oc (3.23)
Rc jX m

1- 16
1 1
I oc = 1 + (pu) (3.24)
Rc2 2
Xm

Then from the open circuit test with measurements of the current and the power, we determine Rc and
Xm.

Short Circuit Test


For the above open circuit test, the low voltage side of the transformer was chosen since it is
easier to apply the lower voltage during that test. Similarly, the rated current is lower on the high
voltage side of the transformer. With the short circuit test, the rated current is commonly applied to
the high voltage side of the transformer (since with a short circuited secondary, the applied voltage
required to reach the rated current is relatively low). With the rated current applied to the high
voltage side, we measure the voltage, Vsc which is V1 for this example, and the power, Psc.

2
Psc = I sc Req = 1 Req (pu) (3.25)

(
Vsc = Isc Req + jX eq ) (3.26)

2 2
Vsc = 1 Req + X eq (pu) (3.27)

i1 Req = R1 + R'2

+ i 1,ex L eq = Ll1+ L'l2 +


i'1
V1 Rc Lm V'2 = 0

Figure 18. Short circuit testing of a 4000/120, 30 (kVA) rated transformer. This
corresponds to the rated current on the high voltage side of 7.5 (A).

Then from the short circuit test with measurements of the current and the power, we determine
Req = R1 + R2 and X m = 2 60(Ll1 + Ll2 ) .

Example
A 60Hz transformer is rated 30 (kVA), 4000(V)/120(V). The open circuit test, with the high
voltage side open, gives Poc = 100 (W), Ioc = 1.1455 (A). The short circuit test, measured with the
low voltage side shorted, gives Psc = 180 (W), Vsc = 129.79 (V). Determine the equivalent circuit for
this transformer by using the per unit system.

1- 17
1. Bases:
V1b = 4000 (V)
S1b = 30 (kVA)

30 103
I1b = = 7.5 (A)
4 103

V2
Z1b = 1b = 533 ()
S1b
V2b = 120 (V)

S 2b = S1b = 30 (kVA)

S 30 103
I 2b = 1b = = 250 (A)
V2b 120
V 120
Z 2b = 2b = = 0.48 ()
I 2b 250

2. Convert to (pu):
180
Psc, pu = = 0.006 (pu)
30 103
129.79
Vsc, pu = = 0.0324 (pu)
4 103
100
Poc, pu = = 0.00333 (pu)
30 103
1.1455
I oc, pu = = 0.00458 (pu)
250

3. Calculate the components of the equivalent circuit {dropping the (pu) subscripts}
2 P
Psc = I sc Req or Req = sc = 1 Psc = 0.006 (pu)
2
I sc
2 2
Vsc = Vsc = Isc Req + jX eq = 1 Req + X eq or X eq = Vsc2 Req
2
= 0.0318 (pu)

V2 V2 1
Poc = oc or Rc = oc = = 300 (pu)
Rc Poc Poc

1- 18
V V 1 1 1
Ioc = I oc = oc + oc = + or Xm = = 318 (pu)
Rc jX m Rc2 2
Xm 2
I oc
1
2
Rc

With this knowledge, we can address the more common example of:

A 60Hz transformer is rated 30 (kVA), 4000(V)/120(V). The short circuit impedance is


0.0324 (pu) and the open circuit current is 0.0046 (pu). The rated core losses are 100 (W) and the
rated winding losses are 180 (W). Calculate the efficiency and the necessary primary voltage when
the load at the secondary is at rated voltage, 20 (kW), and at 0.8 pf lagging.

Using (pu) system:

Z sc = 0.0324 (pu)
2 180
Psc = I sc Req = 1 Req = = 0.006 (pu)
30 103
2 2
X eq = Z sc Req = 0.017 (pu)
1 1 1
Poc = or Rc = = = 300 (pu)
Rc Poc 100

30 103
1 1 1
I oc = + or Xm = = 318 (pu)
Rc2 X m
2
2 1
I oc 2
Rc

Now we have the equivalent circuit components of the transformer and can work on the full
circuit which includes the load that has a power of 20 (kW) or:
20 103
P2 = = 0.6667 (pu)
30 103

but the power at the load is

P2 = V2 I 2 pf or 0.6667 = 1 I 2 0.8 thus I 2 = 0.8333 (pu)

as a phasor: I2 = 0.8333 /-36.87 = 0.6667 j0.5 (pu)

( )
V1 = V2 + I2 Req + jX eq = 1.0125 + j 0.008334 = 1.0125/0.472 (pu)
V1 = 1.0125 (pu)

PReq = I 22 Req = 0.0062 (pu)

1- 19
V2
Pc = c = 0.034 (pu)
Rc
P2
= = 0.986
P2 + PReq + Pc

converting V1 to SI units gives


V1 = V1, pu V1b = 1.0125 4000 = 4050 (V)

3.6 Three-Phase Transformers


If we consider three-phase transformers as consisting of three identical one-phase transformers,
then we have an accurate representation as far as equivalent circuits and two-port models are
concerned, but it does not give insight into the magnetic circuit of the three-phase transformers.
The primaries and the secondaries of the one-phase transformers can be connected either in or
in Y configurations. In either case, the rated power of the three-phase transformer is three times that
of the one-phase transformers.
For the connection:
Vl l = V1 (3.28)
I l = 3I1 (3.29)

For the Y connection:


Vl l = 3V1 (3.30)
I l = I1 (3.31)

Connections to the three-phase transformer that are Y connected in the primary are shown in Figure
19.

1- 20
I1 I2 I1 I2

V1 V2 V1 V2

I1 I2 I1 I2

V1 V2 V1 V2

I1 I2 I1 I2

V1 V2 V1 V2

Figure 19. Y Y and Y connections of three-phase transformers.

Connections to the three-phase transformer that are connected in the primary are shown in Figure
20.

I1 I2 I1 I2

V1 V2 V1 V2

I1 I2 I1 I2

V1 V2 V1 V2

I1 I2 I1 I2

V1 V2 V1 V2

Figure 20. Y and connections of three-phase transformers.

1- 21
3.7 Autotransformers
An autotransformer is a transformer where the two windings (of turns N1 and N2) are not isolated
from each other, but are connected as shown in Figure 21.

I1

N1
V1 I2

N2
I1 I2 V2

Figure 21. An autotransformer.

From this figure that the voltage ratio in an autotransformer is:

V1 N1 + N 2
= (3.32)
V2 N2

and the current ratio is:


I 2 N1 + N 2
= (3.33)
I1 N2

An interesting note on autotransformers is that the coil of turns N1 carries current I1, while the coil of
turns N2 carries the (vectorial) sum of the two currents, I1 I2 . So if the voltage ratio were 1, no
current would flow through the N2 coil. This characteristic leads to a significant reduction in the size
of the autotransformer compared to a similarly rated transformer, especially if the primary and
secondary voltages are of the same order of magnitude. These savings come at a serious
disadvantage of the loss of isolation between the two sides.

1- 22
Chapter Notes:
To understand the operation of transformers we have to use both the Gausss law for magnetism (or
Biot-Savart law) and Faradays law.
Most transformers operate under or near rated voltage. The voltage drop in the winding resistance
and leakage reactance are usually small.
In both transformers and in rotating machines, the net mmf of all the currents must accordingly
adjust itself to create the resultant flux required by this voltage balance.
Leakage fluxes induce voltage in the windings that are accounted for in the equivalent circuit as
leakage reactance (elements Ll1 and Ll2). The leakage-flux paths are dominated by paths through
air, and are thus almost linearly proportional to the currents producing them. The leakage
reactances therefore are often assumed to be constant (independent of the degree of saturation of
the core material).
The open- and short-circuit tests provide the parameters for the equivalent circuit of the
transformer.
Three-phase transformers can be considered to be made of three single-phase transformers for the
purpose of this course. The main issue is then to calculate the ratings, voltages, and currents of
each.
Autotransformers are used mostly to vary the voltage a little. It is seldom that an autotransformer
will have a voltage ratio greater than two.

1- 23
ECE 320
Energy Conversion and Power Electronics
Dr. Tim Hogan

Chapter 4: Concepts of Electrical Machines: DC Motors


(Textbook Sections 3.1-3.4, and 4.1-4.2)

Chapter Objectives
DC machines have faded from use due to their relatively high cost and increased maintenance
requirements. Nevertheless, they remain good examples for electromechanical systems used for
control. Well study DC machines here, at a conceptual level, for two reasons:

1. DC machines although complex in construction, can be useful in establishing the concepts of


emf and torque development, and are described by simple equations.
2. The magnetic fields in them, along with the voltage and torque equations can be used easily to
develop the ideas of field orientation.

In doing so we will develop basic steady state equations, again starting from fundamentals of the
electromagnetic field. We are going to see the same equations in Brushless DC motors, when we
discuss synchronous AC machines.

4.1 Geometry, Fields, Voltages, and Currents


The geometry shown in Figure 1 describes an outer iron frame (stator), through which (i.e. its
center part) a uniform magnetic flux, , is established. The flux could be established by a current in
a coil or by a permanent magnet for example.
In the center part of the frame there is a solid iron cylinder (called rotor), free to rotate around its
axis. A coil of one turn is wound diametrically around the cylinder, parallel to its axis, and as the
stator and its coil rotate, the flux through the coil changes. Figure 2 shows consecutive locations of
the rotor and we can see that the flux through the coil changes both in value and direction. The top
graph of Figure 3 shows how the flux linkages of the coil through the coil would change, if the rotor
were to rotate at a constant angular velocity, .
= cos(t ) (4.1)

Figure 1. Geometry of an elementary DC motor.

1- 1
1 2 3 4 5
Figure 2. Flux through the stator coil of the simple dc motor shown in Figure 1.

Since the flux linking the coil changes with time, than a voltage will be induced in this coil, vcoil, as
shown below:
d
vcoil = = sin (t ) (V) (4.2)
dt
1.5

1
1
2
0.5

3
coil

-0.5 4
5
-1

-1.5
0 100 200 300 400 500 600 700
Angle ()
1.5

0.5

1
dcoil

0
5
v

-0.5 2
4
-1
3
-1.5
0 100 200 300 400 500 600 700
Angle ()
Figure 3. Flux and voltage in a coil of the motor in Figure 2 with coil positions 1-5.

1- 2
The points marked on the cosine and sine waveforms of Figure 3 correspond to the positions of the
rotor as shown in Figure 2.
Torque on the coil follows the Lorentz Force Law: which gives the force, F, in newtons on a
charge, q, in coloumbs that is exposed to a vectoral electric field, E , in volts per meter and magnetic
flux density, B, as
(
F = q E + v B ) (4.3)
where v is the velocity of a charged particle q. The first term gives the force on the electron that
moves it through the wire or the voltage applied to the wire. The second term gives the force on the
charge due to the magnetic field in which the wire is placed. With many electrons contributing to the
electrical current, I, (as a vector) in the wire, the force on the wire is
F = IB (4.4)
The force on the full coil is then twice this value to account for each of the wires shown in Figure 2.
To maintain this torque in a direction that continues the rotary motion of the rotor, the ends of the
rotor coil are connected to metal electrode ring segments called a commutator as shown in Figure 4.
These ring segments are attached to the rotor so as to rotate with it, and are electrically contacted
using two stationary brushes typically made of carbon and copper. The brushes are spring loaded and
pushed against the commutator. As the rotor spins, the brushes make contact with the opposite
segments of the commutator and they switch between segments of the commutator just as the induced
voltage goes through zero and switches sign.

bru
sh
or
utat
m
com

bru
sh

Figure 4. A coil of a DC motor and brushes that are pushed up against the rotating
commutator (wire of the coil is highlighted red and soldered to the
commutator segments).

Because the current switches direction through the coil each time the brushes rotate to the
opposite segment of the commutator (and that this occurs as the voltage goes through zero), the
1- 3
voltage at the terminals of the brushes is a rectified version of the voltage induced in the coil as
shown in Figure 5.

1.5 1.5

1 1

0.5 0.5

terminal
coil

0 0
v

v
-0.5 -0.5

-1 -1

-1.5 -1.5
Time (arb. units) Time (arb. units)

Figure 5. Induced voltage in a coil and terminal voltage in an elementary DC machine.

If a number of coils are placed on the rotor, as shown in Figure 6, with each coil connected to a
different segment of the commutator, then the total induced voltage to the coils, E, will be:
E = k (4.5)
where k is proportional to the number of coils.

Figure 6. Multiple coils on the rotor of a DC machine.

We stated in equation (2.39) that Pelec = Pmech T mech = e i thus with the multiple coil DC
machine:
E i = T (4.6)
k i = T (4.7)
T = ki (4.8)

1- 4
If the DC machine is connected to a load or a source as in Figure 7, then the induced voltage and
terminal voltage will be related by:

Vterminals = E ig Rwdg for a generator (4.9)


Vterminals = E + im Rwdg for a motor (4.10)

im
ig

Rwdg Load

E or

Source

Figure 7. Circuit with a DC machine with the current reference directions defined as
indicated for DC machine used as a generator, ig, or as a motor, im.

