Вы находитесь на странице: 1из 158

March 1, 2001,

Aalborg University
Institute of Energy Technology

Dynamic Motor Modelling


(or the IET-way)

Rasmus Post
Ewen Ritchie
Aalborg University

2000
March 1, 2001,
Introduction
At my university, the form of the education is different from most universities.
In most universities are the education is centred around lectures, with work in the
laboratory and a final project in the end. At Aalborg University is everything centred
around the project. Every semester has a theme and within each semester there is
a project. The project is done by a group of 2-7 students. This group do all their work
together within one semester. The groups change from one semester to the next. The
project is based on the theme of the semester and uses between 40-60% of the time
and must contain theoretical analysis, modelling of the problem, practical
implementation in the laboratory, verification of model. The work in the laboratory
and the theoretical work are written in a rapport describing the work and results.
The courses given in the semesters are of two types: Study-oriented courses and
project oriented courses. The study-oriented courses are about general knowledge,
such as mathematics and physics, whereas project oriented courses are specific
courses within the semester theme. The study-oriented courses are usually evaluated
by a written or an oral exam and the project oriented courses are evaluated through
an oral exam together with the semester project. The courses given consist of five
lectures, starting with the presentation from the teacher followed by problem solving
in the project groups.
I have been teaching electrical motor modelling at the university for some
years. Every time a semester starts there is the problem of which book to select as
the textbook. Many books exist in the area of electrical machine modelling and a
large portion of them are very good. However, I find it difficult to find a book that
covers the required theory. This is because I would like the book to cover the general
theory of electrical machines, both the classical motors such as DC-motors and
induction motors, but also modern motors such as BLDCM and SRM. I would also
like the book to cover how to model the machines and how to get model parameters
from measurements. I would like the book to cover the why aspect, like Why is it
a good idea to a specific thing ths way and not that way. This is the reason for
writing this book.
In the following are scalar values shown with normal weight as rs , a matrix
e.g. showing a three-phase stator resistance will be bold as r s and in the same way
a vector containing a three-phase current also will be bold as i s . Anything describing
6
a spacial orientation will be shown with an arow as is .

i
March 1, 2001,
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . i

Chapter 1
The mechanical system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -1-
1.1 Mechanical motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -1-
1.1.1 Linear motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -1-
1.1.2 Rotating motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -2-
1.1.3 Friction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -4-
1.1.4 The whole mechanical system . . . . . . . . . . . . . . . . . . . . . . . . . . -5-
1.1.5 Some more about moments of inertia . . . . . . . . . . . . . . . . . . . . . -6-
1.2 Determining mechanical properties . . . . . . . . . . . . . . . . . . . . . . -8-
1.2.1 The swinging pendulum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -9-
1.2.2 The run-out test . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -10-
1.2.3 Which method to select? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -11-

Chapter 2
Electromechanical energy conversion . . . . . . . . . . . . . . . . . . . -13-
2.1 The Lorenz-force and Faradays law . . . . . . . . . . . . . . . . . . . . -14-
2.1.1 The Lorenz-force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -14-
2.1.2 Faradays law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -16-
2.1.3 Acting together . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -18-
2.2 Work, force and torque . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -19-
2.2.1 Energy and co-energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -20-
2.2.2 The electric circuit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -22-
2.2.3 The magnetic circuit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -24-
2.2.4 The mechanical circuit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -26-

Chapter 3
Introduction to the DC-motor . . . . . . . . . . . . . . . . . . . . . . . . . . -34-
3.1 The archetypical DC-motor . . . . . . . . . . . . . . . . . . . . . . . . . . . . -34-
3.1.1 Generating torque in a DC-motor . . . . . . . . . . . . . . . . . . . . . . . -34-
3.1.2 Lap or wave winding . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -38-
3.1.3 EMF of a DC-motor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -41-
3.1.4 Torque constant, voltage constant and field constant . . . . . . . -43-
3.2 The DC-motor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -43-
3.2.1 Separately excited DC-motor . . . . . . . . . . . . . . . . . . . . . . . . . . -43-
3.2.2 Shunt motor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -45-
3.2.3 Series motor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -46-
3.2.4 Compound motor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -47-
3.2.5 Permanent magnet DC-motor . . . . . . . . . . . . . . . . . . . . . . . . . . -49-

ii
March 1, 2001,
3.2.6 Universal motor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -49-
3.3 Making the DC-motor work . . . . . . . . . . . . . . . . . . . . . . . . . . . -50-
3.3.1 Armature reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -50-
3.3.2 Incomplete commutation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -53-
3.4 Getting motor parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . -56-
3.4.1 The brushes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -56-
3.4.2 Field and armature resistance . . . . . . . . . . . . . . . . . . . . . . . . . -56-
3.4.3 Field and armature inductance . . . . . . . . . . . . . . . . . . . . . . . . -58-
3.4.4 The field constant kM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -59-
3.5 Simulating the DC-motor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -60-
3.6 The transfer function of a DC-motor . . . . . . . . . . . . . . . . . . . . -61-

Chapter 4
Introduction to the induction motor . . . . . . . . . . . . . . . . . . . . . -64-
4.1 The three phase induction motor . . . . . . . . . . . . . . . . . . . . . . . -64-
4.1.1 Evolving the induction motor . . . . . . . . . . . . . . . . . . . . . . . . . . -64-
4.1.2 Construction of the induction motor . . . . . . . . . . . . . . . . . . . . . -68-
4.2 Modelling the induction motor . . . . . . . . . . . . . . . . . . . . . . . . . -71-
4.2.1 Stator voltages and input power . . . . . . . . . . . . . . . . . . . . . . . . -71-
4.2.2 Magneto-motive force of an induction motor . . . . . . . . . . . . . . -73-
4.2.3 Induced voltages of an induction motor (EMF) . . . . . . . . . . . . -76-
4.2.4 Winding factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -77-
4.2.5 Single-phase star equivalent circuit . . . . . . . . . . . . . . . . . . . . . -80-

Chapter 5
Dynamic modelling of the induction motor . . . . . . . . . . . . . . . -87-
5.1 Phase transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -87-
5.1.1 One rotating coil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -90-
5.1.2 Stationary frame of reference . . . . . . . . . . . . . . . . . . . . . . . . . . -90-
5.1.3 Rotating frame of reference . . . . . . . . . . . . . . . . . . . . . . . . . . . -91-
5.1.4 A bit of anarchy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -92-
5.1.5 Transformation of currents, voltages, flux-linkage, etc. . . . . . -92-
5.2 Transformation of the voltage equation . . . . . . . . . . . . . . . . . . -99-
5.2.1 Transformation of stator-voltages . . . . . . . . . . . . . . . . . . . . . . -99-
5.2.2 Transformation of rotor voltages . . . . . . . . . . . . . . . . . . . . . . -101-
5.3 Transformation of three-phase resistors and flux-linkage . . -102-
5.3.1 Transformation of a three-phase resistor . . . . . . . . . . . . . . . . -102-
5.3.2 Transformation of flux-linkage . . . . . . . . . . . . . . . . . . . . . . . . -103-
5.4 Equivalent circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -104-
5.4.1 Equivalent circuit, rotating frame of reference . . . . . . . . . . . -105-
5.4.2 Equivalent circuit, stationary frame of reference . . . . . . . . . -105-

iii
March 1, 2001,
5.5 The electromagnetic torque . . . . . . . . . . . . . . . . . . . . . . . . . . -106-
5.6 Solving the system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -108-
5.6.1 Solving for the currents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -108-
5.6.2 Solving for the flux-linkage . . . . . . . . . . . . . . . . . . . . . . . . . . . -110-

Appendix 1
Moments of Inertia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -114-

Appendix 2
Simulating differential equations using Matlab/Simulink . -120-

Appendix 3
Litterature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -123-

Exercises
Relevant Exercises for the Chapters . . . . . . . . . . . . . . . . . . . -125-
E.1 Exercises for Chapter 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -125-
E.2 Exercises for Chapter 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -127-
E.3 Exercises for Chapter 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -130-
E.4 Exercises for Chapter 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -133-

Solutions
Solutions to the problems (do not cheat!!) . . . . . . . . . . . . . . -135-
S.1 Solutions for Chapter 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -135-
S.2 Solutions for Chapter 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -142-
S.3 Solutions for Chapter 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -150-
S.4 Solutions for Chapter 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -157-

iv
March 1, 2001,
Chapter 1
The mechanical system
The goal is to be able to model a combined system consisting of an electrical
system, an electromagnetic system and a mechanical system. It is important that the
model of the whole system also includes a model for the mechanical part of the
system, because it is finally the goal to either convert electrical energy into
mechanical energy (as a motor) or mechanical energy into electrical energy (as a
generator).

1.1 Mechanical motion


This is a model of a motor. This means that a mechanical system is involved.
For simplicity, linear motion is looked at first and later the linear motion is
developed into rotating motion.

1.1.1 Linear motion

For linear motion, the forces acting on a body may usually be simplified to a
driving force, Fe, acting on the mass, and an opposing force (or load), Fl, as shown on
Figure 1 .

Figure 1 A body with two forces.


For linear motion the following may be written
dx t dv t d 2x t
vt ' , at ' ' (1)
dt dt dt 2
and the resulting force acting on the body then becomes
dv t
Fm t ' Fe t & Fl t ' m @ a t ' m @ (2)
dt

If the resulting force causes the body to move, the resulting work on the system (the
kinetic energy) becomes

-1-
March 1, 2001,
dWm t ' Fm @ dx t
dWm t dx t
pm t ' ' Fm @ ' Fm @ v t
dt dt (3)
t t v
vJ 1
Wm t ' pm J dJ ' m @ v J @ dJ ' m @ v J du ' @m@vt2
0
dJ 0
2
0

This gives the energy stored in the moving masses, but not the energy lost to the
load. At any given moment the energy stored in the moving mass is therefore
proportional to the speed squared.

1.1.2 Rotating motion

If the motion is linear instead of linear, a situation as shown in Figure 2


arrises.

Figure 2 A rotating body.


By examining an infinitely small volume, the differential force acting upon it
becomes
dv
dF ' dm @ a ' dm @ (4)
dt

By rotating motion, the relationship between speed, angular speed and radius is
v ' r @ Tm (5)

The differential torque acting on the infinitely small volume may therefore be
calculated as
dTm
dT ' r 2 @ dm @ (6)
dt
The differential volume, which the differential torque from (6) acts upon may be
found from Figure 3 .

-2-
March 1, 2001,
Figure 3 The differential volume.
The differential mass of the volume shown in Figure 3 may then be found as
dm ' D @ rd1 @ dr @ dl (7)

where D is the specific mass-density of the body. The differential torque then becomes
dTm
dT ' D @ r 3 dr d1 dl @ (8)
dt
The combined torque of all the differential volumes may then be found by integration
as
r1 2B l
dTm
T ' D@ 3
r dr d1 dl @
m m m dt
0 0 0

B 4 dTm (9)
' @ D @ l @ r1 @
2 dt
dTm
' Jcyl @
dt
A very important term has been defined. This term is the moment of inertia of a
cylinder. The moment of inertia is a term which describes the mass and shape of an
object with regard to how it slows the rate-of-change in angular speed around a given
axis. The moment of inertia is a good term, because describes the physical properties
of an object. The equation in (9) may be simplified by examining the integral. In
Figure 4 is shown a body of mass rotating around the axis O-O.

-3-
March 1, 2001,
Figure 4 A body of mass rotation about the axis O-O.
When the differential element dm is rotated around its axis of rotation, it is
accelerated following a circular curve, with a resulting force as a tangent to the
circle. By summing up for all differential elements of the volume, the moment of
inertia is then found from the integral over the volume as
N
2 2
J ' r0 dm ' ri mi (10)
i'1

where r0 is the radius to the differential element. Finding the moment of inertia then
simply becomes a matter of summing up all the individual parts of the volume. This
is, as will be shown later, also true for non-differential elements of a finite size.
Incidentally, for the rotating body, the stored kinetic energy becomes
1
Wm ' @ J @ T2 (11)
2

A small note on moments of inertia: In most text-books is the letter for the
moment of inertia usually a I, but electrical engineers will like to reserve this
character for the current. Therefore, in text-books on electrical machines, the
moment of inertia is denominated by a J.

1.1.3 Friction

Friction is a little complicated to model, because it is fundamentally non-


linear, with a number of external components influencing it. It is, however, often
sufficient to model it as a 1'st order approximation as shown in the following.
In Figure 5 is shown a sketch of how the frictional torque changes as a

-4-
March 1, 2001,
function of the angular speed.

Figure 5 The frictional torque in a motor as a


function of the angular speed.
Picturing a box resting on the floor. By starting to push the box, one
experiences that a specific force must be used to get the motion started, however,
when the motion has started, the force becomes smaller by some degree. This
phenomenon is called stiction, and only happens when the speed is very small.
Another thing which is experienced is that it requires a minimum force to keep the
motion going. If the applied force drops below this force, the motion will stop
eventually. This phenomenon is called dry-friction (or coulomb friction) . The dry-
friction is seen as the offset on Figure 5 . Lastly, when pushing the box, it is
observed that when the speed increases, the force required to maintain the speed
increases more or less linear with the speed. This phenomenon is called the viscous-
friction and is usually approximately linearly dependent on the speed. Together this
gives the friction as
T fric ' B @ Tr,mec % Tc % T stic (12)

1.1.4 The whole mechanical system

Putting it all together, the developed electrical torque must balance the load
torque, all the frictions and the torque used to accelerate the rotating mass. The
differential equation modelling the mechanical system then becomes

-5-
March 1, 2001,
dTr,mec
Te ' J @ % B @ Tr,mec % T c % Tstic % T l (13)
dt
It is important to remember that the moment of inertia must be both the moment of
inertia for the motor and for the load, unless the load moment of inertia is included
in the load torque. When the individual components in (13) are known, the equation
is integrated to find the angular speed. If the result is once again the rotor-angle is
found. Very importantly, the viscous-friction is a function of the angular speed,
which means that it changes is the motor is reversed. This is not the case for the dry-
friction or the load. It is therefore important to know if the sign must be changed,
otherwise the situation where the load drives the motor appears. However,
sometimes the load does drive the motor. This is the case when the motor runs as a
generator. This is shown in Figure 6 .

Figure 6 Operating modes for an electrical


motor.

1.1.5 Some more about moments of inertia

As shown above, is the moment of inertia a describing factor for the rotating
motion. In (9) was the moment of inertia for a cylindrical body found by performing
a triple-integral. This, however, is not practical, due to a number of reasons. Firstly,
a triple-integral is not easy to perform, and secondly, it takes time and it is easy to
include errors. Therefore, in books on mechanical modelling, tables of commonly used
shapes are precented. In Appendix 1 is a table of commonly used shapes. Looking
in data-sheets for devices, the moment of inertia might not be listed. Replacing the
moment of inertia will be the mass of the device and the radius of gyration. The
radius of gyration is defined as

-6-
March 1, 2001,
J
k ' , J ' k2 @ m (14)
m
This may be interpreted as all the mass is concentrated in a narrow shell with the
radius k.
Many of the shapes used in mechanical system cannot be found in the table
in Appendix 1. The problem is also that many shapes consists of materials of
different mass-density. However, by examining the equation in (9) more closely, it
shows that it is possible to add or subtract different moments of inertia to form the
desired shape. The first shape in the table in Appendix 1 is a circular cylindrical
shell as shown in Figure 7 on the left-hand side.