Example 4.1.1
A DC motor, when connected to a 100 (V) source and with no load connected to the motor runs at
1200 (rpm). Its stator resistance is 2 (). What should be the torque and current if it is fed from a
220 (V) supply and its speed is 1500 (rpm)? Assume the field is constant.
With no load connected to the motor, we will assume the torque is zero (assuming no friction in
bearings). With the torque equal to zero, the current is zero since: T = ki = Ki . This means for this
operation:
V = E = k = K

2 (rad) 1 (min)
and with = 1200 (rpm ) = 125.66 (rad/s) , then
revolution 60 (s)
100 (V) = K 125.66
K = 0.796 (V s)
Then at 1500 (rpm)
2 (rad) 1 (min)
o = 1500 (rpm ) = 157.08 (rad/s)
revolution 60 (s)

E = K = 125 (V)

1- 5
For a motor:
V = E + IRwdg

220 (V) = 125 (V) + IRwdg

I = 47.5 (A)
T = KI = 37.81 (Nm)

4.2 Energy Considerations


Conservation of energy governs that the total energy supplied to an electromechanical system
equals the energy output. Some of the energy output could be mechanical, some could be lost as
heat, and some could be stored. This is summarized in equation (3.10) of your textbook as:
Energy input Mechanical Increase in energy Energy
from electric = energy + stored in the + converted
sources output magnetic field into heat

In a lossless system, the last term is zero. When the losses are negligible, a differential change in
the electrical energy input corresponds to the sum of a differential mechanical energy output and a
differential change in the energy stored in the magnetic field, or
dWelec = dWmech + dWfld (4.11)
Mechanical energy is force times distance, and the time rate of change of electrical energy is
power or dWelec = ei dt and (4.11) can be written as:
ei dt = Ffld dx + dWfld (4.12)
where Ffld is the force produced by the magnetic field. From equation (3.1) in the handout notes, we
d N
saw e = . Using the linear inductor assumption of (2.19) which was L = = , then (4.12)
dt i i
becomes:
dWfld = i d Ffld dx (4.13)
The energy in the field is given as the energy stored in an inductor

1 2 1 2
Wfld = Li = (4.14)
2 2 L
The current and force produced by the magnetic field can then be found from (4.13) as:
Wfld ( , x )
i= (4.15)
x

Wfld ( , x ) i 2 dL(x )
Ffld = = (4.16)
x 2 dx

If the magnetic force results in a torque as in a rotating mechanical terminal, the mechanical energy
from the field is then replaced with torque and angular displacement, Ffld dx Tfld d , as:

1- 6
dWfld = i d Tfld d (4.17)

thus leading to
i 2 dL( )
Tfld = (4.18)
2 d

In a rotary machine with multiple windings, the inductance in equation (4.18) would contain self and
mutual inductances of all coils involved.

Chapter Notes:

The field of a DC motor can be created either by a DC current or a permanent magnet.


The two fields, the one coming from the stator and the one coming from the moving rotor, are both
stationary (despite rotation) and they are perpendicular to each other.
If the directions of current in the stator and in the rotor reverse together, torque will remain in the
same direction. Hence if the same current flows in both windings, it could be AC and the motor
will not reverse.

1- 7
ECE 320
Energy Conversion and Power Electronics
Dr. Tim Hogan

Chapter 5: Three Phase Windings


(Textbook Sections 4.3-4.7)

Chapter Objectives
Flux linkage plays a crucial role in the operation of both DC and AC machines. In this chapter,
the geometry and the operation of windings in AC machines is discussed. The flux varies in time,
and can also vary in position, or be stationary. To understand how these machines operate, the
concept of space vectors (or space phasors) is introduced.

5.1 Introduction
Electric machines often have defined an armature winding which is the winding that is power
producing, and a field winding that generates the magnetic field. Either could be on the stator or
rotor depending on the specific motor or generator; however it is more common with AC machines
such as synchronous or induction machines that the armature winding is on the stator (the stationary
portion of the motor). Synchronous machines have field windings on the rotor that is excited by
direct current delivered to the rotor windings by slip rings or collector rings by carbon brushes. The
field winding produces the north and south poles, thus the image shown in Figure 1 is for a two-pole,
single phase (one armature winding) synchronous generator. The magnetic axis for the armature
winding is perpendicular to the area defined by the armature winding (armature winding is the
perimeter of this area).

Figure 1. Figures 4.4 and 4.5 from your textbook showing a simple two-pole, single phase
synchronous generator, the spatial distribution of the magnetic field relative to the magnetic axis
of the armature winding, and the time dependent induced voltage in the armature winding [1].

1- 1
In a three phase device, the armature has three coils each with a magnetic axis that is rotated
spatially by 120 as shown in Figure 2.

Figure 2. A three-phase, two-pole synchronous generator as shown in Figure


4.12(a) in your textbook [1].

More poles for the field winding and more armature windings are also possible as shown in
Figure 3. Such a configuration can deliver three phase power by interconnecting the armature
windings in a Y connection configuration as an example.

Figure 3. A four-pole synchronous generator with multiple armature windings that can
be wired together in a three-phase Y connection is shown (from Figure 4.12(b) and
4.12(c) in your textbook [1]).

5.2 Control of the Magnetomotive Force Distribution


As generators, the synchronous machines typically use the stator windings as a source of
electrical power. As motors, the stator (or armature) windings are commonly supplied electrical
power to generate a spatially varying field (we will consider the time variation of these fields later).
Insight to the distribution (in space) of the field from the armature windings can be seen by
considering a single N-turn coil on the stator that spans 180 electrical degrees (in 180 it has gone

1- 2
from a +F to a F). Such a coil is known as a full pitch coil. The mmf distribution for such a full
pitch coil is depicted in Figure 4.19 from your textbook as shown below in Figure 4.

Figure 4. A full pitch coil on the stator. Windings on the rotor are left off for clarity. This is
Figure 4.19 in your textbook. (a) Schematic view of flux produced by a concentrated, full-
pitch winding in a machine with a uniform air gap. (b) The air-gap mmf produced by current
in this winding [1].

Assuming the reluctances of the stator and rotor negligible compared to the gaps, then the full
mmf of Ni would drop across two gaps as the flux traverses a full loop. This gives rise to a +(Ni/2)
maximum, and a (Ni/2) minimum for the mmf as one spatially maps F for the stator winding as a
function of the angle relative to its magnetic axis. Plotting F as a function of this angle, a, shows
abrupt changes at the wires of the coil. The Fourier transform of this square wave gives a
fundamental sinusoidal distribution of Fag1 as shown in Figure 4.
Harmonics exist since it is non-sinusoidal. In an attempt to make the spatial distribution more
sinusoidal in form, multiple coils can be placed within specific groves of the stator as shown for one
of the phases in a three-phase winding in Figure 5. Here there are equal numbers of wires in each of
the groves, and equal current in each of the wires.

1- 3
Figure 5. The mmf of one phase of a distributed two-pole, three phase winding with full-pitch
coils. This is image 4.20 from your textbook [1].

By adjusting the number of turns in each slot a more sinusoidal mmf distribution can be also be
obtained - as shown for windings of a rotor in the configuration shown in Figure 6.

Figure 6. The air-gap mmf of a distributed winding on the rotor of a round-rotor generator. This
is image 4.21 from your textbook [1].

Some simple methods for controlling the shape of the magnetomotive force for a set of coils have
been outlined. Now we turn our attention to how we use this information for analyzing the electric
machine through the concept of space vectors.

1- 4
5.3 Current Space Vectors
Consider three identical windings placed in a space of uniform permeability as shown in Figure 7.
Each winding carries a time dependent current, i1(t), i2(t), and i3(t), We also require that
i1(t ) + i2 (t ) + i3 (t ) 0 (5.1)

Each current produces a flux in the direction of the coil axis, and if we assume the magnetic
medium to be linear, we can find the total flux by adding the individual fluxes. This means that we
could produce the same flux by having only one coil, identical to the three, but placed in the direction
of the total flux, carrying an appropriate current. If the coils in Figure 7 carry the following currents
i1 = 5 (A), i2 = -8 (A), i3 = 3 (A), then the vectoral sum of these currents, oriented in the same
direction as the corresponding coil can be shown in Figure 8.

I1
I2
1
I3

3
Figure 7. Three phase windings spatially oriented 120 apart.

The direction of the resultant coil and current it should carry, we create three vectors, each in the
direction of one coil, and equal in amplitude to the current of the coil it represents. If, for example,
the coils are placed at angles of 0, 120, 240. Then their vectoral sum will be:

i = i / = i1 + i2e j120 + i3e j 240 (5.2)

This represents the vector of a current oriented in space, and thus we define i as a space vector. If
i1, i2, and i3 are functions of time, so will be the amplitude and angle of i. If we consider a horizontal
axis in space as a real axis, and a vertical axis as an imaginary axis, then we can find the real
(id = Re{i}) and imaginary (iq = Im{i}) components of the space vector. With this representation, we
can determine the three currents from the space vector as:

1- 5
2
i1(t ) = Re{i (t )}
3
2
{
i2 (t ) = Re i (t )e j
3
} (5.3)

2
{
i3 (t ) = Re i (t )e j 2
3
}
2
= 120 = rad (5.4)
3

I
I1 = 5 (A) I1

I3 = 3 (A)
I2
I2 = - 8 (A)
I
I3

Figure 8. (a) Currents in the three windings of Figure 7. (b) Resultant space
vector and (c) corresponding winding position and current of an
equivalent single coil.

Consider if the three coils in Figure 7 were to represent the three stator windings of an AC
machine as shown in Figure 9, with currents for each phase also shown in Figure 9.

(a) (b)
Figure 9. Simplified two pole 3-phase stator winding and the instantaneous currents
for each phase. These are Figures 4.29 and 4.30 from your textbook [1].
1- 6
Then for:
ia (t ) = I m cos(et )
ib (t ) = I m cos(et 120 ) (5.5)
ic (t ) = I m cos(et + 120 )

the space vector is:

[
i (t ) = I m cos(et ) + e j120 cos(et 120 ) + e j 240 cos(et + 120 ) ]
= I m [cos(et ) + {cos(120 ) + j sin (120 )}cos(et 120 ) + {cos(240 ) + j sin (240 )}cos(et + 120 )]
3
= I m cos(et ) 0.5 cos(et 120 ) 0.5 cos(et + 120 ) + j {cos(et 120 ) cos(et + 120 )}
2
(5.6)

Using: cos( x y ) = cos( x ) cos( y ) sin ( x )sin ( y )

3 3
i (t ) = I m cos(et ) 0.5 0.5 cos(et ) + sin (et ) 0.5 0.5 cos(et ) sin (et )
2 2
3 3 3
+ jI m 0.5 cos(et ) + sin (et ) + 0.5 cos(et ) + sin (et ) (5.7)
2 2 2
3
= I m [cos(et ) j sin (et )]
2

This is a space vector that rotates in space as a function time at an angular frequency of e.
Graphically, this can be seen from summation of ia, ib, and ic in Figure 9 (b) at different points in time
(et = 0, /3, and 2/3), the resultant space vector at each point in time is represented by the
corresponding magnetomotive force, F, associated with that space vector as shown in Figure 10.

Figure 10. The production of a rotating magnetic field by means of three-phase currents.

1- 7
Section 5.2 in these notes described ways of forming magnetomotive force spatial distributions
that were more sinusoidal in profile. Representing such windings as sinusoidally concentrated
windings in the stator as depicted in Figure 11, then the density of turns of a the coil would vary
sinusoidally in space as function of the angle .

Stator Winding

Stator

Air Gap

Rotor

Figure 11. The production of a rotating magnetic field by means of three-phase currents.

Thus the number of turns, dNs, covering an angle d at a position over d is a sinusoidal
function of the angle . The turns density, ns1( ) is then:

dN s
= ns1( ) = n s sin (5.8)
d

For the total number of turns, Ns, in the winding:


Ns =
0 ns1( )d (5.9)

This leads to
Ns
ns1( ) = sin (5.10)
2

With i1 current flowing through this winding, the flux density in the air gap between the rotor and the
stator can be found for the integration path shown in Figure 12. The path of integration is defined by
the angle q and we can notice that because of symmetry of the flux density at the two air gap
segments in the path is the same.

1- 8
Current in positive direction

Stator

Air Gap

Rotor

Current in negative direction Stator Winding

Figure 12. Integration path to calculate flux density in the air gap.

Assuming the permeability of the stator and rotor is infinite, then Hiron = 0 and:

+
2 H g1( )g =
i1ns1( )d
2 Bg1( )
g = i1N s cos (5.11)
0
N s 0
Bg1( ) = i1 cos
2g

For a given current, i1, in the coil the flux density in the air gap varies sinusoidally with angle, but
as shown in Figure 13 it reaches a maximum when the angle is zero. For the same machine and
conditions as in Figure 13, Figure 14 shows the plot of turns density, ns() and flux density, Bg() in
B

cartesian coordinates with as the horizontal axis. For this one coil, if the current, i1, were to vary
sinusoidally in time, then the flux density would also change in time. The direction of the space
vector would be maintained however the amplitude would change in time. The nodes of the flux
density where it is equal to zero will remain at 90 and 270, while the extrema of the flux will be
maintained at 0 and 180.

1- 9
Figure 13. Sketch of the flux (red) in the air gap.

1 1

0.5 0.5
B ()
n ()

0 0
g
s

-0.5 -0.5

-1 -1
0 60 120 180 240 300 360 0 60 120 180 240 300 360

Figure 14. Turns density on the stator and the air gap flux density vs. .

Consider now an additional winding, identical to the first, but rotated with respect to it by 120.
For a current in this winding we will get a similar air gap flux density as before, but with nodes at
210 = 90 + 120 and at 30 = 270 + 120. If a current, i2, is flowing in this winding, then the air
gap flux density due to it will follow a form similar to equation (5.11) but rotated by 120 = (2/3).

N s 0 2
Bg 2 ( ) = i2 cos (5.12)
2g 3
Similarly, a third winding, rotated 240 relative to the first winding and carrying current i3, will
produce an air gap flux density of:
N 4
Bg 3 ( ) = i3 s 0 cos (5.13)
2g 3

1- 10
Combining these three flux densities, we obtain a sinusoidally distributed air gap flux density, that
could equivalently come from a winding placed at an angle and carrying current i as:

N s 0
Bg ( ) = Bg1( ) + Bg 2 ( ) + Bg 3 ( ) = i cos( + ) (5.14)
2g

This means that as the currents change, the flux could be due instead to only one sinusoidally
distributed winding with the same number of turns. The location, (t), and current, i(t), of this
winding can be determined from the current space vector:

i (t ) = i (t ) / = i1(t ) + i2 (t )e j120 + i3 (t )e j 240 (5.15)

5.3.1 Balanced, Symmetric Three-phase Currents


If the currents i1, i2, i3 form a balanced three-phase system of frequency fs = s/2, then we can
write:

i1 = 2 I cos(s t + 1 ) =
2
2
[
I s e j s t + I s e j s t ]

i2 = 2 I cos s t + 1

2
=
3 2
2
[
I s e j ( s t 2 3) + I s e j ( s t 2 3) (5.16) ]

i3 = 2 I cos s t + 1

4
=
3 2
2
[I e (
s
j s t 4 3)
+ I s e j ( s t 4 3 )]

where I is the phasor corresponding to the current in phase 1. The resultant space vector is:

3 2 j s t 3 2 j ( s t +1 )
i s (t ) = Ie = Ie (5.17)
2 2 2 2

The resultant flux density wave is then:

3 N
B( , t ) = 2 I s 0 cos(s t + 1 ) (5.18)
2 2g

3 N
which shows a travelling wave, with a maximum value of Bmax = 2 I s . This wave travels
2 0
around the stator at a constant speed s, as shown Figure 15.