Figure 7 On the left-hand side is a circular cylindrical


shell and on the right-hand side is a circular cylindrical
tube.
The inner and outer radii of the shell are approximately the same. In this case, the
moment of inertia along the z-axis is
Jzz ' mr 2 (15)

which is the same as the radius of gyration in (14). This is not the case of the shape
shown in Figure 7 on the right-hand side. In this case, there is a significant
difference between the two radii. Using the third equation in the table in Appendix
1, the moment on inertia along the z-axis becomes the moment of inertia of a solid
circular cylinder with the outer radius, from where the moment of inertia of a solid
circular cylinder with the inner radius subtracted
2 2
J1 ' 1
2
m1r1 , J2 ' 1
2
m2r2

J ' J2 & J1
2 2
(16)
' 1
2
m1r2 & 1
2
m2r1
2 2
' 1
2
m1r2 & m2r1

Re-phrasing this, the moment of inertia of any shape may be constructed by adding

-7-
March 1, 2001,
or subtracting moments on inertia of simpler shapes.
What remains in the above is what happens when the axis of rotation is
different from the axis described in Appendix 1. In Figure 8 is the shape from
Figure 7 shown displaced by the distance d.

Figure 8 The shape from Figure 7 displaced the


distance d from the axis of rotation.
In this case the moment of inertia becomes
J ' J0 % md 2 (17)

where d is the distance the body is moved away from the axis of rotation. The
method for finding the moment of inertia of a shape then becomes: First subdivide
the shape into simpler shapes and calculate the individual moments of inertia along
the local axis. Then displace, is necessary, the shape to the axis of rotation. Finally
add and subtract the individual moments of inertia to form the final shape.

1.2 Determining mechanical properties


One thing is to have the model for the mechanical system, a different thing is
to find the model parameters. It stands to reason that the friction parameters must
be found from measurement, however, from the above it appears that the moment
of inertia may be found from calculations done on a known shape. For simple shapes
this is possible but for, e.g. the rotor (armature) of a DC-motor it is usually not
feasible. This is due to the complex shape and the varying mass densities of the
materials. One instance where the moment of inertia may be found easily, is when
the design is done in, e.g. a CAD-program. A feature often found in these programs
is to calculate the moment of inertia. However, the best results are usually found
from measurements. Following is precented some methods for finding the moment

-8-
March 1, 2001,
of inertia from measurements.

1.2.1 The swinging pendulum

The first method described requires that the object for which the moment of
inertia must be found is accessible. If the object is a rotor of an electrical machine,
it might not always be possible to remove it. If it is possible to remove the object, the
method is simply to hang it from wires as shown in Figure 9 .

Figure 9 The swinging pendulum.


The method is to displace the object so it starts to swing as a pendulum. By
measuring the frequency of the oscillation the radius of gyration is found as
1 gr
f' @ (18)
2B k2
where g is the gravitational constant. The inertia is then found by using (14) and
(17) and remembering that the mass is displaced by the radius in Figure 9 . This
method only works well if the wires are long compared to the radius of gyration of
the shape and only for small oscillations.

1.2.2 The run-out test

It is not always possible or practical to remove the rotor of a motor to


determine the moment of inertia from the swinging pendulum. A different test is to
exploit that equation (13) is a first order differential equation, and as such always

-9-
March 1, 2001,
will have an exponential increase or decay as a step-response. The time-constant is
measured and used to determine B and J. In the mechanical system the inverse step
is the easiest step to produce. In theory, there should be no difference between a step
or an inverse-step, however, a step would require a constant torque through the
entire startup which is difficult, whereas the inverse step only requires the motor to
be shut off. The test is done by running the motor in no-load on its own power to a
specified angular speed. All windings are then opened, thereby removing the
electromagnetic torque and thus producing the inverse step in torque. Two different
methods exist to find B and J from the run-out test.

Method 1:
One test is done in no-load and the time-constant from the decaying angular
speed is determined. A second test is done, but this time an extra known moment of
inertia is attached to the rotor and a new time-constant is found. The moment of
inertia of the rotor and the viscous-friction are then found from
J
J1 '
B
(19)
J % Jtest
J2 '
B

B and J are then easily found.

Method 2:
It is not always possible or practical to attach an extra moment of inertia. In
this case a different method is used. A test is done in no-load and the angular speed
as a function of time is recorded. Then equation (13) is applied, but because it is no-
load and because the driving torque is absent it becomes
dTr,mec
0 ' J@ % B @ Tr,mec % Tc (20)
dt
For each of the recorded points dT/dt - )T/)t is calculated. Ideally three points
should be sufficient to obtain the model parameters, but to ensure a better fit to the
measured data all the measured points should be used. A function is defined as
dTr,mec
0 ' J@ % B @ Tr,mec % T c
dt
\ (21)
N
)T
F ' j J@ % B @ Ti % Tc
n'1 )t i

-10-
March 1, 2001,
and then an optimizer is used to find the minimum of the function F.

1.2.3 Which method to select?

No clear guide exists to which method to select. In theory, the swinging


pendulum should give the best results on the moment of inertia, but it does not give
dry-friction or the viscous-friction. Method 1 of the run-out test also gives the
possibility of accurately measuring the moment of inertia and it also gives the
viscous-friction. The dry-friction must be found afterwards by comparing a simulated
run-out test with a measured. Alternatively the dry-friction may be found by
attaching an arm to the shaft and pulling on it with a dynamometer. This, however,
does not work well when the stiction is large and/or the dry-friction is small
(relatively speaking). Method 2 of the run-out test gives all the mechanical
parameters, but because it relies on the calculation of dT/dt , it is possibly
susceptible to noise in the measurement. This is why all the measured points should
be used to determine the model parameters. Another problem exists, because the
left-hand side of (20) is zero, and therefore the solution J=B=Tc=0 exist. There will
be another solution. Care must be taken to eliminate this result. The best method
for finding the mechanical parameters is, possibly, a little of all the mentioned
methods.

-11-
March 1, 2001,
This Page is Blank

-12-
March 1, 2001,
Chapter 2
Electromechanical energy conversion
A basic question when starting with systems, which convert electrical energy
into mechanical energy, is: Why electromagnetic energy conversion, why not
electrostatic energy conversion?. This question may easily be answered by looking
at the energy in the two fields. In Figure 10 is shown a basic capacitor and
inductor.

Figure 10 Comparison between electrostatic and


electromagnetic energy.
The energy in a electric and a magnetic fields may be calculated as
1
We ' @ C @ V2
2
1 A
' @ ,r @ , 0 @ @ V2
2 d
(22)
1
Wm ' @ L @ I2
2

1 0 @ r @ N 2 @ h b
' @ @ ln @ I2
2 2@B a
For the benefit of this argument, the volumes of the inductor and capacitor are
approximately the same. As is seen form (22), the energies are proportional to 0 and
,0. The value of the base permitivity and base permeability are

-13-
March 1, 2001,
F
,0 ' 8.854 @ 10&12
m
(23)
H
0 ' 4 @ B @ 10&7 ' 1.257 @ 10&6
m
and the relative permitivity and relative permeability are in the ranges of
,r ' 1 6 70
(24)
0 ' 1 6 200000
depending on the materials. Because the base permeability in (23) is more than five
orders of magnitude larger than the base permitivity and because the relative
permeability nearly always is much larger than the relative permitivity, the energy
in the magnetic field will always be much larger than the electric field independent
on all other parameters.

2.1 The Lorenz-force and Faradays law


At this point, it is known how to solve the mechanical equation for rotating
motion. It is also known that the magnetic field is the best choice for generating a
force to drive the system, due to the fact that the energy-density in the magnetic
field is much larger than the electrical field. However, understanding that the
energy-density is larger, does not help explaining the coupling between the magnetic
fields and the electrical circuits. To do this, two phenomena are used, being The
Lorenz-force and Faradays law.

2.1.1 The Lorenz-force

First is examined what happens when a current carrying is placed inside flux
density field with an approximately linear distribution. The flux density may be
achieved by a setup as shown in Figure 11 , where the field is driven by a
permanent magnet.

-14-
March 1, 2001,
Figure 11 A permanent magnet
(light blue) generating a magnetic
flux (magenta] in an air-gap
created by an iron yoke (blue).
It is known from electromagnetic field theory that the force acting upon a electrical
charge placed into a magnetic field is
F'q@vB (25)

where q is the charge. Picturing a stream of moving charges, a situation like the
moving electrons in a conductor arrises as shown in Figure 12 . The direction of the
force is also easily found by applying The Left-Hand Rule.

Figure 12 On the left-hand side is shown a current-carrying


conductor in a magnetic field and on the right-hand side is
shown The Left-Hand Rule.
In this case where a current carrying wire is placed in the magnetic flux, the force
will become
F'B@I@l (26)

if the flux is uniformly distributed and the length is known.It is, however, important
to remember that a current-carrying conductor in itself generates a magnetic flux.
The two fluxes are then super-imposed on each other. If a small loop of wire is placed
inside the magnetic flux and if a current is sent through the wire, a situation as
shown in Figure 13 arrises.

-15-
March 1, 2001,
F

Figure 13 A current-carrying loop in a


magnetic field.
Compared to the flux shown in Figure 12 , the flux becomes distorted due to the
current in the loop. As is seen in Figure 13 , the forces acting on the loop points in
the direction of lower flux density. If the radius of the loop is known, the torque
acting on the loop may also be calculated as
T ' 2@F@r
(27)
' 2 @ B @ I @ l @ r @ sin2
where 2 is the angle between the flux produced by the loop and the field.

2.1.2 Faradays law

In (25) was shown what happens when a current-carrying coil was placed in
a magnetic field. If a short-circuited coil is placed in the magnetic field and then
moved at a constant speed inside the field, much the same thing happens as with the
Lorenz-force. In this case the electrons in the coil are subjected to a force, and they
will start to move and thereby creating a current in the coil. If the coil is open, no
current can flow. In this case there will be a surplus of electrons at one end of the
coil and a deficit at the other. This causes a voltage to form at the ends of the coil.
This is shown in Figure 14 .

-16-
March 1, 2001,
v

Figure 14 On the left-hand side is shown a short-circuited coil


moving at the speed v in a magnetic field and on the right-hand
side is shown The Right-Hand Rule.
The magnitude of the induced voltage may be determined from Faradays law as
d8
e ' &
dt
(28)
dN@A@B
e ' &
dt

and the orientation of the induced voltage or electromotive force (EMF) may be found
by applying The right-hand rule. Two special cases arises from (28): A field with
constant flux and with a moving coil as opposed to a stationary coil with a changing
flux.
Case 1: Constant flux, moving coil
dN@A@B
e1 ' &
dt

' &N @ B @ d A
dt

' &N @ B @ d b @ l
dt (29)
' &N @ B @ l @ d b
dt

' &N @ B @ l @ db ,
db
' vt
dt dt
' &N @ B @ l @ v t
where l and b are the length and breadth of the coil and db/dt is the speed by which
it moves.

-17-
March 1, 2001,
Case 2: Changing flux, stationary coil
dN@A@B
e2 ' &
dt
(30)
dB
' &N @ A @
dt
In many cases the induced voltage in the coil will be caused due to a change in the
flux and due to movement of the coil as
e ) ' e1 % e2 (31)

Another thing which influences the induced voltage is the angle between the coil and
the field. When the coil is parallel to the field no voltage is induced, because none of
the field cuts through the window of the coil. When the coil is perpendicular to the
field the maximum voltage is induced because the maximum flux goes through the
window of the coil. This is shown in Figure 15 .

Figure 15 The angle between the coil and the field.


All in all the induced voltage becomes
e ' e ) @ sin2 (32)

Due to the fact that the induced voltage is caused by a change in the flux, it is also
known as the electromotive force or EMF for short.

2.1.3 Acting together

Because Faradays law and the Lorenz-force arrises from the same
phenomenon, they are inseparable. A change in the flux will cause a voltage in a coil
and a voltage applied to the coil change the flux. The same way with the forces, if a
current is applied to a coil, it will form a flux, which will produce a force acting on
the coil and a force acting on a coil in a magnetic field may cause a current to flow.
Therefore for all but some very special cases, there will be forces and there will be
voltages and there will be currents. Therefore, as shown in Figure 16 , all circuits

-18-
March 1, 2001,
in a magnetic field will have an induces EMF and all circuits with a current will be
subjected to a force.

Figure 16 A basic circuit with an induced


EMF.

2.2 Work, force and torque


Everywhere, when one form of energy is transformed into a different kind,
there must be a balance. The laws of thermodynamics state that all the energy must
be accounted for. Energy supplied to a system is either lost as heat or as a
permanent phase-transformation, stored or transmitted. Using a simple electro-
mechanical system as the one shown in Figure 17 , this becomes evident.

Figure 17 A system with a electrical part, magnetic part and


mechanical part.
The energy-flow of the system may be explained by looking at Figure 18 .

-19-
March 1, 2001,
WE

Figure 18 The energy balances.


Explaining Figure 18 from left to right: Electrical energy is supplied to or taken
from the system. The electrical energy may go to two things, either changed into
magnetic energy or lost in the resistance of the wires. The magnetic energy may be
changed back to electrical energy, lost as iron losses in the magnetic object or
changed into mechanical energy. The mechanical energy may be changed back to
magnetic energy, lost as mechanical losses or transferred to a mechanical load. It is
important that the energy flow goes in both directions, the only energies that do not
are the loss-energies which under normal conditions may not be reclaimed. The
change in energy is therefore
dWe ' dWf % dWm (33)

The method for finding the work done on the system then becomes a matter of
determining the change in energy of the different parts of the system.

2.2.1 Energy and co-energy

Because everything in electromagnetic energy transformation links through the


magnetic field and the magnetic materials, is it important to be able to calculate the
energy. In many cases is it sufficient to regard the magnetic materials as linear
totally without hysteresis and saturation. This, however, it not the case, but is an
useful approximation. For a linear material as the one shown in Figure 19 there
is a simple relation between the current and the flux linkage.

-20-
March 1, 2001,

i
Figure 19 The definition of linear inductance.
The inductance may be calculated as
)8
L ' (34)
di

However, as is often the case, the materials are not linear. It is therefore convenient
to use the energy and the co-energy to calculate the changes in energy. The energy
and co-energy in a magnetic material is defined as shown in Figure 20 for both
linear and non-linear material.

Figure 20 On the left-hand side is shown the energy and co-energy for linear material
and on the right-hand side is the energy and co-energy shown for material with
saturation.
As may be seen from Figure 20 on the left-hand side, will the co-energy increase in
proportion to the energy with an increase in the relative permeability and the energy
will decrease with the same amount. For non-linear materials are the relation

-21-
March 1, 2001,
between energy and co-energy not as straight forward, due to the fact that the
permeability changes with the scale of the current. Further complicating the matter,
most magnetic materials also exhibit hysteresis together with saturation. This is
shown in Figure 21 .

Figure 21 On the left-hand side is the energy and co-energy shown for material with
saturation and on the right-hand side is shown the energy and co-energy together with
the iron losses for material with saturation and hysteresis.
The energy lost as iron losses must be accounted for in the energy balance. In the
following are the magnetic materials linear and without hysteresis unless otherwise
stated.
The definition of energy and co-energy is useful, because they may be used to
describe how the energy is transferred, e.g. from magnetic to mechanical energy.

2.2.2 The electric circuit

It is not the aim of this text to explain calculations on electrical circuits. It is


sufficient to state that the energy lost in resistors is
Wel ' R @ i t 2 dt (35)

For a capacitor the energy stored is

-22-
March 1, 2001,
duC
iC t ' C@
dt
\
pC t ' uC t @ i C t
duC
' uC t @ C @
dt
\ (36)
t
WC t ' pC t
0

t
duC
' uC t @ C @ dt
dt
0

1
' @ C @ uC t 2
2

and for an inductor the energy stored is


di
uL t ' L @ L
dt
\
pL t ' i L t @ uL t
di L
' iL t @ L @
dt
\ (37)
t
WL t ' pL t
0

t
di L
' iL t @ L @ dt
dt
0

1
' @ L @ iL t 2
2

-23-
March 1, 2001,
2.2.3 The magnetic circuit

Many methods for doing calculations on electrical circuits exist. It would be


nice if a similar method for doing calculations on magnetic circuits were available.
To find if this is the case the example from Figure 22 is used.