1- 11
1

0.5
t t t
1 2 3

B ()
0

g
-0.5

-1
0 60 120 180 240 300 360
Angle ()

Figure 15. Air gap flux density profile vs. for three times t3>t2>t1.

5.4 Phasors and Space Vectors


This is a good point to reflect on the differences between phasors and space vectors. A current
phasor, I = Ie j0 , describes one sinusoidally varying current of frequency , amplitude 2 I , and
initial phase 0. The sinusoid can be reconstructed from the phasor as:

i(t ) =
2
[
2 j t * j t
Ie +I e ]
= 2 I cos( t + 0 ) = Re Ie j t ( ) (5.19)

Although rotation is implicit in the definition of the phasor, no rotation is described by it.
On the other hand, the definition of current space vector requires three currents that sum to zero.
These currents are implicitly in windings that are symmetrically placed, but the currents are not
necessarily sinusoidal. Generally the amplitude and angle of the space vector changes with time, but
no specific pattern is a priori defined. We can reconstruct the three currents that constitute the space
vector from equation (5.3). When these constituent currents form a balanced, symmetric system, of
frequency s, then the resultant space vector is of constant amplitude, rotating at a constant speed. In
that case, the relationship between the phasor of one current and the space vector is shown in
equation (5.17).

5.5 Magnetizing Current, Flux, and Voltage


To see how this rotating magnetic flux influences the windings we use Faradays law. From here
on we will use sinusoidal symmetric three-phase quantities.
The three stationary windings are linked by a rotating flux as shown in Figure 16. When the
current is maximum in phase 1, the flux is as shown in Figure 16(a) and is linking all of the turns in
phase 1. Later, the flux has rotated as show in Figure 16(b) then the flux linkages with coil 1 have
decreased. When the flux has rotated 90 as in Figure 16(c), the flux linkages with the phase 1
windings are zero.
1- 12
(a) (b) (c)

Figure 16. Rotating flux and flux linkages. Sinusoidal windings containing many turns in
the stator are represented by single wires to show current flow direction.

To calculate the flux linkages, l, we determine the flux through a differential number of turns at
an angle as shown in Figure 17.

Figure 17. Flux linkage through a differential number of turns (such as one turn) for
coil 1.

The flux through this differential section of coil is then:


(t , ) =
Bg (t , )dA = lr
Bg (t, )d (5.20)

1- 13
where l is the axial length of the coil (into the page), and r is the radius to the coil.

The number of turns linked by this flux is dns ( ) = ns ( )d , so the flux linkages for these few (or
one) turns is:

d = ns ( )d ( ) (5.21)

To find the flux linkages, 1, for all of coil 1, then we must integrate the flux linkages over all turns
of coil 1 or:

0
1 = d = ns ( )d ( ) (5.22)

When these two integrals are taken, then we find:

N s2lr
1(t ) = 30 2 I cos(st + 1 ) = LM 2 I cos(st + 1 ) (5.23)
8g

which means the flux linkages in coil 1 are in phase with the current in this coil and proportional to it.
The flux linkages of the other two coils, 2 and 3, are identical to that of coil 1, but lagging in time by
120 and 240 respectively. With these three quantities we can create a flux-linkage space vector,
as:
= 1 + 2e j120 + 3e j 240 = LM i (5.24)

Since the flux linkages of each coil vary, and in our case sinusoidally, a voltage is induced in each
of these coils. The induced voltage in each coil is 90 ahead of the current in it, bringing to mind the
relationship of current and voltage of the inductor. Notice though, that it is not just the current in the
winding that causes the flux linkages and the induced voltages, but rather the current in all three
windings. Although this is the case, we sill call the constant LM the magnetizing inductance, and the
induced voltages in each coil can be found as:
d1
e1(t ) = = 2 I cos t + 1 +
dt 2
d2 2
e2 (t ) = = 2 I cos t + 2 + (5.25)
dt 2 3
d 4
e3 (t ) = 3 = 2 I cos t + 3 +
dt 2 3
and the voltage space vector, e can be defined as:

e = e1 + e2e j120 + e3e j 240 = jLM i (5.26)

The flux linkage space vector is aligned with the current space vector, while the voltage space vector,
e, is ahead of both by 90. This agrees with the fact that the individual phase voltages lead the current
by 90, as shown in

1- 14
e

Figure 18. Magnetizing current, flux-linkage, and induced voltage space vectors.

1 A. E. Fitzgerald, C. Kingsley, Jr., S. D. Umans, Electric Machinery, 6th edition, McGraw-Hill, New York, 2003.

1- 15
ECE 320
Energy Conversion and Power Electronics
Dr. Tim Hogan

Chapter 6: Induction Machines


(Textbook Sections 6.1-6.5)

Chapter Objectives
The popularity of induction machines has helped to label them as the workhorse of industry.
They are relatively easy to fabricate, rugged and reliable, and find their way into most applications.
For variable speed applications, inexpensive power electronics can be used along with computer
hardware and this has allowed induction machines to become more versatile. In particular, vector or
field-oriented control allows induction motors to replace DC motors in many applications.

6.1 Description
The stator of an induction machine is a typical three-phase one, as described in the previous
chapter. The rotor can be one of two major types either (a) it is wound in a fashion similar to that
of the stator with the terminals connected to slip rings on the shaft, as shown in Figure 1, or (b) it is
made with shorted bars.

Shaft
Slip
Rings
Rotor
Figure 1. Wound rotor, slip rings, and connections.

Figure 2 shows the rotor of such a machine, while the images in Figure 3 show the shorted bars
and the laminations.
The bars in Figure 3 are formed by casting aluminum in the openings of the rotor laminations. In
this case the iron laminations were chemically removed.

1- 1
Figure 2. (a) Cutaway view of a three-phase induction motor with a wound rotor and slip rings
connected to the three-phase rotor winding shown in Figure 6.1 in your textbook [1]. (b)
Cutaway view of a three-phase squirrel-cage motor as shown in Figure 6.3 in your textbook [1].

(a) (b)

Figure 3. (a) The rotor of a small squirrel-cage motor. (b) The squirrel-cage structure
after the rotor laminations have been chemically etched away as shown in Figure 6.3 in
your textbook [1].

6.2 Concept of Operation


As these rotor windings or bars rotate within the magnetic field created by the stator magnetizing
currents, voltages are induced in them. If the rotor were to stand still, then the induced voltages
would be very similar to those induced in the stator windings. In the case of a squirrel cage rotor, the
voltage induced in a bar will be slightly out of phase with the voltage in the next bar, since the flux
linkages will change in it after a short delay. This is depicted in Figure 4.
If the rotor is moving at synchronous speed, together with the field, no voltage will be induced in
the bars or the windings.

1- 2
1

7 6 0.5
5 bar 7
4 bar 1
bar 2 bar 3
3

e(t)
0
2
13 1
Bg -0.5

-1
0 50 100 150 200 250 300 350
19 t

(a) (b)

Figure 4. (a) Rotor bars in the stator field and (b) voltages in the rotor bars.

Generally when the synchronous speed is s = 2f s , and the rotor speed 0, the frequency of the
induced voltages will be fr, where 2f r = s o . Maxwells equation becomes here:

E = v Bg (6.1)

where E is the electric field and v is the relative velocity of the rotor with respect to the field:
v = (s o )r (6.2)

Since a voltage is induced in the bars, and these are short-circuited, currents will flow in them. The
current density J ( ) will be:
1
J ( ) = E (6.3)

where is the resistivity of the bars.

These currents are out of phase in different bars, just like the induced voltages. To simplify the
analysis we can consider the rotor as one winding carrying currents sinusoidally distributed in space.
This will be clearly the case for a wound rotor. It will also be the case for uniformly distributed rotor
bars, but now each bar, located at an angle will carry different current, as shown in Figure 5(a).

1- 3
Bg Bg

(a) (b)

Figure 5. (a) Currents in rotor bars and (b) equivalent current sheet in the rotor.

1
J= (s o ) B g ( ) (6.4)

1
J ( ) = (s o )B g sin( ) (6.5)

The bars can also be replaced with a conductive cylinder as shown in Figure 5(b) with a
distributed current.
Slip, s, is defined as the ratio:
o
s= s (6.6)
s

Thus at starting, the speed is zero and s = 1, and at synchronous speed, s = o and s = 0. Above
synchronous speed s < 0, and when the rotor rotates in a direction opposite of the magnetic field, then
s > 1.

Example 6.2.1
The rotor of a two-pole 3-phase induction machine rotates at 3300 (rpm), while the stator is fed
by a three-phase system of voltages at 60 (Hz). What are the possible frequencies of the rotor
voltages?

At 3300 (rpm):
2
o = 3300 = 345.6 (rad/s) while s = 377 (rad/s)
60

These two speeds can be in the opposite or the same direction such that:

r = s o = 377 345.6 (rad/s) = 722.58 (rad/s) or 31.43 (rad/s)


or fr = 115 (Hz) or fr = 5 (Hz)

1- 4
To better understand the currents induced in the squirrel cage as caused by the magnetic flux
density from the stator coils consider the simplified view of the squirrel cage in Figure 6. Adjacent
bars of the squirrel cage form a coil since the ends of the cage are electrically shorted and the largest
coils are formed by pairs of opposite side bars of the cage. The bars of the squirrel cage work
collectively to respond to any changing magnetic field. If the squirrel cage is rotating at the same
angular velocity as the stator magnetic flux density, then the loops of the squirrel cage experience a
constant magnetic field, and no current flows through the bars of the cage. Under this condition, no
magnetomotive force is generated by the squirrel cage and no torque on the rotor exists. The rotor
then begins to slow down relative to the stator magnetic field. As is does so, the magnetic flux
density for a given loop of the squirrel cage changes (some loops experiencing an increase in
magnetic flux density, and some loops experiencing a decrease in magnetic flux density). The
response to this change in magnetic field is a current that flows in the bars of the squirrel cage so as
to oppose this change in magnetic field. Since the stator magnetic field is sinusoidally distributed in
space around the rotor, and the most rapid change in this field is determined by its spatial derivative,
then as the squirrel cage slips with respect to s, the field from the rotor caused by this slip is
spatially oriented 90 with respect to the stator magnetic field (derivative of the cosine field
distribution).
For example, if the stator magnetic flux density is oriented from left to right as was shown in
Figure 5 and is rotating clockwise, then as a motor with s near zero (near synchronous speed s > 0)
the rotor is also rotating clockwise, but not quite as fast as the stator magnetic flux density. As this
slip occurs, the largest change in the field occurs from the top to the bottom of the squirrel cage or at
the zero magnetic field points of the stator field where the spatial slope of the stator field is largest.
As the slip occurs, the field through the squirrel cage, Bg in Figure 5, begins to rotate in a clockwise
B

fashion relative to the rotor. The largest change in field occurs in a direction from top to bottom of
the rotor and the coils of the squirrel cage collectively generate a counter magnetomotive force so as
to oppose the change in magnetic field. Thus current flows in the bars of the squirrel cage to generate
a counter field directed from bottom to top, or into the page on the right side, and out of the page on
the left as indicated in Figure 5.

s o

o
Figure 6. Squirrel cage isometric view.
1- 5
6.3 Torque Development
To calculate the torque on the rotor we use the following two equations:

F = Bli and T = F r (6.7)

since the flux density is perpendicular to the current producing it. For the length of the conductor, l,
we will use the depth of the rotor. The thickness, de, of the equivalent conducting sheet shown in
Figure 5 is set equal to the total area of the bars of the squirrel cage.


A = nrotor bars d 2 = 2rd e
4
(6.8)
n d2
d e = rotor bars
8r

where d is the diameter of a single bar of the squirrel cage.


For a small angle d centered at a given angle , we find the contribution to the total force and
torque as:
dF = ( JdA) Bg l
dF = ( Jde rd )Bg
dT = rdF (6.9)
2r 2ld e 2
= 2
T=
dT =
=0

B g (s o )

1
where Bg = B g sin ( ) and J ( ) = (s o )B g sin ( ) from equation (6.5) with equal to the

resistivity of the bars of the squirrel cage.
Using the relationship between flux density (or flux linkages), s, and the rotor voltage, Es, the
torque can be expressed as:
8 de 2
T = s (s o ) where s = Es (6.10)
N 2 l
s s


where the stator voltage is related to the flux density as es = s N slrBg . Focusing on the variables
2
of this equation we see the torque is proportional to the frequency of the rotor currents, (s o ) and
to the square of the flux density. This is so since the torque comes from the interaction of the flux
density, Bg, and the rotor currents, but the rotor currents are induced due to the flux, Bg, and the
relative speed (s o ) . Equation (6.10) gives torque as a function of more accessible quantities of
B B

stator induced voltage, Es, and frequency s. This comes as a result of the simple and direct
relationship between stator induced voltage, flux (or flux linkages), and frequency.

1- 6
6.4 Operation of the Induction Machine near Synchronous Speed
We already determined that the voltages induced in the rotor bars are of slip frequency,
fr = (s - o)/2. At rotor speeds near synchronous, fr is small. The rotor bars in a squirrel cage
machine possess resistance and leakage inductance, but at very low frequencies (near synchronous
speed) we can neglect this leakage inductance. The rotor currents are therefore limited near
synchronous speed by the rotor resistance only.
The induced rotor-bar voltages and currents form space vectors. These are perpendicular to the
stator magnetizing current and in phase with the space vectors of the voltages induced in the stator as
shown in Figure 7.
is,m

ir
Bg

Figure 7. Stator magnetizing current, airgap flux, and rotor currents.

These rotor currents, ir, produce additional airgap flux, which is 90 out of phase of the
magnetizing flux. The stator voltage, es, is applied externally and is proportional to and 90 out of
phase with the airgap flux. Thus additional currents, isr, will flow in the stator windings in order to
cancel the flux due to the rotor currents. These additional currents are shown in Figure 8(a), and the
resulting stator current and space vectors are depicted in Figure 8(b).

is,m is

is,r
ir Bg ir
Bg is

(a) (b)
Figure 8. Rotor and stator currents in an induction motor (a) Rotor and stator current
components (b) total stator current and space vector.

1- 7
i s,r is
es

i s,m
Bg
ir
Figure 9. Space vectors of the stator and rotor current and induced voltages.