Figure 22 A coil on a magnetic toroidal core


with an air-gap.
Using Ampres circuital law around the contour C
H @ dl ' N @ I0 (38)
C

To simplify the matter, the leakage flux is neglected. When there is no leakage flux,
the same flux will flow in both the core and the air-gap, however, due to different
permeability in the core and air, the magnetic field will be different. Therefore
B f ' Bg ' aN @ Bf
Bf
H f ' aN @
(39)
Bf
Hg ' aN @
0

where aN describes the path inside the core. Substituting (39) into (38)

-24-
March 1, 2001,
Bf Bf
N @ I0 ' @ 2B @ r0 & lg % @ lg
0
\
0 @ @ N @ I0
Bf ' aN @
0 @ 2B @ r0 & lg % @ lg
(40)
\
0 @ N @ I0
Hf ' aN @
0 @ 2B @ r0 & lg % @ lg
@ N @ I0
Hg ' aN @
0 @ 2B @ r0 & lg % @ lg

and since Hg/Hf ' /0 , the field intensity in the air-gap is much stronger than in the
core. If the radius of the cross-section compared to the radius of the toroid is much
smaller, the flux density inside the core is approximately constant.
MB@A (41)

This then gives the total flux as


N @ I0
M '
2B @ r0 & l g lg
%
@A 0 @ A

'
U f % Ug
(42)
\
2B @ r0 & lg lf
Uf ' '
@A @A
lg
Ug '
0 @ A

looking at (42), this is much as an electric circuit with a voltage source, N @ I0 and
some resistors, Uf and Ug , resulting in a current, M , flowing through them. This is
shown in Figure 23 .

-25-
March 1, 2001,

Figure 23 the similarity between a electrical circuit and


a reluctance circuit.
The term N @ I is known as the ampere-turns. If the physical dimensions and the
permeability of the magnetic sections is known, the flux is easily found. The method
may be expanded to shapes much more complicated than the one shown in Figure
22 . A nice thing is that the current from the electrical circuit is in the reluctance
circuit and generates a coupling between the two systems. What may complicate
things, is that the permeability of the magnetic sections may change due to the
strong non-linearity of the magnetic materials.

2.2.4 The mechanical circuit

It was shown, in the previous chapter, how the mechanical part of a device
exhibit motion, if a force or torque is precent. It was also shown how the Lorenz-force
may be used to generate a force on a winding. It is, however, not yet clear how this
translates to a system as the one shown in Figure 17 , where the forces obviously it
generated in the air-gap and not on the winding. To better understand this, is in
Figure 24 shown the fluxes and resulting forces acting on two permanent magnets
in close proximity to each other.

-26-
March 1, 2001,
Figure 24 Forces acting on two permanent magnets in
close proximity to each other.
The resulting forces will try to align the magnets and close the air-gap. The
alignment of the magnets increases the area of the air-gap and closing the air-gap
will reduce the length. Examining again the expression in (42), it is seen that
shortening the length and increasing the area will reduce the reluctance of the air-
gap. In Figure 25 is essentially the same figure as in Figure 17 shown, only now
is the yoke displaced by the distance )x.

Figure 25 The yoke is displaced the distance )x.


By displacing the yoke in Figure 25 , the area of the air-gap does not change, but the
length does change. As is known from (42), will the reluctance of the air-gab change
and thereby also the flux. To find the forces, equation (33) is used.

-27-
March 1, 2001,
dWe ' dWf % dWm
dWm ' F m @ dx (43)
dWe ' i @ v @ dt ' i @ d8

The change in energy may come from two places, either electrical as a change in the
current or mechanical as a change in distance, there exists two independent variables
i and x. The change in flux linkage d8 then becomes
M8 M8
d8 ' @ di % @ dx (44)
Mi Mx

This gives the change in magnetic energy as


MWf MWf
dWf ' @ di % @ dx (45)
Mi Mx
and the energy balance as
MWf M8 MWf M8
F m @ dx ' & @ dx % i @ @ dx % & @ di % i @ @ di (46)
Mx dx Mi di

Four cases exist: Case 1, there is no movement of the yoke. Case 2, when the yoke is
moved, there is no change in the flux linkage. Case 3, when the yoke is moved, there
is no change in the current. Case 4, the general case, where both the flux linkage and
the current changes.

Case 1:
A special case exists when the yoke is not moved.
MWf M8 MWf M8
Fm @ dx ' & %i@ @ dx % & %i@ @ di
Mx dx Mi di
dx ' 0
(47)
\

MWf M8
0 ' & %i@ @ di
Mi di

and then (47) becomes Faradays law because the change in energy only is between
the electrical and magnetic parts.

Case 2:

-28-
March 1, 2001,
The yoke is moved quickly, and because the flux cannot change instantly, the
current must therefore change. This is shown in Figure 26 .

Reducing
the air-gap

Figure 26 The change in magnetic energy


when changing the air-gap quickly.
When the flux does not change, there cannot be induced any voltage in the winding
and therefore there cannot be transferred any energy to or from the electric circuit
to the magnetic and the energy goes directly to the mechanical circuit. The energy
balance then becomes
MWf M8 MWf M8
Fm @ dx ' & %i@ @ dx % & %i@ @ di
Mx dx Mi di
d8 ' 0
(48)
\
MWf MWf
Fm @ dx ' & @ dx % & @ di
Mx Mi

Case 3:
The yoke is moved slowly, and because it is slow, the flux now has the time to
change. In this case is the current kept constant by some means. This is shown in
Figure 27 .

-29-
March 1, 2001,

Reducing
the air-gap

Figure 27 The change in magnetic energy


when changing the air-gap slowly.
Because the flux changes, there will be induced a voltage in the electric circuit and
there may be transferred energy from the electric circuit to the magnetic. The energy
balance then becomes
MWf M8 MWf M8
Fm @ dx ' & %i@ @ dx % & %i@ @ di
Mx dx Mi di
di ' 0
(49)
\

MWf M8
Fm @ dx ' & %i@ @ dx
Mx dx
Case 4:
The cases where there is no change in the flux linkage or the current is unusual,
because it would require some special circumstances for them to appear. In most
cases there will be a change in both the current and flux linkage as shown in Figure
28 .

-30-
March 1, 2001,

smaller
air-gap

Figure 28 General change in energy due to


change in air-gap.
For this case the resulting force may only be calculated by using (46).

The four cases above are only correct for linear materials. As is the case, magnetic
materials are non-linear and the energy balances above are only approximations. For
non-linear materials a change in air-gap still results in a change in energy, as shown
in Figure 29 .

Figure 29 The change in magnetic energy


due to the change in air-gap.
The best method for finding the resulting force is in this case to calculate the change
in energy in Figure 29 by integrating the area. Using this method, it is possible to
calculate the forces with good accuracy. This general case then becomes

-31-
March 1, 2001,
MWf i, x M8 i, x
Fm ' & %i@ (50)
Mx Mx
In many cases it is desirable to regard the magnetic materials as linear, at
least for an initial calculation. It is known that the energy in an inductor is
1
2
@ i 2 @ L x , therefore the change in energy due to a change in air-gap then becomes
)Wf 1 2 )L
' @i @
)x 2 )x
(51)
MWf 1 2 ML
' @i @
Mx 2 Mx

where L is not a function of the current. Moreover, since 8 ' i @ L then


M8 ML
i@ ' i2 @ (52)
Mx Mx

which then gives the force as


1 2 ML ML
Fm ' & @i @ % i2 @
2 Mx Mx
1 2 ML
' @i @ (53)
2 Mx
1 MU
' @ M2 @
2 Mx

-32-
March 1, 2001,
This Page is Blank

-33-
March 1, 2001,
Chapter 3
Introduction to the DC-motor
The DC-motor has been and still is one of the work-horses in industry and
modern house-holds. The motor is not cheap, because it has a complicated winding
in the rotor (armature) and has a commutator with brushes to feed current to the
rotor. Its axial length is also large, because the commutator adds to the length, but
does not contribute to the generation of torque. However, the motor is very much
liked, because it is very easy to control both with regard to torque and speed. The DC-
motor is used often as the motor to compare other motors with. The DC-motor is old-
fashioned, but has some very interesting special couplings, which gives it some
distinct and desirable qualities.
As an overview the following may be said on the DC-motor:
Pro Con
Easy to understand The motor is large because the
commutator adds to the length
Easy to model It must be maintained, because the
brushes ware down
Used as the reference for all other There is always the possibility of
motors sparking on the commutator
It is possible to make it as low cost if Unless the motor is low cost, a
the motor is small number of compensating windings are
required, which adds to the price

3.1 The archetypical DC-motor


Some very fundamental phenomena characterise the DC-motor, which are
closely related to the electro-mechanical energy transformations from the previous
chapter.

3.1.1 Generating torque in a DC-motor

One of the most essential qualities of a motor is the ability to generate torque

-34-
March 1, 2001,
(or force for linear motors). If a current-carrying loop is placed in a magnetic field as
shown in Figure 30 , a set of forces will act on it and create a torque.

Figure 30 Forces acting on a current-carrying loop in a


magnetic field.
The forces are known from the Lorenz-force as
F ' B@i@l (54)

and when the radius of the loop is known the torque becomes
T ' 2@B@i@l@r (55)

and if the loop consists of a winding with N turns, the torque becomes
T ' N@2@B@i@l@r (56)

If one winding produces a torque of a specific size, then many windings will produce
higher torque. A way of placing multiple windings is shown in Figure 31 .

Figure 31 The forces acting on a set on current-carrying loops


in a magnetic field.
The arrangement shown in Figure 31 is known as the armature. It is, however, only
the component of the magnetic flux density perpendicular to the axis of the individual

-35-
March 1, 2001,
coils that produces torque. For each of the windings the torque is
T ' N@2@B" @i@l@r (57)

where B " is the perpendicular component of the flux density with respect to the
individual loops. For all the windings together the torque becomes
I

Te ' N @ 2 @ B "i @ i @ l @ r (58)


i'1

and because l and r describes the physical dimensions of the loops, the total torque
may be described by
T e ' kt @ Mf @ I a (59)

where Ia is the current flowing in the armature and ka becomes a factor describing the
physical shape of the winding.. The question is then: How are the individual
windings connected to get the currents flowing as shown in Figure 31 ?. The answer
is to use a commutator as shown in Figure 32 .

Figure 32 The commutator on a DC-motor.


The current is then fed to the armature by a set of brushes running on top of the
commutator. For a two-pole machine, two brushes are used and for machines of a
higher pole-number, the number of brushes usually corresponds to the pole-number.
The brushes are usually made from carbon, but may occasionally be made from other
another material like cupper, bronze or brass. When the armature turns, the brushes
cause the current to change direction in the windings so that the state shown in
Figure 31 is maintained. The situation then is as shown in Figure 33 .

-36-
March 1, 2001,
Figure 33 The individual windings of a DC-motor fed through
a commutator.
There exist a number of ways to connect the armature winding. The two most
commonly used is the lap and the wave winding. The advantages of the two windings
are shown next.

-37-
March 1, 2001,
3.1.2 Lap or wave winding

The decision on using a lap winding or a wave winding depends on the size of
the motor, the current carrying capacity of the brushes and commutator and the
voltage which the motor has to use.
In the lap winding the ends of the coils are connected to commutator bars close
to each other. The induced EMF in each side of the coil must be added together and
therefore they must be located under poles of opposite polarity. Furthermore, if the
size and arrangement of the coils are to be uniform, the number of slots spanned by
a coil must be an integer value less or equal to the total number of slots divided by
the number of poles. From this the pitch of the coil in a lap winding may be
calculated by
C
yb ' k (60)
p

where k ' 0, 1, 2.... . Using an example with 24 slots, 4 poles and k=1 this gives
yb'11 or 13 . Using yb'13 this gives a layout of the armature winding as shown in
Figure 34 .

Figure 34 Four-pole armature lap winding.

-38-
March 1, 2001,
The coils a, g, m and s in Figure 34 are all located under one pole and the EMF of
each of the sides of the coils cancel each other out. This is therefore the ideal
placement of the brushes. Showing the connection of the lap winding in Figure 34
in a schematic form, it is seen that the connection is as shown in Figure 35 .

Figure 35 The connection of a four-pole lap-wound


armature shown in simplex form.
It is also seen from Figure 35 that the number of parallel paths for the current is
equal to the number of pole-pairs. It will therefor be practical to use the lap-wound
armature for a low-voltage high-current motor, because the stresses on the winding,
brushes and commutator are reduced due to the multiple parallel paths and because
the generated torque is proportional to the current and not the voltage.
In a multi-pole DC-motor the air-flux repeats itself after each pole-pair. It is
therefore possible to connect two sides of a coil approximately two pole-pitches away
and yet the voltage across two adjacent commutator bars will be nearly the same as
with the lap winding. Coming back after going around the armature, the winding
formed by the series of coils may not close on itself, because this will short-circuit the
armature. The connection may therefore be at a commutator bar further ahead on the
commutator. A connection like this is known as a wave winding. Going around the
armature consecutive pole-pairs may be connected in series and their EMF will be
additive. Going around the armature, the winding may not close on itself because this
will be short-circuiting the armature winding. The winding must be connected to a

-39-
March 1, 2001,
commutator bar before or after the one that was the start. What must be determined
is the pitch of the coils around the armature. This is found from
C 2
y b % yf ' (61)
p/2

where C is two times the slot number, yb is the back pitch (or the first pitch) and yf
is the front pitch (or the next pitch). The coil pitch is
y 1
yc ' b (62)
2
Using 21 slots is a four-pole machine this gives yb % yf ' 20 or 22 and yc ' 5 or 6 .
Selecting yb ' yf ' 11 and y c ' 5 this gives a layout of the armature as shown in
Figure 36 .

Figure 36 Four-pole armature wave winding.


Note that for the wave winding there are p/2 coils connected in series between each
commutator bar and not like the lap winding where there only is a single coil. This
also means that there are 1/p of the total number of coils between two adjacent
brushes. Between two adjacent brushes there therefore are p/2 @ 1/p ' 1/2 of all the
coils independent of the pole number, always. This is also seen by looking at the coil

-40-
March 1, 2001,
in Figure 36 highlighted with red. The current starts at commutator bar 1 and
travels through the coil a to commutator bar 12 and through the coil l ending up at
the commutator bar 2. Continuing through the remaining commutator bars and coils,
it is found that the polarity of the induced EMF does not change until half the coils
remain. The conclusion is therefore that a wave winding may be commutated by only
two brushes. However, additional brushes may be placed at the other points of
neutral, because it increases the contact area on the commutator and the commutator
therefore becomes shorter. The required brushes are shown in Figure 36 in blue and
the additional brushes in green. Tracing out the armature winding as shown above,
it may be re-drawn as shown in Figure 37 .

Figure 37 Coil connection of a four-pole simplex wave-wound


armature.
It may occur in a wave-wound armature with only two brushes that differences in the
induced EMFs due to small differences in the air-gap or unequal pole flux will cause
a sizable circulating current to appear. This current may be reduced by inserting
equi-potential connections in the armature winding from the points of neutral of
equal potential.

3.1.3 EMF of a DC-motor

As was shown in the previous chapter, whenever there is a change in the


magnetic field due to a time varying field or a moving circuit, there will be an
induced voltage. The same is especially true for the DC-motor, because of the
magnetic field and the rotating armature. This is shown for a single winding in
Figure 38 .

-41-
March 1, 2001,
Figure 38 A coil rotating with the angular speed T in a
homogeneous flux density.
As the coil rotates, the induced EMF changes as
d8 dB" @A@N
EMF ' ea ' & ' & (63)
dt dt

where B " is the perpendicular component of the flux density passing through the
winding, A is the area of the loop and N is the number of turns. Expanding this to all
the windings in the machine, a situation arrises as shown in Figure 39 .

Figure 39 The EMF generated in the armature-winding of a


DC-motor.
and the EMF for the full winding becomes
N
d B "i @ A @ N
ea ' &
dt (64)
i'1

' k e @ Mf @ T
where ke is a constant describing the physical shape of the winding.