A few items to note from the above analysis:

isr is 90 ahead of the stator magnetizing current, is,m. This means that it corresponds to
currents in the windings i1r, i2r, i3r, leading by 90 the magnetizing currents i1m, i2m, i3m.
The amplitude of the magnetizing component of the stator current is proportional to the stator
frequency, fs, and induced voltage. On the other hand, the amplitude of this component of the
stator currents, isr, is proportional to the current in the rotor, ir, which is proportional to the flux
and slip speed, r = s o, or proportional to the developed torque.
The stator current of one phase, is1, can be split into two components. One in phase with the
voltage, isr1, and one 90 behind it, ism1. The first reflects the rotor current, while the second
depends on the voltage and frequency. In an equivalent circuit, this means that isr1 will flow
through a resistor, and ism1 will flow through an inductor.
Since isr1 is equal to the rotor current (through a factor), it will be inversely proportional to
s r, or better stated as proportional to s/(s r). The equivalent circuit for the stator
shown in Figure 10 reflects these considerations.

is,1 is,r

+
is,m
es Xm RR s
s o

Figure 10. Equivalent circuit of one stator phase.

If the inductor motor is supplied with a three-phase, balanced sinusoidal voltage, then it is
expected that the rotor will develop a torque according to equation (6.10). The relationship between
speed, o, and torque near synchronous speed is shown in Figure 11. This curve is accurate as long
as the speed does not vary more than 5% around the rated synchronous speed, s.

1- 8
T

s
o

Figure 11. Torque-speed characteristics near synchronous speed.

As the speed exceeds synchronous, the torque produced by the machine is in the opposite direction to
the speed (i.e. the machine operates as a generator), developing a torque opposite to the rotation (or
counter torque) and transferring power from the shaft to the electrical system.
We already know the relationship of the magnetizing current, Ism, to the induced voltage, Esm,
through our analysis of the three-phase windings. Now we relate the currents, ir and isr to the same
induced voltage.
The current density, J , on the rotor conducting sheet is related to the air gap flux density as:

1
J= (s o ) Bg (6.11)

This current density corresponds to a space vector ir that is opposite to the isr in the stator. This
current space vector will correspond to the same current density:

1
J = isr N s (6.12)
rd

while the stator voltage es is also related to the flux density Bg. The amplitude of es is:
B


es = s N s lrBg (6.13)
2

Substitution of the relating phasors instead of space vectors in equation (6.11) we obtain:

s
Es = RR I (6.14)
s o sr

The torque is then three times (three phases) the power EsIsr over the stator speed or

E 2 1 s o 2 Pg
T =3 s = 3 s r = 3 (6.15)
s RR s RR s

1- 9
s
where = (Es/s). Here Pg is the power transferred to the resistance RR , through the
s o
airgap. Of this power a portion is converted to mechanical power represented by losses on the
o
resistance RR , and the remaining is losses in the rotor resistance, represented by the losses
s o
on resistance RR. In Figure 12 this split in the equivalent circuit is shown; note the resistance
o
RR can be negative, indicating that mechanical power is absorbed in the induction machine.
s o

is,1 is,r

+
is,m RR
es Xm RR s
s o
RR o
s o

Figure 12. Equivalent circuit of one stator phase separating the loss and torque rotor
components.

6.4.1 Example
A 2-pole three-phase induction motor is connected in a Y configuration and is fed from a 60 (Hz),
208 (V) (l-l) system. Its equivalent one-phase rotor resistance is RR = 0.1125 (). At what speed
and slip is the developed torque 28 (Nm)?

2
V 1
T = 3 s r with Vs = 120 (V)
s RR
2
120 1
28 = 3 r
377 0.1125
r = 10.364 (rad/s)
10.364
s= r = = 0.0275
s 377
o = s r = 366.6 (rad/s)

1- 10
6.5 Leakage Inductances and Their Effects
In the previous discussion we assumed that all the flux crosses the airgap and links both the stator
and the rotor windings. In addition to this flux there are flux components which link only the stator
or the rotor windings and are proportional to the currents there, producing voltages in these windings
90 ahead of the stator and rotor currents and proportional to the amplitude of these currents and their
frequency.
This is a simple model for the stator windings, since the equivalent circuit we are using is for the
stator, and we can model the effects of this flux with an inductor added to the circuit. The rotor
leakage flux can be modeled in the rotor circuit with an inductance L1s, as well, but corresponding to
o
frequency of f r = s , the frequency of the rotor currents. Its effects on the stator can be
2
modeled with an inductance L1r at frequency fs, as shown in the complete 1-phase equivalent circuit
in Figure 13.

Rs X ls Xlr
+ +
I s,1 I s,m I s,r RR
Vs Es Xm RR s
s o
RR o
s o

Figure 13. Complete equivalent circuit of one stator phase including leakage flux contributions.

Here Es is the phasor of the voltage induced into the rotor windings from the airgap flux, while
Vs is the phasor of the applied 1-phase stator voltage. The torque equation (6.15) still holds here, but
give us slightly different results. We can develop torque-speed curves, by selecting speeds, solving
the equivalent circuit, calculating power Pg, and using equation (6.15) for the torque. Figure 14
shows the stator current per phase, torque for the three phase induction machine, and the power factor
as a function of o for o given as a percentage of s and ranging from negative values to greater
than s.

1- 11
200 1
I (A)
s1

100 0.8

I (A), and T (Nm)


T (Nm)
0 0.6

pf
pf
-100 0.4
s1

-200 0.2

-300 0
-50 0 50 100 150
(percent of synchronous speed)
o

Figure 14. Stator current (single phase), torque and power factor of an induction machine
vs. speed.

6.6 Operating Characteristics


Figure 14 shows the developed torque, current, and power factor of an induction machine over a
speed range from below zero (slip > 1 or braking region) to above synchronous (slip < 0 or generator
region). There are three regions of interest:
1. For speeds in the range 0 o s the torque is of the same sign as the speed, and the
machine operates as a motor. There are a few interesting points on this curve and on the
corresponding current and power factor curves.
2. For speeds in the range o 0, torque and speed have opposite signs, and the machine is
in breaking mode. Notice the current is very high, resulting in high winding losses.
3. For speeds in the range o s the speed and torque are of opposite signs and the
machine is in the generating mode.

These regions are identified on an extended plot of torque in Figure 15. If we consider the motor
operation region 0 o s the operating point is often designed to be near, or at, the point where
the power factor is maximized. It is for this point that the motor characteristics are given on the
nameplate, rated speed, current, power factor, and torque. When designing an application it is this
point that we have to consider primarily. Will the torque suffice? Will the efficiency and power
factor be acceptable?
Starting the motor is also of interest (where slip s = 1) where the torque is not necessarily high,
but the current often is. When selecting a motor for an application, we have to make sure this starting
torque is adequate to overcome the load torque which may also include a static component. In

1- 12
addition, the starting current is often 3-5 times the rated current of the machine. If the developed
torque at starting is not adequately higher than the load starting torque, their difference, called the
accelerating torque, will be small and it may take too long to reach the operating point. This means
that the current will remain high for a long time, and fuses or circuit breakers could have their limits
exceeded.

Motor
Torque

0
Breaking Motor Generator region
region region
Generator

-100 -50 0 50 100 150 200 250


(percent of synchronous speed)
o

Figure 15. Stator torque vs. speed identifying three regions of operation.

A third point of interest is the maximum torque, Tmax, corresponding to speed Tmax. We can find
it by analytically calculating torque as a function of slip, and equating the derivative to zero. This
point is interesting, since speeds higher than this generally correspond to stable operating conditions,
while lower speeds generally correspond to unstable operating conditions. To study this point of
operation we use the Thevenin equivalent circuit for the left side of the circuit as seen looking into
terminal A-B in Figure 17.

Rs Xls A Xlr
+ +
I s,1 I s,m I s,r RR
Vs Es Xm RR s
s o
RR o
s o
B

Thevenin

Figure 16. Equivalent circuit for induction machine indicating terminals for
Thevenin equivalent circuit.

1- 13
R TH XTH Xlr
A
+ +
I s,r RR
VTH Es RR s
s o
RR o
s o
B

Figure 17. Equivalent circuit for induction machine with the stator circuit
replaced with a Thevenin equivalent circuit.

We find
( X ls + X m ) + jRs X m
RTH + jX TH = (Rs + jX ls ) ( jX m ) = (6.16)
Rs + j ( X ls + X m )
and
jX m
VTH = Vs (6.17)
Rs + j ( X ls + X m )

Then from the circuit in Figure 17, we can find Is , r as:

VTH
Is, r = (6.18)
R
RTH + R + j ( X TH + X lr )
s

where s is in the per unit system, that is s = (s o ) / s . From equation (6.15)

2 RR
VTH
3 3 2 RR 3 s
T= Pg = I s, r = (6.19)
s s s s RR
2
+ ( X TH + X lr )
2
TH
R +
s

This torque reaches a maximum as a slip, smaxT , that can be found by taking the derivative of
equation (6.19) and setting it equal to zero. When this is done we find:

RR
= (RTH )2 + ( X TH + X lr )2
smax T
or (6.20)
RR
smax T =
(RTH )2 + ( X TH + X lr )2
1- 14
This gives the maximum torque of:

2
3 VTH
Tmax = (6.21)
2s R + R 2 + ( X + X )2
TH TH TH lr

If the stator resistance is negligible, then RTH 0 and

2
T Tmax (6.22)
s s
+ max T
smax T s

If we neglect both the stator resistance and the magnetizing inductance, we can find simple equations
for Tmax and Tmax. To do so, we must assume operation near synchronous speed, where the value of
s
RR is much larger than sLlr.
s o

RR
T max s (6.23)
Llr + Lls
2 2
3V 1
Tmax s (s T max ) = 3 Vs
1
(6.24)
2 s RR 2 s Lls + Llr

Es
The slip frequency at this torque, r = s T max , for a constant flux s = is independent
s
of frequency, and is proportional to the resistance RR. We already know that RR is proportional to the
rotor resistance, so if the rotor resistance is increased, the torque-speed characteristic is shifted to the
left, as shown in Figure 18. If we have convenient ways to increase the rotor resistance, we can
increase the starting torque, while decreasing the starting current. Increasing the rotor resistance can
be easily accomplished in a wound-rotor induction machine. For the squirrel cage motor, more
complex structures such as double or deep rotor bars can be used to increase the rotor resistance.
In the formula developed we notice the maximum torque is a function of the flux. This means
that we can change the frequency of the stator voltage, but as long as the voltage amplitude changes
so that the flux stays the same, the maximum torque will also stay the same as shown in Figure 19.
This is called Constant Volts per Hertz Operation and is a first approach to controlling the speed of
the motor through its supply.
Near synchronous speed the effect of the rotor leakage inductance can be neglected, and led to the
torque-speed equation (6.15) repeated below:
E 2 1 s o 2 Pg
T =3 s = 3 s r = 3
s RR s RR s

Figure 20 shows both the exact and the approximate torque-speed characteristics. It is important to
notice that the torque calculated from the approximate equation is grossly incorrect away from
synchronous speed.

1- 15
R = 2r R =r
R 1 R 1

Torque
0 20 40 60 80 100
(percent of synchronous speed)
o

R =r
R = 2r R 1
R 1
s
I

0 20 40 60 80 100
(percent of synchronous speed)
o

Figure 18. Effect of changing the rotor resistance on the torque-speed and
current-speed characteristics of an induction motor.

20 (Hz)
60 (Hz)
Torque

45 (Hz)

-50 0 50 100 150


(rad/s)
o

60 (Hz)
s
I

45 (Hz)
20 (Hz)

-50 0 50 100 150


(rad/s)
o

Figure 19. Effect on the torque-speed characteristic of changing frequency


while keeping the flux constant.

1- 16
Linear
Approx.

Torque
40 50 60 70 80 90 100 110
(percent of synchronous speed)
o

Figure 20. Exact and approximate torque-speed characteristics.

6.7 Starting Characteristics of Induction Motors


From the above analysis we see that a particular challenge with induction motors is the high
current and low torque during starting.
A simple way to decrease the starting current is to decrease the stator terminal voltage during
startup. From equation (6.15), we see the torque is proportional to the stator voltage squared, while
the current is directly proportional to the stator voltage. If a transformer is used to decrease the stator
voltage, then both the developed torque and the line current will decease by the square of the turns
ratio of the transformer.
A commonly used method is to change the connection configuration during starting of a motor.
For example a motor designed to operate with the stator windings connected in are changed to a Y
connection during starting. The voltage ratio of these configurations is:
1
Vs,Y = Vs, (6.25)
3
then
1
I s,Y = I s, (6.26)
3
1
Ts,Y = Ts , (6.27)
3

I line Y I line

+ +
Is,Y +
Is,
Vl-l +
Vs,Y
Vl-l
Vs,

Figure 21. Y Delta starting of an induction motor.

1- 17
In the connection, I line = 3I ph , leading to:
1
I line,Y = I line, (6.28)
3

Once the machine has approached the desired operating point, we can reconfigure the connection
to , and provide better efficiency.
This decrease in current is often adequate to allow a motor to start low load starting torque.
Using a variable frequency and voltage supply we can comfortably increase the starting torque, as
shown in Figure 19, while decreasing the starting current.

6.8 Multiple Poles


If we consider that an induction machine will operate close to synchronous speed {3000 (rpm) for
50 (Hz) and 3600 (rpm) for 60 (Hz)} we may find that the speed of the machine is too high for an
application. If we recall the pictures of the flux in AC machines we have shown before, we notice the
flux has a relatively long path to travel in the stator. This makes the stator heavy and lossy.
In machines with more than one pair of poles, the sinusoidal distribution for the windings covers
a smaller angle. For example, in a 4 pole machine each side of a sinusoidally distributed winding of
one phase covers only 90 instead of 180 (as was the case for a 2 pole machine).
Figure 22 shows at one instant the equivalent windings resulting from rotor windings of a 2-pole,
4-pole, and 6-pole machines.

Figure 22. Multipole induction machines 2-pole, 4-pole, 6-pole rotor windings
(stator windings not shown).

The effects of a large number of poles on the operation of the machine are not difficult to predict.
If the machine has p poles, or p/2 pole pairs, then in one period of the voltage, the flux will travel
2s/p (rad/s). This leads to the rotor speed corresponding to synchronous of sm:
2
sm = s (6.29)
p

We now introduce the actual mechanical speed of the rotor as m, while we keep the term o as
the rotor speed of a two pole motor. We generally measure m in (rad/s), while we measure o in
(electrical rad/s). We retain the same definition for slip based on the electrical speed o.

1- 18
2
m = o (6.30)
p
p
o s 2 m
s= s = (6.31)
s s

This shows that for a 4-pole machine, supplied by a 60 (Hz) source, and operating close to rated
conditions, the speed will be near 1800 (rpm), while for a 6-pole machine, the speed will be near
1200 (rpm). While increasing the number of poles results in a decrease of the synchronous and
operating speeds of the machine, it also results in an increase of the developed torque of the machine
by the same ratio. Hence, the corrected torque formula will be:
p Pg p Pm
T =3 =3 (6.32)
2 s 2 o
Similarly, the torque near the synchronous speed is:

p Es2 1 s o p 2sr p Pg
T =3 =3 =3 (6.33)
2 s RR s 2 RR 2 s
while the previously developed formulas for maximum torque will become:
2
p 1 VTH
Tmax = 3 (6.34)
2 2 s R + R 2 + ( X + X )2
TH TH TH lr

and
2 2
3 p Vs 1
Tmax (s T max ) = 3 p Vs 1
(6.35)
2 2 s RR 2 2 s Lls + Llr

Example 6.8.1
A 3-phase, 2-pole induction motor is rated at 190 (V), 60 (Hz), is connected in the Y
configuration, and has RR = 6.6 (), Rs = 3.1 (), XM = 190 (), Xlr = 10 (), Xls = 3 ().
Calculate the motor starting torque, starting current and starting power factor under rated voltage.
What will be the current and power factor if no load is connected to the shaft?

1. At starting, s = 1
190
Is = 3 = 7.06 /-54.5 (A)
{[3.1 + j3] + j190 (6.6 + j10)}
j190
Ir = Is = 6.7 /-52.6 (A)
6.6 + j10 + j190

T =3
Pg p
=3
( )
6.7 2 6.6 2
= 2.36 (Nm)
s 2 377 2

1- 19
2. Under no load, the speed is synchronous and s = 0

190
Is = 3 = 0.57 /-89.1 (A)
{3.1 + j3 + j190}
I s = 0.57 (A)
pf = 0.016 lagging

Example 6.8.2
A 3-phase, 2-pole induction motor is rated at 190 (V), 60 (Hz), is connected in the Y
configuration, and has RR = 6.6 (), Rs = 3.1 (), XM = 190 (), Xlr = 10 (), Xls = 3 (). It is
operating from a variable speed variable frequency source at a speed of 1910 (rpm), under a
constant (V/f ) policy and the developed torque is 0.8 (Nm). What is the voltage and frequency
of the source? (Hint: First calculate the slip).
1
190
The ratio (Vs/s) stays at 3 .
377

2
p V 1
T = 3 s r
2 s RR
2
110 1
0.8 = 1 3 r r = 20.65 (rad/s)
377 6.6
p
s = m + r = 220.01 + 20.65 = 220.66 (rad/s)
2
110
f s = 35 (Hz) Vs = 220.66 = 64.4 (V) or 110 (Vl -l )
377

Example 6.8.3
A 3-phase, 4-pole induction machine is rated 230 (V), 60 (Hz). It is connected in the Y
configuration, and has RR = 0.191 (), LM = 35 (mH), Lls = 1.2 (mH). It is operated as a generator
connected to a variable frequency/variable voltage source. Its speed is 2036 (rpm), with a
counter-torque of 59 (Nm). What is the efficiency of this generator? (Hint: Here power in is
mechanical, power out is electrical; also first calculate the slip).

Although we do not know the voltage or the frequency, we know their ratio is (132.8/377).
2
p V 1
T = 3 s r
2 s RR
2
132.8 1
59 = 3 2 r
377 0.191
r = 15.14 (rad/s)

1- 20
Now the synchronous speed can be found by adding slip and rotor speeds as:

p 2 2036
s = m + r = 2 15.14 = 411.3 (rad/s)
2 60
132
f s = 65.5 (Hz) Vs = 65.5 = 144 (V)
60
We have to calculate the impedances of the equivalent circuit for the frequency of 65.5 (Hz):
X m = 35 10 3 411.3 = 14.4 ()
X ls = 0.49 ()
X lr = 0.617 ()
then
p
r + m
RR 2 = 5.38 ()
r
144
Is = = 30 /-148 (A)
[0.2 + j 0.49 + j14.4][0.191 5.38 + j 0.617]
Ir = 27.2 /-166.9 (A)

Notice that with generation operation RR < 0. We can calculate now the losses, etc.:

Pm = 3 27.2 2 5.38 = 11.941 (kW)


Protor , loss = 3 27.2 2 0.191 = 423 (W)
Pstator , loss = 3 30 2 0.2 = 540 (W)
Pout = Pm Protor , loss Pstator , loss = 10.98 (kW)
P
= out = 0.919
Pm

1 A. E. Fitzgerald, C. Kingsley, Jr., S. D. Umans, Electric Machinery, 6th edition, McGraw-Hill, New York, 2003.

1- 21
ECE 320
Energy Conversion and Power Electronics
Dr. Tim Hogan

Chapter 7: Synchronous Machines and Drives


(Textbook Chapter 5)

Chapter Objectives
For induction machines, as the rotor approaches synchronous speed, the frequency of the currents
in the rotor decreases, as does the amplitude of these currents. The reason an induction motor
produces no torque at synchronous speed is not that the currents are DC, but that their amplitude is
zero.
It is possible to operate a three-phase machine at synchronous speed if DC is externally applied to
the rotor and the rotor is rotating at synchronous speed. In this case torque will be developed only at
this speed, i.e. if the rotor is rotated at speeds other than synchronous, the average torque will be zero.
Machines operating on this principle are called synchronous machines, and cover a great variety.
As generators they can be quite large, rated a few hundred MVA, and almost all power generation is
done using synchronous machines. Large synchronous motors are not very common, but can be an
attractive alternative to induction machines for some applications. Small synchronous motors with
permanent magnets in the rotor, rather than coils with DC, are rapidly replacing induction motors in
automotive, industrial and residential applications since they lighter and more efficient.

7.1 Design and Principle of Operation


The stator of a synchronous machine is of the type that we have already discussed with three
windings carrying a three-phase system of currents. As we saw in Chapter 5, this results in a stator
magnetic field that spatially rotates around the stator at a constant angular speed, s. Unlike
induction machines, synchronous machines have zero slip, and the rotor maintains the same angular
speed, s, as the stator generated field. In a synchronous machine, the rotor windings carry DC
current, or are composed of permanent magnets as discussed next.

7.1.1 Wound Rotor Carrying DC


In this case the rotor steel structure can be either cylindrical, that that in Figure 1(a), or salient
like the one in Figure 1(b). In either case, the rotor winding carries DC, delivered through slip rings,
or through a rectified voltage of an inside-out synchronous generator mounted on the same shaft. In
this handout, discussion is limited to cylindrical rotors.

7.1.2 Permanent Magnet Rotor


Instead of supplying DC to the rotor, the rotor contains permanent magnets in configurations such
as those shown in Fig. 2. The effects of permanent magnet rotors include:
The rotor flux can no longer be controlled externally. It is defined by the magnets and the
geometry
The machine becomes simpler to construct, at least for small sizes.

1- 1
Sinusoidally
distributed
winding

Concentrated
winding

(a) (b)

Figure 1. Wound rotor configurations of synchronous machines.

N N

S S S S
S S
N N N N

N S S N N S S N

N N N N
S S
S S S S
N N
(a) Surface PM (SPM) synchronous (b) Surface inset PM (SIPM)
machine synchronous machine

N
N S
N S
S S
S
N N S N
S N
N S S N
N S
N N N S
S
S S
S N
N S N

(c) Interior PM (IPM) synchronous (d) Interior PM with circumferential


machine orientation synchronous machine

Figure 2. Possible magnet placements in PMAC motors.

1- 2
7.2 Equivalent Circuit
The flux in the air gap can be considered to be due to two sources: the stator currents, and the
rotor currents or permanent magnet. Recall the stator currents produce a flux that rotates at
synchronous speed, and that this flux could also be produced by one equivalent winding that is
rotating at synchronous speed and carrying a current equal to the magnitude of the stator-current
space vector.
The rotor is itself such a winding, a real one, sinusoidally distributed, carrying DC and rotating at
synchronous speed. It produces an airgap flux, which could also be produced by an additional set of
three phase stator currents, giving a space vector iF. The amplitude of this space vector would be:
NS
iF = if (7.1)
NR
where NS is the number of the stator turns of the one equivalent winding (when the three stator
windings are represented by a single equivalent rotating winding as discussed above) and NR is the
number of the turns in the rotor winding. Its angle R would be the same as the angle of the rotor
position:
R = st + R 0 (7.2)
The stator current space vector has amplitude:
3
is = 2I s (7.3)
2
where Is is the rms current of one phase. The stator current space vector will have an instantaneous
angle, is, of
is = s t + is 0 (7.4)

The airgap flux is then produced by both these current space vectors (rotor and stator). This flux
induces in the stator windings a voltage, es. In quasi steady-state everything is sinusoidal and the
voltage space vector corresponds to three phase voltages E1, E2, E3. In this case we can create an
equivalent circuit for the stator as shown in Figure 3. Here IF is the stator AC current, that if it were
to flow in the stator windings, it would have the same effects as the rotor current, if. In our analysis
we can use as reference either the stator voltage, Vs, or the stator current, Is as shown in Figure 4.
From the relationship of the space vectors in Figure 4, the angle is the power factor angle (the angle
between Is and Vs). We call , the angle between Vs and IF, and the power angle, , is between IF
and Is.

Is

+ IM

Vs Es jXM IF = I F

Figure 3. Stator equivalent circuit for a synchronous machine.

1- 3
Is


Vs


IM
IF
Is

Figure 4. Space vector diagram of a synchronous machine.

From Figure 3 and Figure 4 we can note the following two relationships:
The space vector of the voltages induced in the stator, es, is 90 ahead of the magnetizing
current space vector, IM. This can be understood by the fact that IM is what causes all the
airgap flux that links the stator and induces es. For a given frequency, the amplitude of this
voltage, es, is proportional to the current IM.
A permanent magnet machine can be considered equivalent to that with a winding, carrying a
direct current, if, that is constant (and cannot be controlled).

7.3 Operation of the Machine Connected to a Bus of Constant Voltage and Frequency
This is usually the case for large synchronous generators or motors. We can consider any bus as
one of constant voltage, by making a few modifications to the equivalent circuit as shown in Figure 5.

1- 4
jXS Is

+ IM

Vs IF = I F
jXM

jXS jXM
+ Is
+
Vs VF = jXM IF

Is

+ IM
jXM
Vs IF = I
j(XM+ XS ) j(XM+ XS ) F

Figure 5. Equivalent circuits that account for a non-zero bus impedance in the system.

Synchronous machines are very efficient, and most of the time we can neglect the stator
resistance. All power then is converted to mechanical power and:
2
P = 3Vs I s cos = Ts (7.5)
p
P = 3Vs I F cos (7.6)
IM = Is + IF (7.7)
Vs = jX M I M (7.8)

In this operation, Vs and s (and therefore speed) remain constant. The only input variables are
2
the torque, T, which affects output power, Pout = Ts , and the field current, if, which is
p
proportional to IF; the magnetizing current IM is constant, since it is tied to the voltage Vs.
Assuming the machine is operated so the power to it varies while the frequency and field current
remain constant. Since this is a synchronous machine, the speed will not vary with the load. From
equation (7.6) we can see the power, and therefore the torque, varies sinusoidally with the angle .
Remember that is the angle between the axis of the rotor winding, and the stator voltage space
vector. Since this voltage space is 90 ahead of the space vector of the magnetizing current ( 90)
1- 5
is the angle between the rotor axis and the magnetizing current space vector (same as the airgap flux).
When there is no torque this angle is 0, i.e. the rotor rotates aligned with the flux, when external
torque is applied to the rotor in the direction of rotation the rotor will accelerate. As it accelerates
(with the flux rotating at constant speed) the flux falls behind the rotor, and negative torque is
developed, making the rotor slow down and rotate again at synchronous speed, but now ahead of the
flux.
Similarly, when load torque is applied to the rotor, the rotor decelerates; as it does so, the angle
decreases beyond -90, i.e. the rotor falls behind the flux. Positive torque is developed that brings
the rotor back to synchronous speed, but now rotating behind the stator flux.
In both cases when the load torque on a motor or the torque on a motor or the torque of the prime
mover in a generator increases beyond a maximum, corresponding to cos = 1, the machine cannot
develop adequate torque and it loses synchronization.
Let us now discuss the effect of varying the field current while keeping the power constant. From
equation (7.6), when power and voltage are kept constant, the product I F cos remains constant as
well, but this product is the projection of IF on the horizontal axis. This means that as the field
current changes while power stays the same, the tip of IF travels on a vertical line, as shown in Figure
7(a). Similarly equation (7.5) means that at the same time the tip of Is travels on another vertical line,
also shown in Figure 7(a).
It is clear from Figure 7(a) that once the field current has exceeded a value specific to the power
level, the power factor becomes leading and the machine produces reactive power. This is different
from the operation of an induction machine, which always absorbs reactive power.
When the machine operates as a generator, the input power is negative. Figure 7(b) shows this
operation for both leading and lagging load power factor. Here the angle between stator voltage and
stator current defined in the direction shown in the equivalent circuit (Figure 5) is outside the range
-90 < < 90.

Motor Generator

/2
Torque

0
/2


Figure 6. Torque and angle in a synchronous motor.

1- 6
I s1 I s3

I s2
I s2
Vs Vs

I s1
I s3
IF3

IF1
IM
IM
IF2
IF2
I s1
IF1
IF3

(a) (b)
Figure 7. Equivalent circuits that account for a non-zero bus impedance in the system.

Example 7.3.1
A 3-phase Y-connected synchronous machine is fed from a 2300(V), 60(Hz) source. The ratio of
the AC stator equivalent current to the rotor DC is IF/if = 1.8. The magnetizing inductance of the
machine is 200(mH).