-42-
March 1, 2001,
3.1.4 Torque constant, voltage constant and field constant, kt, ke and kM

An interesting question is: Is there a relation between ka and ke?. Yes there
is. The electrical power is found as
P e ' ea @ i a ' ke @ Mf @ T @ i a (65)

and the mechanical power is found as


P m ' T @ T ' k t @ Mf @ i a @ T (66)

and in steady-state the power balance is


P m ' Pe (67)

and therefore ke and kt be numerical must equal as


Nm V
kt ' ke ' kM
A @ Wb rad (68)
Wb @
s
where kM is known as the field constant.

3.2 The DC-motor


In the previous section the fundamentals for a DC-motor were shown.
However, the description is not sufficient to describe how a DC-motor works. This is
because there has been no mentioning of what generates the magnetic field and how
the armature is connected to the voltage supply. Another thing is that a number of
very specific types of DC-motors exist. They are: Separately excited DC-motor,
permanent magnet DC-motor, shunt motor, series motor, compound motor and the
universal motor. There has been no mentioning how these special motors appear. The
most general of the motors is the separately excited DC-motor and this will be used
to describe how a DC-motor works in general. Later the other types will be described
separately.

3.2.1 Separately excited DC-motor

Previously was shown how the armature was connected, however, there has
been no mention on how the magnetic field is produced. The simplest way of
producing the magnetic field is by using a second winding supplied by a separate
voltage supply. This is shown schematically in Figure 40 .

-43-
March 1, 2001,
Figure 40 Separately excited DC-motor.
The separately excited DC-motor in Figure 40 is very versatile, because, by
controlling both Va and Vf, it is possible to control both the torque and speed
independently. However, it must be remembered that the field winding is wound on
a magnetic core and as such saturate at a given flux density. Because of this, there
will be a non-linear relationship between the current in the field winding and the
resulting flux as shown in Figure 41 .
f

If
Figure 41 The relationship between field
current and the resulting flux in a separately
excited DC-motor.
The relationship shown in Figure 41 is found by driving the motor with another
motor at some angular speed, T, and then slowly increasing the field current. If T is
known, the relationship between If and M may be found. The path for the flux from
the field winding usually has a large air-gap. This is the reason for the flatness of the
curve in Figure 41 .
The separately excited DC-motors have one distinct advantage. If the desired
torque is lower than the maximum torque, it may be seen from (59) that different
values of If and Ia may produce the same torque. Using this, it is possible to run the
motor at a high efficiency for any given torque. This is because resistive power loss
in the armature winding and field winding is reduced by selecting specific currents.
This is especially advantageous when the energy source is limited such as in an
electrical vehicle. It is also an advantage that the current in the armature winding
may be high and the current in the field winding may be low. The speed of the DC-

-44-
March 1, 2001,
motor may therefore be controlled by a circuit with a low current-carrying capability.
It may also be seen form (59), that by reversing the voltage on either Va or Vf, the
direction of the driving torque is also reversed

3.2.2 Shunt motor

Although the separately excited DC-motor has some distinct advantages, it is


not always desirable to have to control two voltage sources. The simplest way of doing
away with the second voltage source is simply to connect the field winding in parallel
to the armature winding as shown in Figure 42 .

Figure 42 Shunt motor.


By connecting the motor as a shunt motor, a specific relation between the field
current and the armature current is introduced. The field current is only dependent
on the supply voltage, and if the supply voltage is sufficiently stiff, the armature
current cannot influence the field current. This relation gives the torque and load
characteristics shown in Figure 43 and Figure 44 .

Figure 43 Load characteristic of a shuntFigure 44 Torque characteristic of a shunt


motor. motor.
The characteristics are relatively flat, because the armature current may change
without changing the field current much. The only reason that the characteristics are
not fully flat, is because of the resistive voltage drop in the armature winding.

-45-
March 1, 2001,
Because the two characteristics are flat, the shunt motor is very well suited to
applications where a constant torque is required, independent on the speed. This
could be machine-tools, centrifugal pumps or a conveyor belt.

3.2.3 Series motor

A different option is to connect the field winding in series with the armature
winding. This is shown in Figure 45 .

Figure 45 Series motor.


For the series motor the field and armature current is the same. When the motor is
loaded, the angular speed will be decreased and when the angular speed is decreased
the induced voltage will also be decreased. This will cause the armature current to
increase and looking at (59) it is seen that the developed torque will increase. This
relation gives the torque and load characteristics shown in Figure 46 and Figure
47 .

Figure 46 Load characteristic of a seriesFigure 47 Torque characteristic of a shunt


motor. motor.
The series motor has two very distinct advantages: It has high starting torque and
when the speed drops due to loading, the torque increases. The problem with the
series motor is that when it runs close to no-load, the angular speed may be very
high. The motor is very useful for use as locomotion in hoists and locomotives. As

-46-
March 1, 2001,
general rule the load should never be less than 15% of full-load, otherwise a run-
away situation may arise.

3.2.4 Compound motor

The next question is: Is it possible for a motor to have both a shunt winding
and a series winding?. This is possible and is known as the compound motor. The
compound motors combines some of the characteristics on the series motor and shunt
motor. The compound motors may be coupled as commutative compound motor where
the flux from the shunt and series winding work together or as differential compound
motor where the two fluxes oppose each other. The motor may also be coupled as a
long shunt compound motor where the shunt winding is parallel to the series winding
and armature or as a short shunt compound motor where the shunt winding is
parallel to the armature only. This is shown in Figure 48 .

Figure 48 On the top left-hand side is a commutative long shunt compound motor, on the
top right-hand side is a differential long shunt compound motor, on the bottom left-hand
side is a commutative short shunt compound motor and, on the bottom right-hand side is
a differential short shunt compound motor.
The different compound motors have some distinct characteristics, where they
combine to a different degree the characteristics of a series and a shunt motor.
Most of the motors used either are separately excited DC-motors, series motors
or shunt motors. A compound motor is only used when a specific characteristic is
desired and an active control of the motors is undesirable.

-47-
March 1, 2001,
3.2.5 Permanent magnet DC-motor

Obviously, a permanent magnet may be used to create the magnetic flux. The
reasons for using a permanent magnet are: It is possible to make the DC-motor as
low-cost, because there is no need for a field winding and when there is no field
winding, there are no losses associated with it. kM is always the same, and the control
becomes easy, because the system may be described by two time-constants. However,
some disadvantages also exist: Permanent magnets cannot produce as high a flux
density as a field winding and in case of short-circuiting, the permanent magnets
may be partially de-magnetized due to the high currents.

3.2.6 Universal motor

What might sound as a silly question is:What kind of voltage does a DC-motor
run on?. One answer to the question is offcourse DC-voltage, but it is not the whole
answer. Using the series motor as an example and supplying it with an AC-voltage
with the same RMS-value as the DC-voltage. By using the left-hand rule, it is seen
that during the positive half-period, flux from the field winding and the armature
current produces torque in one direction. When the voltage is reversed in the
negative half-period, both the armature current and the field from the field winding
changes and the torque is still produced in the same direction as for the positive half-
period. The motor therefore produces a torque in the same direction for the whole
period, because the current is the same in both armature and field winding. The
developed torque will be pulsating due to the sinusoidal voltage, but because the
mechanical time-constant usually is much larger than both the electrical time-
constant and the period of a 50 Hz voltage, this is not a big problem. The voltage used
to drive the universal motor may be much different form sinusoidal and the motor
therefore is very robust in an application. Using an AC-voltage with shunt motors
and compound motors will also work, only not as well. This is because the time-
constants of the different windings are different and the currents in the armature
and field winding are not the same. The developed torque may be negative for some
of the positive and negative half-period. Using an AC-voltage on a permanent magnet
motor does not work, because the field does not change and the developed torque will
be alternating positive and negative. One thing that makes the universal motor
different from a series motor is the need for a laminated core in the field winding.
Usually in a DC-motor the magnetic core in the field winding is made from solid
magnetic material or thick laminations, but because of the AC-voltage, there will be
large eddy-currents if the core was not made from thin laminations. The universal
motor has many of the same characteristics as other DC-motors, but does not need
a rectifying circuit. The universal motor is usually used in devises where a regulated
speed is desired, but where the cost must be low. The universal motor is used in

-48-
March 1, 2001,
hand-drills, washing machines, kitchen machines ect.

3.3 Making the DC-motor work


In the previous two sections was shown how a DC-motor work in general and
how the different types of DC-motors work. However, so far it has been assumed that
the armature current cannot distort the flux distribution from the field winding and
that the commutations of the armature windings are ideal. Because this is not so,
special compensation windings must be used.

3.3.1 Armature reaction

The general idea was that the current in the armature did not influence the
distribution of the flux from the field winding as shown in Figure 49 .

Figure 49 The field winding of a DC-motor.


This, however, is only approximately true in a no-load situation, where the armature
current is very small. When the armature current is not small, there will be induced
a flux in the field poles by this current as shown in Figure 50 .

Figure 50 The armature reaction shown with blue.


The green line shown the shift of neutral axises due to
the armature reaction.

-49-
March 1, 2001,
The problem is that on one side of the field pole the armature reaction adds to flux
density and on the other side is subtracts from the flux density. The brushes must be
located at the point of neutral on the commutator, because it otherwise will cause
sparking at the brushes. The problem is, that the armature reaction moves the point
of neutral as shown in Figure 51 , where the armature and yoke is unfolded.

Figure 51 On the top is shown the armature and yoke


unfolded. On the bottom is shown the flux densities. Cyan
line is the flux density due to the field winding, red line is
the armature reaction at load and blue line is the resulting
flux density.
In some older DC-motors the brushes are mounted so they may be moved to the point
of neutral depending on the load. The voltage over the brushes is measured and the
point where the maximum voltage is measured is the point of neutral. It should be
noted that the armature reaction moves according to the load and with the rotational
direction. When the direction changes, so does the point of neutral.
Another problem with the armature reaction is that it changes the maximum
flux density in the field poles. For linear materials, this only causes that the point of
neutral moves, however, for non-linear material there is an additional problem. The
ampere-turns of the armature are added or subtracted equally to the flux density
from the field, but because of the non-linear material, this will not result in the same
increase in the flux density as the decrease. This is shown in Figure 52 .

-50-
March 1, 2001,
B
+B
-B

N@I N@I N@I


Figure 52 The armature reactions influence
on the flux density level.
Because %)B and &)B in Figure 52 are not equal, it will cause a decrease in the
average flux and therefore also a decrease in the induced voltage in the armature
winding.
The solution to the problem of armature reaction is to insert a compensating
winding in the pole-face of the field-poles as shown in Figure 53 .

Figure 53 On the top is shown the compensation winding


inserted in the pole-face and on the bottom is shown
schematically the connection of the compensation winding .
The compensating winding is connected in series with the armature. The
compensating winding is tuned to reduce the armature reaction and the DC-motor
with the compensation winding therefore behaves more like the archetypical DC-
motor as described previously. It is usually not possible to fully compensate for the
armature winding, but in series, shunt and compound motors, where the rotational
direction is fixed, the position of the brushes may permanently be moved to partly

-51-
March 1, 2001,
compensate for the full-load armature reaction. There will usually not be a
compensation winding in permanent magnet DC-motors and in low-cost DC-motors,
however, it is possible for shunt and series motors to displace the brushes
permanently as a compromise between no-load and full-load, because the motor
always turns the same way.

3.3.2 Incomplete commutation

The armature reaction is not the only thing that may cause sparking on the
commutator. As is seen in Figure 54 , when the brush passes on the commutator, the
current in the individual winding sections must change direction.

brush

Figure 54 The change in direction of the


current when the brush passes on the
commutator
Because any circuit has an inductance, there is a limit to how quickly it is possible
to change the direction of the current. Furthermore, because the armature reactance
is highest in the inter-polar region, there may be induced an extra voltage in the
winding sections. Ideally, the commutation should happen when the brush passes on
the commutator. The brush usually covers two to three sections of the commutator
and thereby effectively short-circuits the part of the winding being commutated.
However, if the current in the armature is so large that it has not had the time to
change fully or if the armature reaction induces a voltage large enough to delay the
commutation, this will result in sparking on the commutator. This is shown in
Figure 55 . By improper operating of the DC-motor, it is possible that sparks jump
over the whole commutator between the brushes. The sparking in it self is
undesirable and it will damage the commutator quickly.

-52-
March 1, 2001,
Figure 55 On the top is ideal commutation, in the middle is there
an incomplete commutation and on the bottom is an incomplete
commutation with voltage from the armature reaction.
The method for reducing the risk of sparking due to incomplete commutation is the
use of commutator poles. A commutator pole is a small extra pole connected as shown
in Figure 56 , where the armature current through it opposing the flux from the
armature.

commutator pole

Figure 56 On the left-hand side is schematically shown the connection of a


commutator pole and on the right-hand side is shown a picture of a DC-motor
with a commutator pole.
A question is: Is the function of the commutator pole not exactly the same as the
compensating winding for the armature reaction? They both reduce the field from the
armature. Yes, in theory either a compensation winding or a commutator pole
should be sufficient to ensure spark-less commutation. However, it has been shown
that either one of them alone cannot do the job satisfactory.

-53-
March 1, 2001,
The same way as with the compensation winding, low-cost DC-motors and
small permanent DC-motors does not have a commutator pole, but a permanent
displacement of the brushes may help the commutation in series and shunt motors.
In the case of linear commutation as shown in Figure 55 , the commutator
pole, however, cannot reduce sparking if there is not sufficient time to change the
direction of the current in the armature windings. This may happen when the motor
runs at over-speed at full-load or if the load suddenly is disconnected. (This, by the
way, looks spectacular, because sparks erupt form every opening of the motor.)

3.4 Getting motor parameters


To be able to simulate the motor, it is a necessity to have good parameters for
the motor. In chapter 1 it was described how to get the parameters to model the
mechanical system. The same way as for the mechanical system, it is necessary to be
able to correctly measure the parameters describing the electric and electromagnetic
system.

3.4.1 The brushes

The brushes feeding the current through the commutator to the armature
winding has a resistance significantly high, so that it in most cases is necessary to
include it in a simulation of a DC-motor. If not, the voltage-drop over the brushes is
large enough to give a significant difference between measured and simulated
angular speeds. The brushes are usually made from carbon. Carbon acts differently
form normal conducting materials, because it in some aspects acts more like an
insulator than a conductor. The temperature coefficient of the resistance of carbon
is large and negative, which causes the resistance to decrease when the temperature
increases. It is not easy to precisely measure the resistance of the brushes, but as a
rule-of-thumb is the voltage-drop over the brushes in the range of 0.5-1.5V. This
value does not change much with the load current. It is possible to get an estimate
on the magnitude of the voltage-drop over the brushes from some of the following
measurements.

3.4.2 Field and armature resistance

One may say that the measurement of the field and armature resistance is
easy, and may simply be done by using an ohm-metre. This, however, is not the case,
due to a number of reasons. Firstly, one must be able to separate the individual
windings. Secondly, some of the resistence, e.g. the armature resistance is small and
therefore difficult to measure correctly, especially for large DC-motors. Thirdly, the

-54-
March 1, 2001,
brushes short-circuit a different number of the armature windings depending on the
position on the commutator. The best way for measuring the armature resistance
should be by sending a current through the winding and measuring the voltage. The
armature should be rotating in order to obtain a mean value of the resistance. This
is usually done by connecting the DC-motor to another DC-motor. For permanent
magnet DC-motors the problem of the induced EMF due to the field from the magnets
exists. This also exists in all other DC-motors to a lesser degree, because of a
remnant field form the field winding. When the field current is removed, the
magnetic material will relax to a state usually different from the origin. This state
is known as the remnant flux density, and is shown in Figure 57 .

Figure 57 Remnant flux density.


The test for finding the armature current must therefore be done first by rotating the
motor by another motor at a specific angular speed and measure the generated
voltage. This voltage is then recorded. Then is a voltage supply connected to the
armature and a set of more than three measurements at different currents is made.
The measured current versus measured voltage is then plotted as shown in Figure
58 .