The machine is operated as a motor and is absorbing 110 (kW) at 0.89 pf leading. Calculate the
required field current and the load angle. Draw the corresponding phasor diagrams.

Using Figure 8:
Is

27.1
Vs

-131

IM
IF

Figure 8. Synchronous machine as a motor.

We find:
X M = 2 60 0.2 = 75.4 ()

1- 7
2300
Vs = = 1328 (V)
3
110 103

3
Is = / 27.1 = 31/ 27.1 (A)
1328 0.89
using the stator voltage we can calculate IM and from it IF.
1328
IM = = 17.62 / -90 (A)
75.4

I F = I M I s = 42 / -131 (A)

I
i f = F = 23.4 (A)
1.8

Repeating this process for operation as a generator at 110 (kW), 0.82 pf leading gives Figure 9

IG

35 IF
Vs
3.56

-145

Is IM

Figure 9. Synchronous machine as a generator.

I G = 33.7 / 35 (A) I s = 33.7 / -145 (A)

I F = I M I s = 27.66 / 3.56 (A)

I
i f = F = 15.37 (A)
1.8

What is the maximum power the above machine can produce (or absorb when operating as a
generator and at the field current just calculated?
1- 8
We know that absorbed power is:
P = 3Vs I F cos

for if = 15.37 (A) we have IF = 27.66 (A), and P generated is maximum for = 0, thus:

P = 3 1328 27.66 = 110.2 (kW)

If the terminal voltage remains at 2300 (V), 60 (Hz), what is the minimal field current required to
maintain operation as a motor with load 70 (kW)?

P = 3 Vs I F cos = 3 1328 I F = 70 103 (W)


or
I F = 17.57 (A)

7.4 Operation from a Source of Variable Frequency and Voltage


This operation requires that our synchronous machine is supplied by an inverter. The operation
now is entirely different than before. We no longer have an infinite bus, but rather whatever stator
voltage or current and frequency we desire. Moreover, with a space-vector controlled inverter, the
phase of this voltage or current can be arbitrarily set at any instant, i.e. we can define the stator
current or voltage space vector, and obtain it at will. The considerations for the motor operation are
also different:

There is no concern for absorbing or supplying reactive power. Instead, there is a limit on the
total stator current, determined by thermal considerations.
There is a limit to the maximum voltage the source can supply, which leads to modifications of
the machine mode of operation at high speeds.

Operation from a source of variable frequency and voltage is most common for permanent
magnet machines, where the value of |IF| is constant.
In simple terms, when the machine is starting as a motor the frequency applied should be zero,
but the voltage space vector should be of such angle with respect to the rotor that torque is developed
as discussed in the previous section. As torque develops, the machine accelerates, and the applied
stator currents have to create a rotating space vector leading the rotor flux. Voltage and frequency
have to be increased, so that this torque is maintained. It is important therefore to monitor the
position of the rotor in order to determine the location of the stator current or voltage space vector.
Two possible control techniques are implemented: either voltage control, where the stator voltage
space vector is determined and applied, or current control, where the stator current space vector is
applied.
For a fixed stator voltage and power (and torque) level, the stator losses are minimal when the
stator voltage and current are in phase. Figure 10 shows this condition.

1- 9
Is
Vs

IF
IM

Figure 10. Operation of a synchronous PM drive at constant voltage and frequency.

Notice that as the power changes with the voltage constant two things happen:

1. The voltage space vector varies in amplitude and the magnetizing current changes with it.
2. The amplitude of IF stays constant, but its angle with respect to the voltage changes.

From the developed torque and speed we can calculate the frequency, the values of IM and Is, and
the angle between the stator voltage space vector and the rotor, since:
3p 3p
T= Vs I s = LM I M I s (7.9)
2s 2

I F2 = I M
2
+ I s2 (7.10)
where IF is a constant in permanent magnet machines.
More common is the case when the stator voltage is not constant. Here we monitor the position
of the rotor and since the rotor flux and rotor space current are attached to it, we are actually
monitoring the position of IF. To make matters simple we use this current rather than the stator
voltage as reference, as shown in Figure 11.

jIs XM
jIF XM
Vs

Is IM


IF

Figure 11. Operation of a synchronous PM drive below base speed.

Although the previous equations for power and torque are still true, they are not as useful. New
formulae can be created that have the stator current, Is, and magnetizing current, IM, as variables. We

1- 10
also make use of the angle , between IF and Is, since we can control it. Starting from what we
already know:

[ ] [
Pg = Re VsI*F = Re jX M (I F + I s )I*F ] (7.11)

= Re[ jX I I ]+ Re[ jX
*
M F F
*
M IsIF ] (7.12)

= X Im[I I ] = X I I
M
*
s F M s F sin (7.13)

For a given torque, minimum losses require minimum value of the stator current. To minimize the
value of Is with constant power and IF we choose = 90 to obtain:

[ ]
Pg = X M Im I sI*F = X M I s I F (7.14)

= 3 L Im[I I ] = 3 L
p Pg p p*
T =3 M s F M IsIF (7.15)
2 s 2 2
which means that for constant power, the projection of the stator current on an axis perpendicular to
IM is constant.
As the rotor speed increases, even if IM stays constant, the stator voltage Vs = sLMIM increases.
At some base speed, sB, the required voltage exceeds the maximum that the power source can
provide. To increase the speed beyond the base speed, we no longer keep = 90. On the other hand,
at that speed we know that the voltage has reached its upper limit, Vs = Vs,max, therefore the value of
IM = Vs,max/XM is known. In this case equations (7.9) and (7.10) become:
3p 3p
T= Vs I s cos = LM I M I s cos (7.16)
2s 2

I F2 = I M
2
+ I s2 + 2 I M I s sin (7.17)

Figure 12 shows such an operation with the variable having the subscript 1. Note that we
calculate the torque from power.
P = 3 X M I s I F sin (7.18)

T=
P p
s 2
p
2
[ ]p
= 3 LM Im I s I*F = 3 LM I s I F sin
2
(7.19)

1- 11
jIs1 XM
jIF XM
Vs1
IM2
Is2 Is1 IM1
1
jIs2 XM 1 2
2
Vs2 IF

Figure 12. Field weakening of a PM AC motor. The two diagrams are at the same
frequency, but the second one has > 90 and a lower Vs.

Example 7.4.1
A permanent magnet, Y connected, three-phase, 2-pole motor has IF = 40 (A), and
XM = 0.9 () at 100 (Hz).

1. If it is absorbing P = 1.5 (kW) at 100 (Hz) with minimum stator current Is, calculate this
current, the angle between Is and IF, the speed, the stator voltage (line-neutral) and the
power factor.

The minimum current Is will exist when the angle between Is and IF, , is 90. Under this
condition:

1500
P = 3X M Is I F or Is = = 13.89 (A)
3 0.9 40
I M = I F + I s = 40 + 13.89 /90 = 42.34/19.15 (A)
Vs = j s LM I M = 38.12 /109.15 (V)

where s = 2 100 = 628 (rad/s) or 6000 (rpm)

The power factor is then calculated as:

pf = cos(109.14 90 ) = 0.946 lagging

2. It is desired to increase the motor speed to 6900 (rpm) while keeping the power the same,
P = 1500 (W), but the supplied voltage has reached its upper limit of Vs = 38.12 (V). Now
the motor absorbs the same power at the voltage calculated in the previous question, but at
frequency of 115 (Hz). This can be accomplished by having the stator current no longer at
a minimum value and g no longer at 90. Calculate again the angle between Is and IF, the
speed, and the power factor.

1- 12
1500
P = 3Vs I F cos( ) or cos( ) = = 0.32
3 38.12 40
= 109.14 and Vs = 38.12 /109.14 (V)
V 38.12 / 109.15
IM = s = = 36.83 /19.15 (A)
jX M 115
j 0.9
100
I s = I M I F = 13.16 /113.18 (A)

s = 2 115 = 722.57 (rad/s) or 6900 (rpm)

The power factor is then calculated as:

pf = cos(109.14 113.18 ) = 0.997 leading

jIs XM
jIF XM
Vs
IM
I
s

IF
Case #1

jIs XM jIF XM

Vs

IM
Is

IF
Case #2

1- 13
Example 7.4.2
A permanent magnet machine has 4-poles, IF = 40 (A), and the value of the magnetizing
reactance is XM = 2.9 () at 100 (Hz). The maximum stator current is Is,max = 50 (A) and the
maximum voltage the inverter can provide is Vln,max = 200 (V).

First calculate the maximum speed the machine can operate with minimal stator Joule losses.
Xm
Lm = = 4.6 (mH)
2 100
For minimal losses sin() = 1, so:

T =3
p
2
[
4
]
Lm I s I F sin ( ) = 3 4.6 103 (H ) [50(A )][40(A )](1) = 55 (Nm)
2

Then calculate the maximum speed the machine can reach for operation as above:

I m = Is + IF = 64(A )

Vs 200(V )
s, max1 = =
[
Lm I m [64(A )] 4.6 103 (H ) ]
= 679 (rad/s)

and the maximum mechanical speed under these conditions will be:

2
mech, max1 = s, max1 = 339.5 (rad/s)
p

Next assume operation above this speed, mech,max1, for example at 400 (rad/s). Obviously this
cannot be done with minimal stator Joule losses, so the field has to be weakened by increasing
above 90. Calculate the maximum torque that can be developed.

We now have the values of Vs and of Is, since they correspond to the limits, Vs = 200 (V),
Is = 50 (A). From the voltage and speed we can find

p
Vs = [400(rad/s )]Lm I m = 200(V )
2

200
Im = = 54.35(A )
800 4.6 103

From the triangle of Is, Im, IF (case 2 in the last example):

I s2 = I F2 + I m
2
2 I F I m cos (/IFIm)
50 2 = 40 2 + 54.352 2 40 54.35 cos (/IFIm)
(/IFIm) = 61.8
b = 61.8 + 90 = 151.8

1- 14
p
T = 3 I m I F cos( ) = 52.8(N m )
2

Note that the demagnetizing component of Is is

I sd = I s sin ( ) = 23.6 (A )
which might be too high.

7.5 Controllers for Permanent Magnet AC Machines


Figure 13 shows a typical controller for an AC Machine. It requires a DC power supply, usually a
rectifier fed from an AC source, an inverter and a controller. Figure 14 shows the controller in
slightly higher detail.

Variable
voltage and
3-phase frequency
AC DC
Rectifier Inverter Motor
Link

Voltage or
Current
Command

Controller
Speed or Torque
Command

Figure 13. Generic Controller for a PMAC Machine.

PMAC
Motor Optical
*
T Is encoder
PI Calculate
+ Calculate
Inverter
Controller Is * * *
_ isa , isb , isc

Position of Is
+

+
Rotor position

1
s
Rotor speed

Figure 14. Field Oriented controller for a PMAC Machine. The calculations for Is are based on
* * *
equation (7.19) and the calculation of isa , isb , isc are calculated from the space vector Is
from equations (5.3).
1- 15
7.6 Brushless DC Machines
While it would be difficult to find the differences between a PM AC machine described above and a
brushless DC machine by just looking at them, the concept of operation is quite different as is the
analysis. The windings in the stator in a brushless DC machine are not sinusoidally distributed but
instead they are concentrated, each occupying one third of the pole pitch. The flux density on the
magnet surface and in the airgap is also not sinusoidally distributed over the magnet but almost
uniform in the air gap.
As the stator currents interact with the flux coming from the magnet torque is developed. It
should be clear that for the same direction of flux, currents in opposite directions result in opposite
forces, and therefore in reduction of the total torque. This in turn makes it necessary that all the
current in the stator above the rotor is in the same direction. To accomplish this, the following are
needed:
Sensors on the stator that sense the direction of the flux coming from the rotor,
A fast power supply that will provide currents to the appropriate stator windings as determined
by the flux direction
A way to control these currents, e.g. through Pulse Width Modulation
A controller with inputs of the desired speed, the flux direction in the stator, and the stator
currents; and outputs of the desired currents in the stator

Figs show the rotor positions, the stator currents and the switches of the supply inverter for two
rotor positions.

b'

rth
c

a
No
a
b
c th
Sou

c'
a'

b
Figure 15. Energizing the windings in a brushless DC motor.

b'
c

a
North

b
South

c
c'
a'

b
Figure 16. Energizing the windings in a brushless DC motor, a little later.

1- 16
The operation of the system can be described simply since the developed torque is proportional to
the stator currents as:
T = k Is (7.20)

At the same time, the rotating flux induces a voltage in the energized windings:

E = k (7.21)

Finally the terminal voltage differs from the induced voltage by a resistive voltage drop:

Vterm = E + I s R (7.22)

The equations are similar to those for a DC motor (4.6), (4.7), (4.8). This is the reason that although
this machine is entirely different from a DC motor, it is commonly referred to as a brushless DC
motor.

1- 17
ECE 320
Energy Conversion and Power Electronics
Dr. Tim Hogan

Chapter 8: Power Electronics


(Textbook Chapter 10, and Sections 11.2, 11.3, and reserve book: Power Electronics)

Chapter Objectives
As we saw in the last chapter, control over the torque and speed of the motor can be gained through
voltage and frequency control to the motor. This can be accomplished by converting the input AC source
power to a DC source (rectifying it), then filtering it to reduce harmonics, and finally converting it back to
an AC source having the desired frequency and amplitude (inverter).

8.1 Line Controlled Rectifiers


We start with a description of how to draw power from a 1-phase or 3-phase system to provide DC to
a load. The characteristics of the systems here include that the devices used will turn themselves off
(commutate) and that the systems draw reactive power from the loads.

8.1.1 One-Phase and Three-Phase Circuits with Diodes


If the source is 1-phase, a diode is used and the load is purely resistive, as shown in Figure 1 then it is
a relatively simple configuration. When the source voltage is positive, the current flows through the diode
and the voltage of the source equals the voltage of the load.

v,v
s d

vdiode i
+ i

+
+ v
diode t
vs vd R

v,v
s diode

Figure 1. Simple circuit with diode and resistive load.

If the load includes an inductance and a source (such as a battery we wish to charge), as in Figure 2,
then the diode will continue to conduct even when the load voltage becomes negative as long as the
current is maintained. This comes from the characteristics of the inductor:
1 t3
L 0
v L dt = i (t3 ) i (0) = 0 (8.1)

Thus, the shaded area A in Figure 2 must equal the shaded area B.

vdiode vL i
+ +
+ L
+
+
vs vd Ed

00

Ed

vs i
0 t
t1 t2 t3

00

0V
t

0V

vdiode
0V
0s
A 5ms 10ms 15ms 20ms 25ms

vL
0V
t

0V
B

Figure 2. Simple circuit with diode and inductive load with voltage source.