-55-
March 1, 2001,
Figure 58 Relationship between measured
voltage and current.
The resistance is then found as the slope of the curve in Figure 58 . As is seen an
estimate of the voltage drop of the brushes is also found from the measurement. Note
that the sign of the voltage induced by the remnant flux density must be the same
as the supply voltage. If not, the voltage will be subtracted from the supply voltage.
The resistance of the field winding is found the same way, only this time it is not
necessary to rotate the armature and there will be no induced voltage from the
remnant flux density and there are no brushes.
It is not always desirable to rotate the armature when measuring the
armature resistance. Especially for permanent magnet DC-motors the EMF from the
magnets may obscure the measurement. In this case the armature is mechanically
blocked and a measurement is done as above. Because the brushes short-circuit a
different number of sections on the commutator, the measurement should be done at
several different positions and the armature resistance will be the average value.
Measuring the field resistance is done the same way as for the armature,
however, because there are no commutator and brushes, this is easy.

3.4.3 Field and armature inductance

With a good measurement of the resistance of the armature winding and field
winding, is the easiest way of finding the inductances simply to measure the time
constants of the windings. This is done by measuring the response to a step or
inverse-step. Properly the easiest step to use is the inverse step as shown in Figure
59 .

-56-
March 1, 2001,
Figure 59 Measurement of the time constant of the
armature using an inverse step.
At the time t1, the switch is closed and the armature is short-circuited and the
current is recorded. The time constant is found from the measurement and the
inductance is found from
La
Ja ' (69)
Ra
The resistor R1 is used to limit the current after the short-circuit. It is, however, not
practical to use a switch as shown in Figure 59 . The best way is to use a diode
connected over the terminals of the winding. When the voltage is cut, the diode starts
to conduct and the winding is short-circuited. Depending on the resistance of the
winding, it may be necessary to include the on-resistance of the diode in the
calculations of the inductance. If possible, the field winding should be open, because
otherwise there may be induced a current in it.

3.4.4 The field constant kM

Finding the field constant kM is simply to drive the motor by another motor,
establish the field in the motor, measure the angular speed and measure the voltage
generated in the armature. The field constant is then found as
e if
kM @ Mf i f ' (70)
T
It is usually not necessary to know the flux from the field winding. The term kM @ Mf is
sometimes also referred to as the field constant. The measurement of the field
constant is especially easy for the permanent magnet DC-motor, because the field is
always the same. For motors with field windings, the field windings must be supplied
with a current and the field constant then becomes a function of the current in the
field winding as shown in Figure 41 . The non-linear relationship between the field
current and the field constant is due to the non-linearity of the used materials. It
should be noted that because the compound motor has both a series and shunt
winding. Saturation will therefore cause a problem, because the current in the series

-57-
March 1, 2001,
winding will influence the flux-level in the shunt winding. It will also be important
weather the motor is cumulatively or differentially coupled, they must either be
added or subtracted.

3.5 Simulating the DC-motor


What remains is to show through a simulation how a voltage supplied to the
DC-motor results in a torque on the shaft. It is known from chapter 1 how a driving
torque result in rotation of the shaft.
dTr,mec
Te ' J @ % B @ Tr,mec % T c % Tstic % T l (71)
dt
It is also known from chapter 3 how the current in the armature results in a torque
on the shaft
Te ' kM @ Mf @ I a (72)

and how the rotation of the armature results in an induces voltage


ea ' kM @ Mf @ T (73)

and that the voltage is introduced in the armature circuit as shown in Figure 60 .

Figure 60 The armature with the


induced EMF.
Using the circuit shown in Figure 60 , the voltage equation becomes
di t
Va ' Ra @ ia t % La @ a % e a
dt
\
(74)
t
1
ia t ' @ Va & e a & Ra @ i a t dt
La
0

from the calculated current the driving torque is found from (72) and the

-58-
March 1, 2001,
dT
Te ' J @ % B @ T % T c % Tstic % T l
dt
t (75)
T ' 1
J
@ T e & Tc & T stic & Tl & B @ T dt
0
Putting this together this gives for a permanent magnet DC-motor the block-diagram
shown in Figure 61 .

Figure 61 Block-diagram for a permanent magnet DC-motor.


For DC-motors with a field winding there will be a section of the block-diagram
dedicated to finding the field current. What distingue the different DC-motors is how
the field and armature winding is connected in the block-diagram and how the field
constant is implemented.

3.6 The transfer function of a DC-motor


Above was shown the block diagram of the DC-motor. From this is it possible
to analyse the transfer function of the whole system. The general form of a block
diagram for a system is shown in Figure 62 .

Figure 62 General form of a block diagram


for a system.
The transfer function for such a system is
Cs Gs
' (76)
Rs 1%Gs @Hs

Using this on the system shown in Figure 61 and neglecting the load and the dry-
friction, the transfer function of a Permanent magnet DC-motor becomes

-59-
March 1, 2001,
kM
Cs Ra % La @ s @ B % J @ s
'
Rs kM
2
1%
Ra % La @ s @ B % J @ s
kM
'
2
Ra % La @ s @ B % J @ s % kM
kM
'
2
L a @ J @ s 2 % Ra @ J % B @ L a @ s % Ra @ B % kM
1
@ kM (77)
La @ J
'
2
Ra
B Ra @ B kM
s %
2
% @s% %
La J La @ J La @ J
1
@ kM
La @ J
'
2
1 1 1 kM
s %
2
% @s% %
Je J m Je @ Jm La @ J
2
Tn
' K@
2
s 2 % 2 @ . @ Tn @ s % Tn
As is seen from (77), the undamped natural frequency, Tn, and the damping ratio, H,
aside from being dependent on the electrical and mechanical parameters, also are
dependent on the field constant. This may cause a problem when designing a
regulator, because the system changes with the value of kM. When designing a
regulator it is therefore necessary to check the results for the span of kM. It should
also be noted that the amplification of the system changes with the value of kM. A
thing that many times help in the regulator design is that the electrical and the
mechanical time constants usually are orders of magnitude different. It is therefore
easy to regard the system as two 1'st order systems and then design one regulator to
control the torque through the armature current and the speed through the armature
voltage.

-60-
March 1, 2001,
This Page is Blank

-61-
March 1, 2001,
Chapter 4
Introduction to the induction motor
4.1 The three phase induction motor
The induction motor has been and still is the big work-horse for the industry
and in private homes. In private homes, however, it has always had a serious
competitor in the DC-motor or the universal-motor. The death-sentence for the
induction motor has been written several times, because more modern motor types
have arrived as competitors. The induction motor is, however, still alive and well, and
assumably will be for many more years.
As an overview the following may be said on the induction motor:
Pro Con
Robust construction Load-dependent slip
Can run directly from the power grid Nm/m^3 is small
Old well-known motor-technology with Positions-sensor may be required
many variations
Power electronics is capable of Difficult model
increasing performance significantly
Brush-less

4.1.1 Evolving the induction motor

The DC-motor may be used as a starting point to describe the induction motor.
In the DC-motor the armature carries a current and as it turns, the commutator
changes the direction of the current. The magnetic field is created by a set of field
windings. It was shown in the previous chapter how this results in a driving torque
acting on the armature. Picturing that the armature winding is not connected to the
commutator, but instead to as many voltage sources as there are bars on the
commutator. The voltage sources produce a square shaped voltage and each of them
have a phase-shift equal to the position of the winding on the commutator. This
should produce exactly the same currents in the armature as with the commutator.
It will be difficult to connect all these voltages, because the armature rotates.
However, the currents and torque will be exactly the same if the motor is turned

-62-
March 1, 2001,
inside out, with the armature on the outside and the field winding inside. This gives
a situation as shown in Figure 63 .

Figure 63 Inside out DC-motor without a Figure 64 The resulting rotating magnetic
commutator. field.
Now is the time to simplify. All the square-shaped voltage produces together a flux
with an orientation in space. This flux may be equivalented by a set of magnetic poles
rotating with the angular speed T, as shown in Figure 64 . Approximately the same
flux may be achieved by using a three-phase winding, where the windings are
displaced in space by 120E. This winding is known as the stator winding and is shown
in Figure 65 .

Figure 65 The connection of the individual phases in the stator of


a two-pole induction motor.
The three-phase winding then produces a rotating flux-vector which acts as the

-63-
March 1, 2001,
rotating poles in Figure 64 . This is shown in Figure 66 .

Figure 66 Resulting rotating flux from


a three-phase winding.
The angular speed of the flux is known as the synchronous speed. The field winding
in Figure 63 must have a current flowing through it in order to produce a flux. Two
options exist. The first option is simply to leave the winding as it is, and feed a DC-
voltage to it through by a set of slip-rings located on the shaft. The resulting motor
is the synchronous motor and will be described in a later chapter. The second option
is simply to short-circuit the winding. The three-phase stator winding will induce a
current in the field winding, which then produce the required magnetic field. The
rotating winding becomes the rotor winging and the whole rotating part becomes the
rotor. A thing should be noted. The only way there may be induced an
electromagnetic force is, according to Faradays law, if the flux changes. This change
is either from the flux changing in amplitude or if the winding is moved through the
magnetic field. If the rotor winding rotates with the same angular speed as the flux
from the stator, there is no change and therefore there are no induced
electromagnetic force and no current and therefore no torque. If, however, there is a
small difference in the angular speed of the rotor winding and the flux from the
stator, there will are induced electromagnetic force and current and therefore is
torque. It is therefore necessary for the rotor winding to rotate with a angular speed
different from the angular speed of the flux from stator winding. This is known as the
slip and is usually described by the difference between the rotors angular speed
compared to the synchronous angular speed. As an example, the synchronous angular
speed of a two-pole motor supplied by 50 Hz is Tsyn ' 2 @ B @ f ' 314.15 rad
s
, if the slip is
s ' 0.01 , then the angular speed of the rotor is Tr ' 1 & s @ Tsyn ' 311.02 rad s
. What
remains is to do some modifications to the rotor. In Figure 64 was the rotor shown
with four distinct poles. However, this is usually done differently. Due to the slip of
the rotor, the flux from the stator moves with regard to the stator. If the rotor has

-64-
March 1, 2001,
distinct poles, the induced current in the rotor will change because of the change in
the air-gap due to the poles. There will therefore be a oscillation of the torque due
small changes in the rotor currents. To avoid this, the rotor is usually constructed
with a phase-number much higher than the stator-winding. The way this is done, is
by constructing the rotor as a squirrel-cage rotor as shown in Figure 67 .

Figure 67 The blue bars are the squirrel cage


and the red rings are the end-rings short-
circuiting the squirrel cage.
The bars in the squirrel cage then carries a current induced by the stator winding
and the currents together form a flux in the rotor.

4.1.2 Construction of the induction motor

The construction of the induction motor is very simple. It consists of a stator


with the stator-winding, the rotor, with a short-circuited winding (the squirrel-cage),
a fan for cooling, a connection-box on top, two bearings for support and the shaft for
connection to the load. A cut-through of a typical induction motor is shown in Figure
68 .

-65-
March 1, 2001,
Figure 68 A cut-through of a squirrel-cage induction motor.
The stator is constructed from a stacked of lamination punched into the
designed shape. One sheet is shown in Figure 69 .

Figure 69 Figure 70 On the right-hand side is shown


Stator lamination. the stacked stator with slot-insulation and on
the left-hand side is shown the stator-
winding inserted.
The stator is then made by stacking the punched sheets to the designed height and
then the stator is welded or screwed together. The stator winding then is inserted
into the stator and the windings are fixed in place. This is shown in Figure 70 . In
order to distribute the magneto-motive force over a larger section of the air-gap the
stator-winding and because the flux-density in the air-gap is desirably sinusoidally
distributed and the stator-coils for each of the phases are placed into several stator-
slots. The connection of the individual turns of the coils is shown in Figure 71 and

-66-
March 1, 2001,
each of the coils is usually connected as a series connection.

Figure 71 The connection of the individual


phases in the stator of a two-pole induction
motor.
The rotor is produced in much the same way as with the stator. It is made from
laminations punches into the designed shape as shown in Figure 72 .

Figure 72 Figure 73 A finished rotor, with a cast aluminium


Rotor lamination. winding.
The rotor is produced by stacking the laminations and the rotor-winding are made
by injecting molten aluminium into the rotor-slots. The rotor-winding are produced
this for a number of reasons. Firstly, the rotor is usually made with a skew of the
rotor-slots. This is done, because the magneto-motive force produced in steps, due to
the stator-slots. By skewing the rotor, the pulsating torque produced by the steps is
reduced. A rotor-winding made from cupper-wire is difficult to skew (not impossible),
but it is made possible by the molten aluminium. Secondly, the rotor has to be fixed.
The aluminium is therefore also used to bind the rotor together. In larger motors, the
rotor-winding is not made from molten aluminium, but is made from bars of cupper

-67-
March 1, 2001,
or aluminium placed into the rotor-slots, with a short-circuited ring welded on the
ends of the rotor. Sometimes it is desirable to be able to introduce a voltage or some
extra resistance into the rotor. To do this, a three-phase wound rotor is made, with
slip-rings attached to the shaft to allow access to the winding when the rotor rotates.
A slip-ring rotor is shown in Figure 74 .

Figure 74 A wound rotor with slip-rings.


Looking at Figure 72 , it is seen that the rotor is a multi-phase winding, with a
phase for each of the rotor-slots. The multi-phase rotor-winding is un-handy when
modelling the motor, and the rotor is usually approximated by a three-phase short-
circuited winding, making it comparable to the wound-rotor in Figure 74 . From an
electrical viewpoint, this gives for a three-phase motor three stator phases and three
rotor-phases as shown in Figure 75 .

Figure 75 The stator- and rotor-winding of


an induction motor, seen from an electrical
view-point.
The three stator-windings are then connected to the three-phase system as required
and the rotor-windings are either short-circuited as with a squirrel-cage rotor or to
external voltages or resistors as with a wound rotor.

-68-
March 1, 2001,
4.2 Modelling the induction motor
The induction motor has a simple construction, however, it is not easy to
model, because many aspects have to be considered in order to get good results. There
are two approaches: Steady-state modelling and dynamic modelling. Steady-state
modelling will be precented in this chapter and dynamic modelling will be precented
in the following chapter.
Steady-state for an induction motor is a little misleading, because this would
imply that there are no changes of voltage, current, power and torque. For a three-
phase system, this is exactly what happens, the voltages and currents change as
sinus-functions. However, the RMS-values of the voltages and currents does not
change and the fundamental frequency is constant. By defining the steady-state this
way, the calculations are also valid for quasi-stationary waveforms such as PWM-
voltages.

4.2.1 Stator voltages and input power

The stator winding is supplied by a three-phase sinusoidal voltage defined by


the RMS-voltage and the frequency as
Va ' V @ sin 2 @ B f @ t ' 2 @ VRMS @ sin 2 @ B f @ t

2@B 2@B
Vb ' V @ sin 2 @ B f @ t & ' 2 @ VRMS @ sin 2 @ B f @ t &
3 3 (78)

4@B 4@B
Vc ' V @ sin 2 @ B f @ t & ' 2 @ VRMS @ sin 2 @ B f @ t &
3 3

where 2 is the difference between the peak voltage and the RMS-voltage. The
resulting voltages are as shown in Figure 76 .

Figure 76 Three-phase sinusoidal supply voltages with a RMS-


voltage of 400V.

-69-
March 1, 2001,
There exists two ways of connecting the stator winding to the three-phase supply-
voltage. They are the star-connection and the delta-connection, and are shown in
Figure 77 .

Figure 77 On the left-hand side is shown the star-


connection and on the right-hand side is shown the delta-
connection.
The voltages Vab, Vbc and Vca are the same for the two connections and the voltage
voltages VaN, VbN and VcN are found as
1 1 1
VaN ' VbN ' VcN ' @ Vab ' @ Vbc ' @ Vca (79)
3 3 3
in RMS-voltages. The same happens with the currents. In the star-connection the
current is flowing from either of the phases the same as the current flowing in the
phase windings and in the delta-connection the current is split between two phase
windings. Just as with the voltages in the star-connection is the current in the phase
windings in the delta-connection reduced by 3 . Using the star-connection the input
power in one phase winding is calculated as
)
Pin ' VaN @ Ia @ cos n
1 (80)
' @ Vab @ I a @ cos n
3

and for all the phases this becomes

-70-
March 1, 2001,
)
Pin ' 3 @ P in
1
' 3@ @ Vab @ I a @ cos n
3
(81)
' 3 @ Vab @ I a @ cos n

' 3 @ 3 @ VaN @ Ia @ cos n


' 3 @ VaN @ Ia @ cos n
this, however, is exactly the same result as for the delta-connection.