8.1.2 One-Phase Full Wave Rectifier


More common is a single phase diode bridge rectifier such as shown in Figure 3. The load can be
modeled with one of two extremes: either as a constant current source, representing the case of a large
inductance that keeps the current through it almost constant, or as a resistor, representing the case of
minimum line inductance. We will study the first case with AC and DC side current and voltage
waveforms shown in Figure 4 for the ideal case of Ls = 0.
id
+

is

Ls
+
vs Cd vd

Figure 3. One-phase full wave rectifier.

0
vs

is
Id
0
t

0
0 5 10 15 20 25

vd
0
id = Id

0
t

Figure 4. Waveforms for a one-phase full wave rectifier with inductive load.
If we analyze these waveforms, the output voltage will have a DC component, Vdo (where the
subscript o represents that this is the ideal case with Ls = 0):
2
Vdo = 2Vs 0.9Vs (8.2)

where Vs is the RMS value of the input AC voltage. On the other hand the RMS value of the output
voltage will be
Vs = Vd (8.3)
containing components of higher frequency.
Similarly on the AC side the current is not sinusoidal, rather it changes abruptly between Id and Id.
2
I s1 = 2 I d 0.9 I d (8.4)

and again, the RMS values are the same
Id = Is (8.5)
The total harmonic distortion, THD, can then be found as:

I s2 I s21
THD = 48.43% (8.6)
I s1
It is important to note here that if the source has some inductance (and it usually does), then
commutation will be delayed after the voltage reaches zero, until the current has dropped to zero as shown
in Figure 5. This will lead to a decrease of the output DC voltage below what is expected from (8.2)
id

is

Ls
+
Id
vs vd

0V

vd

0V t

vs

0V
0V
16 20 24 28 32 36 40

vL

0V t

0V

is

A t

Figure 5. One-phase full wave rectifier with inductive load and source inductance.
8.1.3 Three-Phase Full Diode Rectifiers

The circuit of Figure 3 can be modified to handle three phases, by using 6 diodes as shown in Figure 6.
Figure 7 shows the AC side currents and DC side voltage for the case of high load inductance. Similar
analysis as before shows that on the DC side, the voltage is:
3
Vdo = 2Vl l = 1.35Vl l (8.7)

From Figure 6, we see that on the AC side, the RMS current, Is is:
2
Is = I d = 0.816 I d (8.8)
3
while the fundamental current, i.e. the current at power frequency is:
1
I s1 = 6 I d = 0.78I d (8.9)

Again, inductance on the AC side will delay commutation, causing a voltage loss, i.e. the DC voltage
will be less than that predicted by equation (8.7).
id

ia
+ a

Ls
+ b R
n Cd vd

+ c

Figure 6. Three-phase full-wave rectifier with diodes.


vd

0V
Vdo
van vbn vcn

0V t

0V
0s 10ms 20ms 30ms 40ms

ia
D4 120
0A t
D1 120
60
ib
D5
A t
D2

D6
ic
0A t
D3

Figure 7. Waveforms of a three-phase full-wave rectifier with diodes and inductive load.

8.1.4 Controlled Rectifiers with Tyristors

Thyristors give us the ability to vary the DC voltage. Remember that to make a thyristor start conducting,
the thyristor must be forward biased and a gate pulse provided to its gate. Also, to turn off the thyristor
the current through it must reverse direction for a short period of time, trr, and return to zero.

8.1.5 One-Phase Controlled Rectifiers

Fig. shows the same 1-phase bridge we have already studied, now with thyristors instead of diodes, and
fig shows the output voltage and input current waveforms. In this figure is the delay angle,
corresponding to the time we delay triggering the thyristors after they became forward biased.
+
T1 T3

is

+
Id
vs vd

T4 T2

Figure 8. One-phase full wave converter with thyristors.

Thyristors 1 and 2 are triggered together and of course so are 3 and 4. Each pair of thyristors is turned off
immediately (or shortly) after the other pair is turned on by gating. Analysis similar to that for diode
circuits will give:
2
Vdo = 2Vs cos( ) = 0.9Vs cos( ) (8.10)

and the relation for the currents is the same

2
I s1 = 2 I d cos( ) = 0.9 I d cos( ) (8.11)

It should also be pointed out that in Figure 9 the current waveform on the AC side is offset in time with
respect to the corresponding voltage by the same angle , hence so is the fundamental of the current,
resulting in a lagging power factor.
vd

is

=0 vs

00V

vd

0V t

>0
vs

00V
14 16 20 24 28 32 35

100

is

0 t

vs
-100

Figure 9. Waveforms of one-phase full wave converter with thyristors.

On the DC side, only the DC component of the voltage carries power, since there is no harmonic
content in the current. On the AC side the power is carried only by the fundamental, since there are no
harmonics in the voltage.
P = Vs I s1 cos( ) = Vd I d (8.12)

8.1.5.1 Inverter Mode


If the current on the DC side is sustained even if the voltage reverses polarity, then power will be
transferred from the DC to the AC side. The voltage on the DC side can reverse polarity when the delay
angle exceeds 90, as long as the current is maintained. This can only happen when the load voltage is as
shown in Figure 10, e.g. a battery.
id L

+
T1 T3

is

+
Ed
vs vd +

T4 T2

Figure 10. Operation of a one-phase controlled converter as an inverter.

8.1.6 Three-Phase Controlled Converters

As with diodes, only 6 thyristors are needed to accommodate three phases. Figure 11 shows the
schematic of the system, and Figure 12 shows the output voltage waveforms.

id

+
T1 T3 T5
ia
+ a

+ b ia Id
n vd

+ c ia

T4 T6 T2

Figure 11. Schematic of a three-phase full-wave converter based on thyristors.


vd
00
Vdo
van vbn vcn
ia
0 t

00
15ms =2 0 m0s 25ms 30ms 35ms

vd
Vd
van vbn vcn
ia
t

s 16ms >0 20ms 24ms 28ms 32ms

Figure 12. Waveforms of a three-phase full-wave converter based on thyristors.

The delay angle is again measured from the point that a thyristor becomes forward biased, but in this
case the point is at the intersection of the voltage waveforms of two different phases. The voltage on the
DC side is then (the subscript o here again meaning Ls = 0):
3
Vdo = 2Vl l cos( ) = 1.35Vl l cos( ) (8.13)

while the power for both the AC and DC side is
P = Vdo I d = 1.35Vl l I d cos( ) = 3 Vl l I s1 cos( ) (8.14)
which leads to:
I s1 = 0.78 I d (8.15)

and the relationship between Vdo and Vd in Figure 12 is:

Vd = Vdo cos( ) (8.16)

Again, if the delay angle is extended beyond 90, the converter transfers power from the DC side to
the AC side, becoming an inverter. We should keep in mind, though that even in this case the converter is
drawing reactive power from the AC side.
8.1.7 Notes
For both 1-phase and 3-phase controlled rectifiers, a delay in creates a phase displacement of the
phase current with respect to the phase voltage, equal to . The cosine of this angle is the power
factor of the fundamental harmonic.
For both motor and generator modes the controlled rectifier absorbs reactive power from the three-
phase AC system, although it can either absorb or produce real power. It also needs the power line
to commutate the thyristors. This means that inverter operation is possible only with the rectifier is
connected to a power line.
When a DC motor or a battery is connected to the terminals of a controlled rectifier and becomes
greater than 90, the terminal DC voltage changes polarity, but the direction of the current stays the
same. This means that in order for the rectifier to draw power from the battery or a motor that
operates as a generator turning in the same direction, the terminals have to be switched.

8.2 Inverters
Here we study systems that can convert DC to AC through the use of devices that can be turned on and
off such as GTOs, BJTs, IGBTs, and MOSFETs, which allows the transfer of power from the DC source
to any AC load, and gives considerable control over the resulting AC signal. The general block diagram
of the complete system is shown in Figure 13.

+
60 (Hz) AC
AC Vd Motor

Rectifier Filter Switch-mode


Capacitor inverter

Figure 13. Typical variable voltage and variable frequency system.

8.2.1 One-Phase Inverter


Figure 14 shows the operation of one leg of an inverter regardless of the number of phases. To
illustrate the point better, the input DC voltage is divided into two equal parts. When the upper switch,
V
S1, is closed while S2 is open, the output voltage VAo will be + d , and when the lower switch, S2, is
2
Vd
closed while S1 is open, the output voltage will be
2
+
Vd S1
2
Vd vAo +
+
Vd S2 vAN
2

Figure 14. One leg of an inverter.


To control the output waveform the switches can be controlled by pulsed width modulation, PWM,
where the time each switch is closed can be determined by the difference between a control waveform,
and a carrier (or triangular) waveform as shown in Figure 15. When the control wave is greater than the
triangular wave, S1 is closed, and S2 is open. When the control wave is less than the triangular wave, S1
is open, and S2 is closed. In this way, the width of the output is modulated (hence the name).

1.0V

0.5V

0.0V

-0.5V

-1.0V
16 m s 1 8 ms 20ms 2 2m s 24ms 26 m s 2 8 ms 30ms 3 2m s 34 m s
V 1 (V t r i ) V 1( V c o nt )
T im e

Figure 15. Control (red sinewave) and carrier or triangular waveform (green) determine when the
switches are closed. For the times when the control wave is greater than the triangular wave, S1
is closed, and S2 is open. When the control wave is less than the triangular wave, S1 is open, and
S2 is closed.

60V

40V

20V

-0V

-20V

-40V

-60V
16ms 18ms 20ms 22ms 24ms 26ms 28ms 30ms 32ms 34ms
V (3)- V(2)
Time

Figure 16. Output voltage VAo corresponding to the control and carrier waves shown in Figure 15.
The frequency of the triangular wave is fc, and the frequency of the control wave is fo. We define the
ratio of these as the frequency modulation index, mf.
f
mf = c (8.17)
fo
Likewise, the amplitude modulation index, ma, is defined as the ratio of the control voltage to the
triangular wave voltage.
Vcontrol
ma = (8.18)
Vtriangular

A one phase, bull wave inverter is shown in Figure 17. It has four controlled switches, each with an
antiparallel diode.

+
Vd S1 S4
2
Vd o A B
Vd S2 S3
2

+
vAB = vAo- vBo

Figure 17. One-phase full wave inverter.

The diagonal switches operate together such that S1 and S3 open and close together, and S2 and S4
V V
open and close together. The output will oscillate between + d and d . If the Fourier
2 2
transformation of the pulse width modulated square wave shown in Figure 18 is taken, the amplitude of
the fundamental will be a linear function of the amplitude index Vo = maVd/2 as long as ma 1. Then the
RMS value of the output voltage will be:
ma Vd
Vo1 = = 0.353maVd (8.19)
2 2
12 0V

8 0V

4 0V

- 0V

-4 0V

-8 0V

-12 0V
16ms 18ms 20ms 22ms 24ms 26m s 28m s 30 ms 32 ms 34m s
V(2)- V(3)
Time

Figure 18. Output voltage VAB for the one-phase, full wave inverter of Figure 17.

When ma increases beyond 1, the output voltage increases also, but not linearly with ma. The output
4
amplitude can reach a peak value of Vd when the reference signal becomes infinite and the output is a

square wave. Under this condition, the RMS value is:
2 2 Vd
Vo1 = = 0.45Vd (8.20)
2
Equating the power of the DC side with that of the AC side gives
P = Vd I do = Vo1I o1 pf (8.21)
Thus for normal operation:
I do = 0.353ma I o1 pf (8.22)
and in the limit for a square wave:
I do = 0.45I o1 pf (8.23)
8.2.2 Three-Phase Inverter

For three-phase loads it makes more sense to use a three-phase inverter shown in Fig., rather than using
three one-phase inverters.

S1 S2 S3
+
Vd
S4 S5 S6

A B C

Figure 19. Three-phase, full wave inverter.

The basic PWM scheme for a three-phase inverter has one common carrier and three separate control
waveforms. If the waveforms we want to achieve are sinusoidal the frequency modulation index, mf, is
low we use a synchronized carrier signal with mf an integer multiple of 3.