4.2.2 Magneto-motive force of an induction motor (MMF)

One of the things that makes the induction motor difficult to understand is the
fact that the three phases carry currents that are displaced by f 1 3 seconds in time
@

and the phase windings are displaced by 2 3 B radians in space. Each of the three
@

phase-windings have a sinusoidal current flowing through them and each of these
current causes a magneto-motive force to form. Considering the currents as
ia ' I m @ cos 2 @ B f @ t

2@B
ib ' I m @ cos 2 @ B f @ t &
3 (82)
4@B
ic ' I m @ cos 2 @ B f @ t &
3
The resulting MMF therefore are a result of the spacial orientation of the individual
phase-windings and the current flowing through them. This is known as the space-
6
vector for the MMF and is denoted by res and is found by
a ' N @ i a cos 2

2@B
b ' N @ i b cos 2 &
3
4@B
c ' N @ i c cos 2 & (83)
3
6 6 6 6
res ' a % b % c

2@B 4@B
' N @ i a cos 2 % N @ i b cos 2 & % N @ i c cos 2 &
3 3

-71-
March 1, 2001,
and the currents are known from (82), so the MMF becomes
6
res ' N @ Im @ cos 2 @ B f @ t @ cos 2

2@B 2@B
% N @ Im @ cos 2 @ B f @ t & @ cos 2 & (84)
3 3
4@B 4@B
% N @ Im @ cos 2 @ B f @ t & @ cos 2 &
3 3

using the trigonometric identity


1 1
cos A @ cos B ' @ cos A & B % @ cos A % B (85)
2 2

the MMF becomes


6 1 1
res ' @ N @ Im @ cos 2 @ B @ f @ t & 2 % @ N @ Im @ cos 2 @ B @ f @ t % 2
2 2
1 1 4@B
% @ N @ Im @ cos 2 @ B @ f @ t & 2 % @ N @ Im @ cos 2 @ B @ f @ t % 2 &
2 2 3
1 1 8@B
% @ N @ Im @ cos 2 @ B @ f @ t & 2 % @ N @ Im @ cos 2 @ B @ f @ t % 2 &
2 2 3
(86)
1 1
' @ N @ Im @ cos 2 @ B @ f @ t & 2 % @ N @ Im @ cos 2 @ B @ f @ t % 2
2 2
1 1 4@B
% @ N @ Im @ cos 2 @ B @ f @ t & 2 % @ N @ Im @ cos 2 @ B @ f @ t % 2 &
2 2 3
1 1 2@B
% @ N @ Im @ cos 2 @ B @ f @ t & 2 % @ N @ Im @ cos 2 @ B @ f @ t % 2 &
2 2 3
Fortunately all the right-hand terms in (86) cancel each other out and the MMF
becomes
6 3 3
res ' @ N @ I m @ cos 2 @ B @ f @ t & 2 ' @ N @ I m @ cos T @ t & 2 (87)
2 2
It should be noted that the right-hand terms in (86) are backward rotating and will
be precent if the currents for some reason become un-balanced. In Figure 78 is
shown the rotation of the space-vector for the MMF.

-72-
6
res

March 1, 2001,
6 6
6 b c
c 6 6
a res

6
b

6
6 b
res 6 6
6 c b
a 6
c

6
res

Figure 78 Rotation of the MMF due


to the change in the currents.
4.2.3 Induced voltages of an induction motor (EMF)

As with the DC-motor, there will be induced an EMF in the winding of an


induction motor. For a two-pole machine, if the flux is distributed sinusoidally in the
air-gap, the flux density may be described as
B 2 ' Bmax @ cos 2 (88)

and the flux per pole then becomes


B
Mp ' 2
B 2 @ l @ r d2 ' Bmax @ l @ r (89)
&B
2

where l is the axial length of the motor and r is the radius of the stator in the air-gap.
Considering a full-pitch coil, where the coil spans 180 electrical degrees, the flux
linkage becomes
8a ' N @ Mp @ cos T @ t (90)

and the induced EMF in phase a, becomes


d8
ea ' & a ' T @ N @ Mp @ sin T @ t
dt (91)
' Emax @ sin T @ t

and for the remaining phases

-73-
March 1, 2001,
2@B
eb ' Emax @ sin T @ t &
3
(92)
4@B
ec ' Emax @ sin T @ t &
3

and as RMS-voltage it becomes


T @ N @ Mp
Erms '
2
2@B@f (93)
' @ N @ Mp
2
' 4.44 @ f @ N @ Mp

4.2.4 Winding factors

So far, all the coils of a phase have been put into the same stator slot and the
pitch of the winding were 180 electrical degrees. However, coils spanning 180
electrical degrees is not the optimal design, because the section of the coil going from
one slot to the next, the end-winding, is long and therefore increases the resistance
of the winding without contributing to the development of torque. The end-winding
also increase the iron losses of the iron core. The MMF of the stator-winding in an
ideal motor should be sinusoidally distributed, but with the windings spanning 180
electrical degrees this is not so. The way of making the winding shorter and at the
same time making the MMF more sinusoidal is by not letting the winding span 180
electrical degrees and by distributing the winding over several slots. This will cause
the addition of winding factors in (93). There exist three winding factors: A
distribution factor, pitch factor and skew factor. When the coils of the winding are
distributed over several slots, this will cause a lowering of the combined induced
EMF. The individual coils will have a different spacial orientation due to the slots
and there will be a phase difference between them. In Figure 79 is shown an
example where the phase windings are distributed over two slots.

-74-
March 1, 2001,
Figure 79 An example of a distributed
winding of a two-pole induction motor.
To calculate the EMF in this case it is necessary to calculate the difference between
the arithmetic and the geometric sum of the voltages from each of the slots. In
Figure 80 is shown a geometric sum of voltages of different slots.

"

"
2 "

n@"
n@"
2

Figure 80 Coil voltages in a distributed


winding.
The distribution factor is defined as

-75-
March 1, 2001,
geometric sum of voltages
Kd ' (94)
arithmetic sum of voltages

Defining " as the angle between two adjacent slots and n as the number of slots per
pole per phase or the number of slots in a phase belt. From Figure 80 and (94) the
distribution factor becomes
RU 2 @ Rx Rx
Kd ' ' '
n @ RS n @ 2 @ Ry n @ Ry
n@" n@"
OR @ sin sin (95)
2 2
' '
" "
n @ OR @ sin n @ sin
2 2
The next thing to consider is the shortening of the cols. In Figure 81 is shown an
example of a winding that is both distributed and short pitched.

Figure 81 An example of a distributed and


short pitch winding of a two-pole induction
motor.
The pitch factor is defined as
voltage induced in a short&pitched coil
Kp ' (96)
voltage induced in a full&pitched coil

-76-
March 1, 2001,
In Figure 82 is shown the voltages of a full-pitch and a short-pitch coil.

Figure 82 on the left-hand side is shown a full-pitch coil and a


short-pitch coil and on the right-hand side is shown the voltages of
the same coils.
Using (96) and Figure 82 the pitch-factor is found as
(
2 @ e @ cos
2
Kp '
2@e (97)
(
' cos
2

Finally, sometimes the stator slots are skewed in order to minimize the pulsating
torques from the interaction between the stator and rotor slots. By calculating how
much the area of the coil is reduced by this, a skew factor is found, Ks.
Using the winding factors, the induced EMF is found as
Erms ' 4.44 @ f @ N @ Mp @ K w
(98)
Kw ' K d @ Kp @ K s

4.2.5 Single-phase star equivalent circuit

In (81) was it shown that the power of an induction motor may be calculated
from the voltage and current in a single phase of the motor (in steady-state only).
This may be used to do calculations on the motor in steady state. This is known as
the single-phase star equivalent circuit. It is constructed by three parts: The stator,
the rotor and the iron-core linking the stator and rotor. The stator and rotor is shown
in Figure 83 .

-77-
March 1, 2001,
Figure 83 Three-phase stator and rotor of
the induction motor.
Looking at one phase of the stator, it consists of the winding resistance and leakage
inductance as shown in Figure 84 .

Figure 84 Stator in single-phase equivalent


circuit.
Exactly the same happens with the rotor as shown in Figure 85 .

Figure 85 Rotor in single-phase equivalent


circuit.
There is one difference between the stator circuit and the rotor circuit. Firstly,
considering the rotor rotating at synchronous speed. As shown above in this case
there will be no current in the rotor. Secondly, if the rotor speed is zero, there will be
induced the maximum current. This is modelled by making the rotor resistance slip-
dependent. The iron-core linking the stator and rotor is represented by an inductor
modelling the magnetization of the corre parallel with a resistor modelling the iron
losses in the core as shown in Figure 86 .

-78-
March 1, 2001,
Figure 86 The core in single-phase
equivalent circuit
What remains is to put everything together. Looking at the stator-rotor they act just
as a transformer, with the ratio a. this is shown in Figure 87 .

Figure 87 Single-phase stator-rotor equivalent circuit.


It is often used to eliminate the ratio between the stator and rotor circuit, and simply
denoting the rotor circuit with a , and thereby indicating that the ratio is included
in the value of the current, resistor and inductor. The circuit in Figure 87 is further
modified because the developed power is
P mech ' Te @ T (99)

therefore, at synchronous speed there is no mechanical power, because there is no


current in the rotor circuit and as standstill there is no mechanical power because T
is zero. The rotor resistance is therefore divided in two separate resistors, one for the
rotor resistive loss and one for the power transferred to the shaft of the motor. This
is shown in Figure 88 .

-79-
March 1, 2001,
Figure 88 Single-phase stator-rotor equivalent circuit with a
separate resistor for the shaft-power.
To calculate the developed power and torque, it is a matter of calculating the power
dissipated in the slip-dependent rotor resistor. Using Rs'0.14S , Rr '0.20S ,
)

Xs'0.45S , Xr '0.40S , Rfe'300S , Xc'13S , U'230V and fn'50Hz the power and torque
)

may be calculated as a function of the slip. The results are shown in Figure 89 .

Figure 89 On the top right and left is shown the calculated stator
and rotor currents, on the bottom left is shown the shaft power and
on the right is shown the shaft torque.
Two important features of the induction motor may be seen from Figure 89 . Firstly,
at the synchronous speed of the rotor, there cannot be developed any shaft power or
shaft torque, because there runs no current in the rotor. Secondly, at standstill,
where the slip is s'1 , there cannot be developed any shaft power, because the rotor
is not rotating, however, there will be a shaft torque, because there runs a current
in the rotor. A second thing the steady-state calculation may be used for, is to select

-80-
March 1, 2001,
the motor from a known load. The first criteria is that the motor must be able to start
from stand-still. In Figure 90 is shown some different loads together with the
developed torque. Looking at the cyan curve, the developed torque is sufficient to
start the motor, however, the difference between the two torques is small, causing the
startup to be slow and a small increase in the load torque may cause the motor to
stall. The steady-state operation point will be the intersection of the blue and cyan
curves. The red graph in Figure 90 is better, because there is a large difference
between the red and blue curves. The steady-state operation point will be the
intersection between the blue and red curves. However, this operation point is not
very well suited to the motor, because looking at the curve for the efficiency, it is very
poor for this point.

Figure 90 On the top shown, in blue, the developed torque and the
other colours are different loads. On the bottom is shown the
calculated efficiency.
The best selection for a motor is to select is so that the steady state operation point
is close to the point of maximum efficiency, and at the same time ensure that the
motor has sufficient torque to quickly accelerate the load.
Sometimes it will be desirable to temporarily be able to increase the developed
starting torque. This may be done by introducing more resistance into the rotor-
circuit. There are two ways of doing this: By using a wound-type rotor with external
resistors connected via slip-rings on the shaft or by special design of the rotor-bars
in the squirrel-cage. In Figure 91 is shown the developed shaft torque of an
induction motor with external resistors connected via slip-rings. As is seen, the

-81-
March 1, 2001,
maximum torque is the same, but the starting torque is significantly increased.
Using extra resistance in the rotor, however, reduces the efficiency of the motor.

Figure 91 Torque curves of an induction motor with slip-


rings by increasing rotor resistance.
The second option for increasing the starting torque is use the phenomenon of
skin-depth. At stand-still, the frequency in the rotor is equal to the frequency of the
stator and when running at the steady-state operation point, the frequency is much
lower. In Figure 92 is shown a deep rotor-bar. At stand-still, where the frequency
in the rotor is high, the changing magnetic flux causes the currents in the rotor-bars
to be pushed outwards and thereby reducing the effective conducting area and
therefore increasing the resistance.

Figure 92 Deep rotor-bar.


The same way as with the wound rotor and the slip-rings, the starting torque
becomes larger and as the speed of the rotor increases, the resistance drops, giving

-82-
March 1, 2001,
a good efficiency at steady-state.

This Page is Blank

-83-
March 1, 2001,
Chapter 5
Dynamic modelling of the induction motor
The steady-state modelling of the induction motor is good for predicting things
as load and torque and to some extend losses and efficiency. However, in many cases
are induction motors involved in very dynamic loads, and the steady-state model of
the motor cannot model the dynamic behaviour. For this reason a dynamic model for
the induction motor must be known.

5.1 Phase transformations


At this point it is known that the three phases together produce a rotating
space vector. The windings in the motor may be described by their resistance, leakage
inductance and magnetizing inductance. Together they describe the electromagnetic
system by a set of 6x6 matrices of differential equations as
d8
v ' R@i % , 8 ' L@i
dt
(100)
di dL
' R@i % L@ % @i
dt dt
where the voltages and currents are vectors as
v sa i sa
v sb i sb
v sc i sc
v ' ) , i ' ) (101)
v ra i ra
) )
v rb i rb
) )
v rc i rc

and where the resistance and inductance is known as

-84-
March 1, 2001,
rs 0 0 0 0 0 L sasa Lsasb L sasc Lsara L sarb Lsarc
0 rs 0 0 0 0 L sbsa Lsbsb L sbsc Lsbra L sbrb Lsbrc
0 0 rs 0 0 0 L scsa Lscsb L scsc Lscra L scrb Lscrc
R ' , L ' (102)
0 0 0 rr 0 0 L rasa Lrasb L rasc Lrara L rarb Lrarc
0 0 0 0 rr 0 L rbsa Lrbsb L rbsc Lrbra L rbrb Lrbrc
0 0 0 0 0 rr L rcsa Lrcsb L rcsc Lrcra L rcrb Lrcrc

An electrical motor is usually modelled as if it were a two-pole motor. This is because


the equations describing the currents and fluxes are independent of the pole-number.
The number of poles only becomes important when calculating the electromagnetic
torque and the angular speed of the shaft. All the voltages and currents in the rotor
are converted to the stator side by the ratio between the stator and rotor windings
as with a transformer. This is usually indicated by a ' as shown in (101). In the
following this is omitted due to several other indexes and thereby enhancing the
clarity. All rotor values are in the following converted to the stator-side even though
the ' is absent. The system in (100), (101) and (102) is a well-defined system, with
a substantial problem: The coupling between stator and rotor is dependent on the
rotor-position. To clear up the system, it is usually divided into a stator-part and a
rotor-part, together with the mutual coupling as.
abc abc abc abc
8s Lss Lsr is
' @ (103)
abc abc abc abc
8r Lrs Lrr ir

The sub-matrices then become


cos 2r cos 2r % 2B cos 2r % 4B
L ls % L ss L sm L sm 3 3
abc
ss ' L sm L ls % L ss L sm abc
, Lsr ' L sr @ cos 2r % 4B cos 2r cos 2r % 2B
3 3
L sm L sm L ls % L ss
cos 2r % 2B cos 2r % 4B cos 2r
3 3
(104)
cos 2r cos 2r % 4B cos 2r % 2B
3 3 L lr % L rr L rm L rm
2B 4B L lr % L rr
abc
rs ' L sr @ cos 2r % cos 2r cos 2r % abc
, Lrr ' L rm L rm
3 3
L rm L rm L lr % L rr
cos 2r % 4B cos 2r % 2B cos 2r
3 3

In numerical terms, there is no problem in solving the system shown in (101), (103)
and (104), but it is unclear what exactly must be done to get a specific result. A goal

-85-
March 1, 2001,
could be if it were possible to get a system comparable to a DC-motor. A point where
a measure of fingerspitzgefl is necessary is how, at a given instance, to calculate
the coupling between the stator and rotor, due to the fact they rotate in regard to
each other.
The problem of the rotating rotor-winding has been known for many years, and
many more or less fanciful suggestion to the solution has been precented. The
question then becomes:Does anything work then?. Many methods work, but as it
is written in the description of EMTDC: It is a good tool in the hands of a good
artisan., and just because it works, it is not necessarily a good idea! A method which
is a good idea is Parks transformation (or two-phase transformation, or to-axis
transformation, or dq-axis transformation, many names for a good tool).
What is the method then? Starting with balanced voltages and linear
inductances a rotating space-vector is produced as shown in Figure 93 .
2 @ Im

t ' 0 t ' t ' 2


3 3

! t
! t s

!t
s

t ' 5 t ' 4 t '


3 3

! t
s

! t
s ! t
s

Figure 93 On the top is shown the balanced three-phase


current and the resulting space-vector is shown on the
bottom.
The rotating space-vector form Figure 93 draws a figure in space. In steady-state
and with linear inductances, this figure becomes a circle. The circle is interesting,
because it has some interesting properties. If the goal is to create a rotating space-
vector describing a circle, three phases with sinusoidal currents are not necessary.
The interesting question becomes: Is it possible to model the three-phase system

-86-
March 1, 2001,
with a different number of phases, possibly with different voltages and currents?.

5.1.1 One rotating coil

If the only requirement were to produce the circular rotating space-vector, this
may be done by a single coil rotating at the same speed as the rotating flux, and
where the current flowing through it is a DC-current as shown in Figure 94 .

Figure 94 Equivalent rotating coil


producing the space-vector.
The problem is, that with only one coil, it is only possible to model steady-state,
because the cross-couplings between the individual coils are neglected. Even quasi-
stationary voltages as PWM-voltages may become erroneous. It is clear from this,
that a minimum of two coil-sets is required to be able to model the system. In the
following there will be described the stationary frame of reference and the rotating
frame of reference .

5.1.2 Stationary frame of reference

From analytic geometry it is known that the circle may be described by two
coordinates in space (x and y). This may be used in this case, by placing two coils at
90E and by supplying them with sinusoidal current displaced by p 90E (or B/2).
These two coils are usually named the "-coil and the $-coil. The reference for the "-
coil, called the "-axis is usually placed together with phase A. This is shown in
Figure 95 .

-87-
March 1, 2001,
Figure 95 Stationary frame of
reference.
This method reduces the three-phase system to a two-phase system.

5.1.3 Rotating frame of reference

By combining the single rotating coil and the stationary frame of reference, a
set of two rotating coils is produced as shown in Figure 96 .

Figure 96 Rotating frame of reference.


Doing this, it is possible to model the cross-couplings between the individual coils. A
further advantage is that in steady-state, the currents flowing in the coils are a DC-
current. Using the rotating frame of reference, the differentials of any state value
(d/dt) are zero in steady-state and when the differentials are different from zero, they
give the change from steady-state only. The rotating frame of reference also has the
advantage that the rotor-angle is known (it is a state). In the rotating frame of

-88-
March 1, 2001,
reference the frame of reference in regard to the phase A is named the d-axis (for
direct axis) and the other axis is named the q-axis (for quadrature axis).

5.1.4 A bit of anarchy

At this point the fundamentals of phase transformations are established, but


aside for the DC-currents in the rotating frame of reference, none of the other
advantages have been shown. Unfortunately it is necessary to note a thing: Be
careful when reading books on motor-modelling using phase transformations!! When
the method was shown in Figure 95 and Figure 96 , the A-phase gave the reference
to the "-axis and the d-axis. The problem is that this is different in some text-books.
Two conventions exist. On Figure 97 is the two conventions shown.

Figure 97 The two axis conventions.


At a time, the convention shown on the left-hand side of Figure 97 was used in
Europe and the one shown on the right-hand side was used in the USA. This,
however, is not the case today, and it is necessary to check the method used in the
individual book. The change in convention causes a change in sign in some of the
equations, and may therefore make it difficult to compare methods used in different
books. In this text there will be used the convention shown on the left-hand side of
Figure 97 .

5.1.5 Transformation of currents, voltages, flux-linkage, etc.

What remains is to define a method for performing the phase transformations


to the stationary or rotating frame of reference. From the definition there is no reason
to regard the two transformations as two different transformations. This is because
in the stationary frame of reference the angle is constant (and zero) and for the
rotating frame of reference the angle is changing in time. The description in the
following will therefore be for the dq-system. For the "$-system the transformation
are similar, only the angle will always be zero.
The transformation is done by defining a transformation-matrix for the
systems as
fdq0 ' Tdq0 2 @ f abc (105)

-89-
March 1, 2001,
where fabc is the voltage, current, flux-linkage, etc in the three-phase system that is
to be transformed to the two-phase system. In the following is the angle-index in
(105) removed, unless the angle is different from 2. The equation in (105) does not
correspond exactly to the transformation described above. The d-axis and q-axis has
been defined, but what is the 0-axis in (105)? This is because the d- and q-axis is
insufficient to describe all cases. Therefore something is borrowed from the
Symmetrical components-method. The use of symmetrical components is widespread
in calculations on high-voltage systems in steady-state. The method works by
defining a transformation-matrix as in (105) as
f120 ' T120 @ fabc

a a2 1
1
T120 ' @ a2 a 1 (106)
3
1 1 1
2B
j@
a ' e 3

and the inverse transformation as


&1
f abc ' T120 @ f120

a2 a 1
(107)
&1
T120 ' a a2 1
1 1 1

The transformation results in a vector containing three components (1,2,0). In


Figure 98 is shown how symmetrical components may be used to describe any three-
phase un-balanced voltages in steady-state.

-90-
March 1, 2001,
Va1
Vb2
Va2 Va

Vc2

Vc1
Vb1

Va0 0 Vb
Vb
Vc0

Vc

Figure 98 Symmetrical components.


Un-balanced currents or flux-linkages may be caused an imbalance in the three-
phase winding or by the operating condition of the machine. This is the reason why
a current in the neutral wire is possible in a balanced voltage system. In (105) is the
zero-sequence added due to the same conditions as for the symmetrical components,
because operating conditions may cause a zero-sequence. For the dq0-system the
transformation-matrix then becomes
fdq0 ' Tdq0 @ fabc

2B 4B
cos 2d cos 2d & cos 2d &
3 3
(108)
2
Tdq0 ' @ &sin 2d &sin 2d & 2B &sin 2d &
4B
3 3 3
1 1 1
2 2 2
where 2d is the angle between the d-axis and the axis for the A-phase. The inverse
transformation becomes
&1
f abc ' Tdq0 @ f dq0

cos 2d &sin 2d 1
(109)
2B 2B
&1 cos 2d & &sin 2d & 1
Tdq0 ' 3 3
4B 4B
cos 2d & &sin 2d & 1
3 3

-91-
March 1, 2001,
For the stator frame of reference the results are similar, just with the angle replaced
by a zero and the transformation-matrix becomes.
f"$0 ' T"$0 @ f abc

1 1
1 & &
2 2
(110)
2 0 1 1
T"$0 ' @ 3 & 3
3 2 2
1 1 1
2 2 2
and the inverse transformation becomes
&1
fabc ' T"$0 @ f"$0

1 0 1
1 1 (111)
&1 & 3 1
T"$0 ' 2 2
1 1
& & 3 1
2 2
What remains is to check if the transformations act as expected. In Figure 99 is
shown a balanced three-phase supply voltage.

Figure 99 Three-phase balanced supply voltage.

-92-
March 1, 2001,
In Figure 100 are the voltages from Figure 99 transformed to a stationary frame
of reference and in Figure 101 are the voltages transformed to a rotating frame of
reference.

Figure 100 Balanced supply voltages Figure 101 Balanced supply voltages
transformed to the stationary frame of transformed to the rotating frame of reference.
reference.
The stationary frame of reference should produce two voltages displaced by 90E and
the rotating frame of reference should produce DC-voltages. This is exactly what is
shown in Figure 100 and Figure 101 . What happens when the voltages are
unbalanced? In Figure 102 is shown a supply voltage, where one phase-voltage is
raised by 20%.

Figure 102 Un-balanced three-phase supply voltages.


In Figure 103 the voltages from Figure 102 are transformed to the stationary
frame of reference and in Figure 104 are the voltages transformed to the rotating
frame of reference.

-93-
March 1, 2001,
Figure 103 The unbalanced supply voltageFigure 104 The unbalanced supply voltage
from Figure 102 , transformed to thefrom Figure 102 , transformed to the
stationary frame of reference. rotating frame of reference.
As it is seen, the result is different from the balanced case. The existence of a zero-
voltage, which is the same for the two transformations, should be noted. It should
also be noted that there exists a frequency component in all the voltages and the
frequency in the zero-sequence is different from the other frequencies.
A transformation between the two frames of reference exists as
fd cos 2 &sin 2 0 f"
fq ' sin 2 cos 2 0 @ f$ (112)
f0 0 0 1 f0
The transformation between the two systems is useful, because in a system where the
machine is connected to a large network, it is a good thing to do all the calculations
on the two-phase frame of reference. It is, however, not a good idea to do all the
calculations in e.g. the rotating frame of reference.
One important thing must be noted on the phase-transformation as shown
t &1
above. The method is not power-invariant!! This is because Tdq0 Tdq0 . This may be
shown by

-94-
P abc ' va @ i a % vb @ i b % vc @ i c

March 1, 2001,
t
va ia

' vb @ ib
vc ic
(113)
t
' vabc @ iabc
&1 t &1
' Tdq0 @ vdq0 @ Tdq0 @ i dq0
&1 t &1
' vdq0 @ Tdq0 @ Tdq0 @ idq0

and when
3
0 0
2
&1 t &1 3
Tdq0 @ Tdq0 ' 0 0 (114)
2
1
0 0
3
this means that
3 1
P abc ' @ v d @ i d % v q @ i q % @ v0 @ i 0 (115)
2 3

This is a fact of the method. It is possible to change the transformation to the "$0-
system or the dq0-system to be power-invariant. This is, however, not the norm for
doing calculations on electrical machines. For calculations on synchronous generators
coupled to extended power grids it is usually done.

5.2 Transformation of the voltage equation


At this point are known space vectors and phase-transformations. What has
yet to be described is what happens to the voltage equations when they are
transformed from the abc-system to the dq0-system. The following step involves a
number of matrix-operations. To enhance the clarity it is necessary to define

dF t .
' F t ' pF t (116)
dt

-95-
March 1, 2001,
The dot-notation for the differential is a well-known way of depicting a time-
differential. However, the dot is easily missed. In this text it is used to put a p in
front of a system parameter, if it is differentiated in time. Furthermore, the
transformation used in the following is the rotating frame of reference and not the
stationary frame of reference. This is because the rotating frame of reference is the
more general case, with the stationary frame of reference as a special case. Later the
stationary frame of reference will be derived from the rotating frame of reference.

5.2.1 Transformation of stator-voltages

The best place to start is with what is known to be right. The starting point is
the voltage equation for the stator in the abc-system.
abc
abc abc abc d8s
vs ' rs @ is %
dt (117)
abc abc abc
' rs @ is % p8s

The first step is the three-phase voltage from the resistance of the stator winding
dq0 abc abc &1 dq0
vr s ' Tdq0 @ vr s ' Tdq0 @ rs @ Tdq0 @ is (118)
&1 dq0 abc
Looking at (118) from right to left, it is obvious that Tdq0 @ is ' is , in accordance
abc
with the phase-transformations above. When r s is multiplied, the result must be
abc
the three-phase voltage from the resistance of the stator winding, vr s , which
corresponds to the first part of (117). This voltage may then be transformed to the
dq0-system by Tdq0 . The same way for the flux-linkage
dq0 abc &1 dq0
v8s ' Tdq0 @ v8s ' Tdq0 @ p Tdq0 @ 8s (119)
&1 dq0 abc
In (119) it is obvious that Tdq0 @ 8s ' 8s . The flux-linkage is differentiated and
dq0
then transformed to v8s by Tdq0 . Together (118) and (119) produces the voltage
equation for the stator in the dq0-system.
dq0 abc &1 dq0 &1 dq0
vs ' Tdq0 @ r s @ Tdq0 @ is % Tdq0 @ p Tdq0 @ 8s (120)
dq0
In (119) it is assumed that 8s is known from somewhere. It will be shown later how
dq0
8s is obtained. It is necessary to examine the result from (120) closer. Firstly the
&1 dq0
last part of (120), p Tdq0 @ 8s must be examined more closely. The flux-linkage, 8,
consists of the product of i and L, and because the inductance-matrix is dependent
on the rotor-position it is necessary to use the chain-rule while differentiating. The
differentiation then becomes
&1 dq0 &1 dq0 &1 dq0
p Tdq0 @ 8s ' pTdq0 @ 8s % Tdq0 @ p8s (121)
&1 dq0
The last part in (121), Tdq0 @ p8s is straight-forward, because it is the flux-linkage

-96-
March 1, 2001,
differentiated and the resulting voltage is transformed to the abc-system. On the
&1 dq0
other hand pTdq0 @ 8s is less straight-forward, because how may the transformation-
matrix be differentiated in time, without making it extremely complicated? The
solution is a little trick, which is to multiply by d2/d2 . ( d2/d2 ' 1 , so no overall
change happens)
&1
&1 dq0 dTdq0 dq0
pTdq0 @ 8s ' @ 8s
dt
&1
dTdq0 d2 dq0
' @ @ 8s (122)
dt d2
&1
dTdq0 d2 dq0
' @ @ 8s
d2 dt
&1
This is a good idea, because d2/dt ' T and Tdq0 is defined from 2, which makes it
possible to differentiate it.
&sin 2 cos 2 0
&1 2B 2B
dTdq0 &sin 2 & cos 2 & 0
' 3 3 (123)
d2
4B 4B
&sin 2 & cos 2 & 0
3 3
Collecting the bits, (121) becomes
&sin 2 cos 2 0
2B 2B
&1 dq0 &sin 2 & cos 2 & 0 dq0 &1 dq0
p Tdq0 @ 8s ' 3 3 @ T @ 8s % Tdq0 @ p8s (124)
4B 4B
&sin 2 & cos 2 & 0
3 3
Because this is the voltage equation for the stator, the angle is known ( 2 ' 0 ), and
the matrix in (124) is simplified to
&sin 2 cos 2 0
2B 2B 0 1 0
&sin 2 & cos 2 & 0
3 3 ' &1 0 0 (125)
4B 4B 0 0 0
&sin 2 & cos 2 & 0
3 3
Collectively this gives the stator voltage equations as

-97-
March 1, 2001,
0 1 0
dq0 dq0 dq0 dq0 &1 dq0
vs ' rs @ is % T @ &1 0 0 @ 8s % Tdq0 @ p8s (126)
0 0 0

5.2.2 Transformation of rotor voltages

In exactly the same way as with the stator voltages, the rotor voltages are
transformed to the dq0-system.
abc
abc abc abc d8r
vr ' rr @ ir %
dt (127)
abc abc abc
' rr @ ir % p8r

The procedure is exactly the same as for the stator, with the exception of the used
angle. Because the voltage is transformed to the stator-side, the angle becomes
2 & 2r , due to the rotor slip. The phase/transformation then becomes Tdq0 2 & 2r . The
resulting voltage equation then becomes.
0 1 0
dq0 dq0 dq0 dq0 &1 dq0
vr ' rr @ ir % T & Tr @ &1 0 0 @ 8r % Tdq0 @ p8r (128)
0 0 0

5.3 Transformation of three-phase resistors and flux-linkage


What remains is to show what happens to the resistors and inductances in the
transformed system, because when the voltages change, the resistors and inductances
is not necessary the same.