1.0V

0.5V

0.0V

-0.5V

-1.0V
16ms 17ms 18ms 19ms 20ms 21ms 22ms 23ms 24ms
V1(Vcont1) V1(Vcont2) V1(Vcont3) V1(Vtri)
Time

(a)
110V

80V

40V

0V

16m s 17 ms 1 8ms 19ms 20ms 21ms 22ms 23ms 24 m s


V(2)
Time

(b)
110V

80V

40V

0V

16m s 17 ms 1 8ms 1 9 ms 20ms 21ms 22ms 23 m s 24 m s


V( 3 )
Time

(c)
120V

80V

40V

-0V

-40V

-80V

-120V
16ms 17ms 18ms 19ms 20ms 21ms 22ms 23ms 24ms
V(2)- V(3)
Time

(d)

Figure 20.Three-phase, full wave inverter showing (a) control and triangular wave, (b) phase A line-to-
neutral, (c) phase B line-to-neutral, and (d) line-to-line signal from phase A to B.
110V

80V

40V

0V

1 7.00ms 17 .04ms 17. 08ms 17. 12ms 17.1 6ms 17.20 ms 17.24m s 17.28ms 17.32ms 17.36ms 17.40m s
V(2)
Time

(a)
110V

80V

40V

0V

1 7.00ms 17 .04ms 17. 08ms 17. 12ms 17.1 6ms 17.20 ms 17.24m s 17.28ms 17.32ms 17.36ms 17.40m s
V(3)
Time

(b)
120V

80V

40V

-0V

-40V

-80V

-120V
17.00ms 17.04ms 1 7.08ms 17.12ms 17. 16ms 17.20ms 17.24 ms 17.28ms 17.32ms 17 .36ms 17.40ms
V(2)- V( 3)
Tim e

(c)

Figure 21.Shorter time span for the (a) phase A line-to-neutral, (b) phase B line-to-neutral, and (c) line-to-
line signal from phase A to B of the outputs in Figure 20.
The PSPICE netlist for this three phase PWM circuit is given below:

3 phase FULL-BRIDGE INVERTER


************ OUTPUT IS V(2,3) ****************
**********INPUT PARAMETERS ****************
.PARAM Vsource = 100 ; DC input to inverter
.PARAM Fo = 200 ; fundamental frequency
.PARAM Mf = 21 ; carrier, multiple of Fo
.PARAM Ma = .8 ; amplitude ratio
.PARAM Fc = {Mf*Fo} ; carrier frequency

VS 1 0 DC {Vsource} ; dc source
******* VOLTAGE-CONTROLLED SWITCHES *****
S1 1 2 40 30 SWITCH
S2 1 3 50 30 SWITCH
S3 1 4 60 30 SWITCH
S4 2 0 30 40 SWITCH
S5 3 0 30 50 SWITCH
S6 4 0 30 60 SWITCH
************** FEEDBACK DIODES *************
D1 2 1 DMOD
D2 3 1 DMOD
D3 4 1 DMOD
D4 0 2 DMOD
D5 0 3 DMOD
D6 0 4 DMOD
******************** LOAD ******************
R1 2 5 10 ; load between nodes 2 and 0
L1 5 0 60MH
R2 3 6 10 ; load between nodes 3 and 0
L2 6 0 60MH
R3 4 7 10 ; load between nodes 4 and 0
L3 7 0 60MH
*************** TRIANGLE CARRIER **************
Vtri 30 0 PULSE (1 -1 0 {1/(2*Fc)} {1/(2*Fc)} 1ns {1/Fc})
******************** REFERENCE *****************
Vcont1 40 0 SIN(0 {Ma} {Fo} 0 0 {-90/Mf})
Vcont2 50 0 SIN(0 {Ma} {Fo} 0 0 {-90/Mf - 120})
Vcont3 60 0 SIN(0 {Ma} {Fo} 0 0 {-90/Mf - 240})
************ MODELS AND COMMANDS *************
.MODEL SWITCH VSWITCH(RON=0.001 VON=.007 VOFF=-.007)
.MODEL DMOD D ; default diode
.PROBE
.TRAN 0.5MS 33.33MS 16MS 0.1MS
.FOUR 200 25 I(R1) ; Fourier transform
.OPTIONS NOPAGE ITL5=0
.END
As long as ma is less than 1, the RMS value of the fundamental of the output voltage is a linear
function of it:
3
Vl l1 = maVd 0.612maVd (8.24)
2 2
In the limit, when the control voltage becomes infinite, the RMS value of the fundamental of the
output is then:
3 4 Vd
Vl l1 = 0.78Vd (8.25)
2 2
Again, equating the power on the DC and AC sides we obtain:
P = Vd I do = 3Vl l1I o1 pf (8.26)
or for normal PWM operation:
I do = 1.06ma I o1 pf (8.27)
and in the limit for the square wave:
I do = 1.35 I o1 pf (8.28)

Finally, there are other ways to control the operation of an inverter. If it is not he output voltage
waveform we want to control, but rather the current, we can either impose a fast controller on the voltage
waveform, driven by the error between the current signal and the reference, or we can apply a hysteresis
band controller, shown for one leg of the inverter in Figure 22.
+
Vd S1
2
Vd vAo +
+
Vd S2 vAN
2

4 .0A

2 .0A

0A

-2 .0A

-4 .0A
16 ms 18 ms 20 ms 22 ms 24 ms 26 ms 28 ms 30 ms 32 ms 34ms
I( R)
Ti me

Comparator
tolerance
band

i*A i Switch-
i Mode
+ Inverter
_
iA

Figure 22. Current control with hysteresis band.

8.2.3 Inverter Notes


With a sine-triangle PWM the harmonics of the output voltages are of frequency around nfn, where
n is an integer and fn is the frequency of the carrier (triangle) waveform. The higher this frequency
is the easier to filter out these harmonics. On the other hand, increasing the switching frequency
also increases proportionally the switching losses. For 6-step operation of a 3-phase inverter the
harmonics are even, except the tripled ones, i.e. they are of order 5, 7, 11, 13, 17 etc.
When the load of an inverter is inductive the current in each phase remains positive after the voltage
in that phase became negative, i.e. after the top switch has been turned off. The current then flows
through the antiparallel diode of the bottom switch, returning power to the DC link. The same
happens of course when the bottom switch is turned off and the current flows through the
antiparallel diode of the top switch.

8.2.4 Example 1
A three-phase controlled rectifier is supplying a DC motor that has k = 1 (Vs) and R = 1 (). The
rectifier is fed from a 208 (Vl-l) source.

208 (V)
Filter

Figure 23. Figure for example problem 8.2.4.

a) Calculate the maximum no-load speed of the DC motor:

Without a load the current is zero, therefore:

V = k + IR = k

The maximum speed is then found by the maximum DC voltage

Vmax = kmax

The maximum DC voltage is provided by the controlled rectifier for = 0.

Vmax = 1.35Vl l = 281.8 (V)

Therefore,
max = 280.8 (rad/s)

b) The motor now is producing torque of 20 (Nm). What is the maximum speed the motor can achieve?

Now that there is load torque, there is also current:

T = kI I = 20 (A)
Then
V IR 280.8 20 1
= = = 260.8 (rad/s)
k 1

c) For the case in b) calculate the total RMS current of the fundamental and the power factor at the AC
side

At the maximum, the fundamental of the AC current is:


I s1 0.78I d = 15.6 (A)
The power factor is then 1.

d) If the motor is now connected as a generator with a counter torque of 20 (Nm) at 1500 (rpm). What
should be the delay angle and AC current?

For a DC generator:
T 2 2
V = k IR = k R 1 1500 1 = 137.08 (V)
k 60 1

Since this is a generator, the voltage is negative for the inverter.

137.08 (V) = 1.35 208 cos( )


= 119.22

8.2.5 Example 2
For the system shown in Figure 24, the AC source is constant, and the load voltage is 150 (Vl-l),
20 (A), 52 (Hz), 0.85 pf lagging.

208 (V)
6-step
Filter
inverte

Figure 24. Figure for example problem 8.2.5.

a) Calculate the voltage on the DC side and the DC component of the current.

For the 6-step inverter


Vl l ,1 = 0.78Vd Vd = 192 (V)
1.35 150 20 0.85
P = 3Vl l I l pf = Vd I d 0 I d 0 = = 23 (A)
192

b) Calculate the AC source side RMS and fundamental current and power factor.

For a 3-phase rectifier


Vd = 1.35Vl l cos( ) 192 = 1.35 208 cos( ) cos( ) = 0.685
I s1 = 0.78 I d = 17.94 (A)
pf = cos( ) = 0.685 lagging
8.3 DC-DC Conversion
DC to DC conversion is often associated with stabilizing the output while the input varies, however
the converse is also required in some applications, which is to produce a variable DC from a fixed or
variable source. The issues of selecting component parameters and calculating the performance of the
system is the focus of this section. Since these converters are switched mode systems, they are often
referred to as choppers.

8.3.1 Step-Down or Buck Converters


The basic circuit of this converter is shown in Figure 25 connected to a purely resistive load. If we
remove the low pass filter shown and the diode, the output voltage vo(t) is equal to the input voltage Vd
when the switch is closed, and zero when the switch is open. The average output, Vo, is then:

1 ton Ts ton
Vo =
Ts 0
Vd dt +
t on0 dt =
Ts
Vd (8.29)

+
Low-pass filter
iL io
L
Vd +
+ vL
C vo = Vo R
(load)

vo

Vd
Vo

0 t
t on
t off

1
Ts = f
s

Figure 25. Topology of the buck chopper.


The ratio of ton/Ts = D, the duty ratio.
The low pass filter attenuates the high frequencies (multiples of the switching frequency) and leaves
almost only the DC component. The energy stored in the filter inductor (or the load inductor) has to be
absorbed somewhere other than the switch, hence the diode, which conducts when the switch is open.
Here we study this converter in the continuous mode of operation such that the current through the
inductor never becomes zero. As the switch opens and closes the circuit assumes one of the topologies of
Figure 26.
vL

Vd - Vo
A
t

- Vo
1
Ts = f
iL s

i L = Io

0 t
t on
t off

iL iL
L L
+ + +
+ vL + vL
Vd C Vo R C Vo R

Figure 26. Operation of the buck chopper.

We will use the fact that the average voltage across the inductor is zero, and assume a perfect filter
such that the voltage across the inductor is (Vd Vo) during ton, and Vo for the remainder of the cycle.

ton Ts
Vo =
0 (Vd Vo )dt + ( Vo )dt = 0
ton
(8.30)
(Vd Vo )ton Vo (Ts ton ) = 0 (8.31)
Vo ton
= =D (8.32)
Vd Ts
Also using the fact that the input and output powers are the same gives:
Vd I d = Vo I o (8.33)
I d Vo
= =D (8.34)
I o Vd
We analyzed this under the assumption of continuous mode of operation. In the discontinuous mode,
the output DC voltage is less than what is given here, and the chopper is less easy to control. At the
boundary between continuous and discontinuous modes, the inductor current reaches zero for one instant
every cycle, as shown in Figure 27.

vo
iL

Vd - Vo IL

iL
i L, avg= I L T s Vd
t I L,max=
0 8L

D
- Vo 0 0.5 1.0

t on
t off

1
Ts = f
s

Figure 27. Operation of the buck converter at the boundary of continuous conduction.

From this figure we can see that at this operating point, the average inductor current is IL = iL, and:

1 DTs
I L = ton (Vd Vo ) = (Vd Vo ) (8.35)
2 2L
Since the average inductor current is the average output current (the average capacitor current is zero),
equation (8.35) defines the minimum load current that will sustain continuous condition.
Finally, a consideration is the output voltage ripple. We assume the ripple current is absorbed by the
capacitor, i.e. the voltage ripple is small. The ripple voltage is then due to the deviation from the average
of the inductor current as show in fig. Under these conditions:
Q 1 1 I L Ts
Vo = = (8.36)
C L2 2 2
where
V
I L = o (1 D )Ts (8.37)
L

Vo 1 Ts2
= (1 D ) (8.38)
Vo 8 LC

vL

Vd - Vo
A
t

- Vo
1
Ts = f
iL s

i L = Io
2
0 t
t on
t off

vo

Vo

Vo

0 t

Figure 28. Analysis of the output voltage ripple of the buck converter.
Another way to view this is to define the switching frequency fs = 1/Ts and use the corner frequency of the
( )
filter f corn = 1 2 LC :
2
Vo 2
= (1 D ) f corn (8.39)
Vo 2 fs

8.3.2 Step-Up or Boost Converter


Here the output voltage is always higher than the input. The schematic is shown in Figure 29.

iL
L
+ +
+ vL
Vd C Vo R

Figure 29. Schematic diagram of a boost converter.

Based on the condition of the switch, there are two possible topologies as shown in Figure 30. Again,
the way to calculate the relationship between input and output voltage we take the average current of the
inductor to be zero, and the output power equal to the input power giving:

Vd ton + (Vd Vo )(Ts ton ) = 0 (8.40)


V 1
o = (8.41)
Vd 1 D
I
o =1 D (8.42)
Id
vL

Vd - Vo
A
t

- Vo
1
Ts = f
iL s

i L = Io

0 t
t on
t off

iL iL
L L
+ + vL + + +
+ vL
Vd C Vo R Vd C Vo R

Figure 30. Two circuit topologies of the boost converter.

To determine the values of inductance and capacitance we study the boundary of continuous
conduction like before and the output voltage ripple. At the boundary of the continuous conduction, the
geometry of the current waveform gives:
TV
I o = s o D(1 D )2 (8.43)
2L
The output current must exceed this value for continuous conduction. Using the ripple analysis as
shown in Figure 31 we find:
Vo DTs
= (8.44)
Vo RC
It is important to note that the operation of a boost converter depends on parasitic components,
especially for duty cycle approaching unity. These components will limit the output voltage to levels well
below those given by equation (8.41).

iD

Q i D= Io
t

vo

Vo

Vo

0 t
DTs (1-D)Ts

Figure 31. Calculating the output voltage ripple for a boost inverter.

8.3.3 Buck-Boost Converter

This converter has a schematic shown in fig. and can provide output voltage that can be lower or higher
than the input voltage.

Again the operation of the converter can be analyzed using the two topologies resulting from the
operation of the switch as shown in fig.
By equating the integral of the inductor voltage to zero we get:
Vd DTs + ( Vo )(1 D )Ts = 0 (8.45)
V D
o = (8.46)
Vd 1 D
At the boundary between continuous and discontinuous conduction we find
TV
I o = s o (1 D )2 (8.47)
2L
The output ripple, as calculated from Fig. is
Vo DTs
= (8.48)
Vo RC

iD

+
+ iL
Vd vL L C Vo R

io
Figure 32. Basic Buck-Boost converter.
vL

Vd
A
t

- Vo
1
Ts = f
iL s

I L = (Id + Io )

0 t
DTs (1-D)Ts

iD

+ +
+ iL + iL
Vd vL L C Vo R Vd vL L C Vo R

+ +

io io

Figure 33. Operation of a Buck-Boost chopper.

8.3 Example
The input of a step down converter varies from 30 (V) to 40 (V) and the output voltage is to be
constant at 20 (V), with output power varying between 100 (W) and 200 (W). The switch is operating at
10 (kHz). What is the inductor needed to keep the inductor current continuous? What is then the filter
capacitor needed to keep the output ripple below 2%?

The duty cycle will vary between D1 = 20/30 = 0.667 and D2 = 20/40 = 0.5. The load current will
range between Io1 = 100/20 = 5 (A), and Io2 = 200/20 = 10 (A).
The minimum current needed to keep the inductor current continuous is:

DTs
I o, min = (Vd Vo )
2L
Since the constant is the output voltage, Vo, and the minimum load current must be greater than Io,min, we
can express it as a function of Vo and make it less than or equal to 5 (A), or:

DTs
5(A ) I o, min = (Vd Vo ) = VoTs (1 D )
2L 2L

Ts = 1/10(kHz), Vo = 20 (V), and the maximum value is achieved for D = 0.5, leading to Lmin = 50 (H).
For the ripple, the highest will occur at 1-D = 0.5, thus:

2
2 f
0.02 = 0.5 corn f corn = 900 (Hz)
2 10 103

1
= 900 (Hz)
6
2 50 10 C

C = 625 (F)

You might also like