5.3.1 Transformation of a three-phase resistor

The three-phase resistor is the least complicated to transform. By taking the


equation from (120) the result is

-98-
&1

March 1, 2001,
dq0 abc dq0
vr s ' Tdq0 @ r s @ Tdq0 @ is

1 0 0
&1 dq0
' Tdq0 @ rs @ 0 1 0 @ Tdq0 @ is
0 0 1
1 0 0
&1 dq0
' rs @ Tdq0 @ 0 1 0 @ Tdq0 @ is
0 0 1
1 0 0
&1 dq0
' rs @ 0 1 0 @ Tdq0 @ Tdq0 @ is
(129)
0 0 1
1 0 0
dq0
' rs @ 0 1 0 @ is
0 0 1
\
abc dq0
rs ' rs

1 0 0
' rs @ 0 1 0
0 0 1
This means that it is the same resistor in both abc-system and dq0-system.

5.3.2 Transformation of flux-linkage

The way to solve this, id exactly the same procedure as above. Starting with
the abc-system
abc abc abc abc abc
8s ' Lss @ is % Lrs @ ir (130)

which transformed to the dq0-system gives


dq0 abc
8s ' Tdq0 @ 8s
abc abc abc abc
' Tdq0 @ Lss @ is % Lsr @ ir (131)
abc &1 dq0 abc &1 dq0
' Tdq0 2 @ Lss @ Tdq0 2 @ is % Tdq0 2 @ Lsr @ Tdq0 2 & 2r @ i r

-99-
The angle 2 & 2r in the last part is important due to the fact that the rotor slips. It

March 1, 2001,
is only the angle 2 that is used when transforming back to the dq0-system, because
it is the stator flux-linkage. To solve (131) extended equation-gymnastics is used,
involving trigonometric relations and matrix operations. These gymnastics are rather
comprehensive and not very interesting, however, the result is interesting.
(It is a good idea to try to solve (131), because it is a good exercise)
3
Lls % @ L
ss 0 0 3
2 @ L 0 0
ss
2
3
8dq0
s ' 0 Lls % @ L
ss 0 @
dq0
is % 3 @ i
r
dq0
(132)
2 0 @ L
ss 0
2
3
0 0 Lls % @ L
ss 0 0 0
2

In the same way is the rotor flux-linkage found. Starting in the abc-system
abc abc abc abc abc
8r ' Lrs @ is % Lrr @ ir (133)

which transformed to the dq0-system gives


dq0 abc
8r ' Tdq0 @ 8r
abc abc abc abc
' Tdq0 @ Lrs @ is % Lrr @ ir (134)
abc &1 dq0 abc &1 dq0
' Tdq0 2 & 2r @ Lrs @ Tdq0 2 @ is % Tdq0 2 & 2r @ Lrr @ Tdq0 2 & 2r @ ir

and with some equation-gymnastics this gives


3
3 Lls % @ L
ss 0 0
@ L 0 0 2
ss
2
3
8dq0
r ' 3 @ i
s
dq0
% 0 Lls % @ L
ss 0 @
dq0
ir (135)
0 @ L
ss 0 2
2
3
0 0 0 0 0 Lls % @ L
ss
2

Looking at (132) and (135), some very interesting things has happened:
There are no angles in the matrices.
The system has become decoupled. (There are many zeros outside the
diagonal)
The zero-sequence in the stator and rotor are entirely independent on
each other. (Which does not mean that there is no zero-sequence in stator
and rotor)

-100-
March 1, 2001,
The flux-linkage in the dq0-system in a compact form then becomes.
8dq0 ' Ldq0 @ idq0
\
8 Lls % L m 0 0 Lm 0 0 i
ds ds

8qs 0 Lls % L m 0 0 Lm 0 i qs
80s 0 0 Lls 0 0 0 i0s (136)
' ) @
8dr Lm 0 0 L lr % Lm 0 0 i rd
8qr 0 Lm 0 0
)
L lr % Lm 0 i rq
80r 0 0 0 0 0 L lr i0r
)

The electromagnetic part of the induction motor may then be solved. What remains
is to find the developed electromagnetic torque.

5.4 Equivalent circuits


The result from the above is a set of equations describing the electromagnetic
system in the rotating frame of reference. The equations describing the system may
be interpreted as equivalent circuits, which may help in understanding the dynamics
of the system.

5.4.1 Equivalent circuit, rotating frame of reference

Looking at equation (126) and (128), it is seen that three components exists:
one associated with the resistance of the wiring, one associated with the exchange of
magnetic energy and one associated with the rotation of the rotor (the electromotive
force). Using this knowledge, it is possible to construct an equivalent-diagram of the
d-, q- and 0-axis individually, remembering that the exchange of magnetic energy
happens through the leakage inductances and magnetizing inductance. For the
rotating frame of reference the resulting equivalent diagram for each of the axis is
shown in Figure 105 .

-101-
March 1, 2001,
Figure 105 The equivalent diagram in the rotating frame
of reference.
Looking at the model as it is shown in Figure 105 , it is clear that the d-, q-and 0-
axis are decoupled from each other.

5.4.2 Equivalent circuit, stationary frame of reference

The stationary frame of reference is a special case of the rotating frame of


reference, because the stator windings are stationary. This causes the electromotive
forces in the stator-winding in Figure 105 to become zero (it is not moving), and the
electromotive force in the rotor becomes proportional to the rotors angular speed and
not the difference between the rotors angular speed and the synchronous angular
speed. The equivalent circuits in the stationary frame of reference then becomes as
shown in Figure 106 .

-102-
March 1, 2001,
Figure 106 The equivalent diagram in the stationary
frame of reference.

5.5 The electromagnetic torque


Properly the most important task for the induction motor is to produce a
torque on the shaft. The easiest way of finding the torque is by looking at the input
power in the three-phase system.
pin ' vsa @ i sa % vsb @ i sb % vsc @ i sc % vra @ i ra % vrb @ i rb % vrc @ i rc (137)

The input power goes to four things: losses, electromagnetic storage, mechanical
storage and the load. By using (115) it is possible to re-write (137) to
3
pin ' @ vsd @ i sd % vsq @ i sq % 2 @ vs0 @ i s0 % vrd @ i rd % vrq @ i rq % 2 @ vr0 @ i r0 (138)
2

and by substituting from (126) and (128) this gives

-103-
March 1, 2001,
2
rs @ isd % T @ 8sq @ i sd % p8sd @ i sd
2
% rs @ isq & T @ 8sd @ i sq % p8sq @ i sq
2
% 2 @ rs @ is0 % 0 % 2 @ p8s0 @ i s0
3
pin ' @ (139)
2 % r @ i2 % T & Tr @ 8rq @ ird % p8rd @ i rd
r rd

2
% rr @ irq & T & Tr @ 8rd @ irq % p8rq @ i rq
2
% 2 @ rr @ ir0 % 0 % 2 @ p8r0 @ i r0
Looking closer at(139) three types of powers may be identified. The first type is on
r @ i 2 -form and is the power lost as the cupper-loss in the stator and rotor windings.
The second type is on p8 @ i -form and describes the exchange of electromagnetic
energy between the magnetic parts and the winding. The last type is on T @ 8 @ i -form
and describes the energy converted to mechanical energy or the mechanical energy
converted to electrical energy. Note that zero-sequence currents and flux-linkage does
not contribute to the mechanical power The mechanical power then becomes
3
pel ' @ T @ 8sq @ isd & T @ 8sd @ i sq % T & Tr @ 8rq @ i rd & T & Tr @ 8rd @ i rq
2
(140)
3
' @ T @ 8sq @ isd & 8sd @ i sq % T & Tr @ 8rq @ ird & 8rd @ i rq
2
By using the relation pel ' Tr @ Tel , the developed torque becomes
) 3 1
Tel ' @ @ T @ 8sq @ i sd & 8sd @ isq % T & Tr @ 8rq @ i rd & 8rd @ irq (141)
2 Tr

At this stage it must be remembered that the model is developed for a two-pole
motor. Do the following thought-experiment: If the slip is disregarded, a two-pole
machine in steady-state supplied by 50 Hz will rotate at 3000 rev/min ,whereas a
four-pole machine will rotate at 1500 rev/min . If the motor rotates slower, it is as if
there was a gear connected. Using a gear to reduce the speed causes the torque to
increase. The correct torque for an arbitrary number of poles then becomes
) p 3 p
T el ' Tel @ ' @ @ T @ 8sq @ i sd & 8sd @ isq % T & Tr @ 8rq @ i rd & 8rd @ irq (142)
2 2 Tr @ 2

where p is the number of poles. Looking at (136) it is given that


8sd @ i sq & 8sq @ isd ' & 8rd @ i rq & 8rq @ ird ' Lm @ i rd @ isq & i rq @ isd (143)

Inserting into (142) the developed torque may be written on the flowing forms

-104-
March 1, 2001,
3 p
T el ' @ @ 8rq @ ird & 8rd @ i rq
2 2
3 p
' @ @ 8sd @ isq & 8sq @ i sd (144)
2 2
3 p
' @ @ Lm @ i rd @ isq & i rq @ isd
2 2
The same argument as for the developed electrical torque in (142) must be used when
calculating the angular speed. This is because the ratio between the angular speed
in electrical radians and mechanical radians is dependent on the number of poles of
the machine. The angular speed in electrical radians must be known, because it is
used in the equations (126) and (128) . The angular speed in electrical radians
becomes
p
Tr,el ' Tr,mek @ (145)
2

5.6 Solving the system


What remains is to find a strategy for solving the differential equations. Two
possibilities exist: Solving for the currents and then calculating the flux-linkages or
solving for the flux linkages and then calculating the currents.

5.6.1 Solving for the currents

Of the two strategies, is to solve for the currents perhaps the most obvious.
Looking first at a simple example shown in Figure 107 .
I

V L

Figure 107 A simple magnetic device.


The equation for the system is
d8 di
v ' r@i% ' r@i%L@ (146)
dt dt

and solving for the current this becomes

-105-
March 1, 2001,
1
i ' @ v & r @ i dt
L (147)
8 ' L@i
In this simple case, there is no real distinction between solving for the current or the
flux-linkage. Expanding (146) to the windings for an induction motor, but locking the
rotor to make the inductances constant in time the voltages on matrix-notation
becomes
d8 di
v ' r@i% ' r@i%L@ (148)
dt dt

and the currents are then found as


1
i ' @ v & r @ i dt
L (149)
8 ' L@i
The problem is, that in (149) is the inverse of the inductance-matrix used. Using
Gauss-Jordan elimination to find the inverse
Lls % L ss L sm L sm cos 2r cos 2r % 2B cos 2r % 4B /
1 0 0 0 0 0
3 3 0
0

L sm Lls % L ss L sm cos 2r % 4B cos 2r cos 2r % 2B /


0 1 0 0 0 0
3 3 0
0

L sm L sm Lls % L ss cos 2r % 2B cos 2r % 4B cos 2r /


0 0 1 0 0 0
3 3 0
0

cos 2r cos 2r % 4B cos 2r % 2B L lr % Lrr Lrm Lrm /


0 0 0 1 0 0
3 3 0
0

cos 2r % 2B cos 2r cos 2r % 4B Lrm L lr % Lrr Lrm /


0 0 0 0 1 0
3 3 0
0

cos 2r % 4B cos 2r % 2B cos 2r Lrm Lrm L lr % Lrr /


0 0 0 0 0 1
3 3 0
0

\
(150)
L ls % Lss % 2 @ L sm Lls % L ss % 2 @ Lsm L ls % Lss % 2 @ L sm 0 0 0 /
0
0
1 1 1 0 0 0 1%2%3
L ls % Lss % 2 @ L sm Lls % L ss % 2 @ Lsm L ls % Lss % 2 @ L sm 0 0 0 /
0
0
1 1 1 0 0 0 1%2%3
L ls % Lss % 2 @ L sm Lls % L ss % 2 @ Lsm L ls % Lss % 2 @ L sm 0 0 0 /
0
0
1 1 1 0 0 0 1%2%3
0 0 0 Llr % L rr % 2@ L rm Llr % L rr % 2@ L rm Llr % L rr % 2 @ Lrm /
0
0
0 0 0 1 1 1
4%5%6
0 0 0 Llr % L rr % 2@ L rm Llr % L rr % 2@ L rm Llr % L rr % 2 @ Lrm /
0
0
0 0 0 1 1 1
4%5%6
Llr % L rr % 2@ L rm Llr % L rr % 2@ L rm Llr % L rr % 2 @ Lrm
0 0 0 /
0
0
0 0 0 1 1 1
4%5%6
As is seen in (150), the inverse cannot be found, and this is because the matrix is
singular. The conclusion is therefore, that the system cannot be solved from the
currents (not this way at least).

-106-
March 1, 2001,
5.6.2 Solving for the flux-linkage

One way of successfully solving the system is to use the flux-linkage as the
tool. Using the equations (126) and (128), and modifying them in accordance with
Figure 106 is the stationary frame of reference is used, the flux linkage associated
with the d-, q-, and 0-axis are calculated as
Rotating frame of reference Stationary frame of reference

8sd ' usd & rsd @ isd & T @ 8sq dt 8s " ' us" & rs @ is" dt

8sq ' usq & rsq @ isq % T @ 8sd dt 8s $ ' us$ & rs @ is$ dt
(151)

8rd ' urd & rrd @ ird & T & Tr @ 8rq dt 8r " ' urd & rr @ ir" & Tr @ 8r$ dt

8rq ' urq & rrq @ irq % T & Tr @ 8rd dt 8r $ ' ur$ & rr @ ir$ % Tr @ 8r" dt

What remains is to find the currents from the flux-linkages. Observing that in both
the rotating frame of reference (Figure 105 ) and in the stationary frame of reference
(Figure 106 ) there is a T-joint in the middle as shown in Figure 108 .

Figure 108 The centre T-joint of the model.


Using this observation, the currents are calculated as

-107-
March 1, 2001,
lm
8s & @ 8r
lrl
is '
2
lm
lsl &
lrl
(152)
lm
8r & @ 8s
lrl
ir '
2
lm
lrl &
lsl

for the d- and q-axis by using mask-equations on the circuit and for the 0-axis it
becomes even simpler.

-108-
March 1, 2001,
This Page is Blank

-109-
March 1, 2001,
Appendix 1
Moments of Inertia

The bellow is a table of commonly used shapes and their moments of inertia.

Name Shape Centre of Moment of inertia


gravity
1 Circular J xx ' 1
2
mr 2 % 1
12
ml 2
Cylindrical
Shell J x1x1 ' 1
2
mr 2 % 1
3
ml 2

J zz ' mr 2

2r
2 Half x' J xx ' Jyy ' 12 mr 2 % 1
ml 2
Cylindrical B 12

Shell J x1x1 ' J y1y1 ' 12 mr 2 % 1


3
ml 2

J zz ' mr 2
J zz ' 1 & 4 mr 2
B2

3 Circular J xx ' 1
4
mr 2 % 1
12
ml 2
Cylinder
J x1x1 ' 1
4
mr 2 % 1
3
ml 2

J zz ' 1
2
mr 2

-110-
4r

March 1, 2001,
4 Semi-cylinder x' J xx ' Jyy ' 14 mr 2 % 121 ml 2
3B
J x1x1 ' J y1y1 ' 14 mr 2 % 13 ml 2

J zz ' 1
2
mr 2

J zz ' 1
2
& 16
mr 2
9B
2

5 Rectangular J xx ' 12
1
ma2 % l2
Parallelepipe
d J x1x1 ' 12
1
mb2 % l2

J zz ' 12
1
ma2 % b2

J y1y1 ' 12
1
mb 2 % 13 ml 2

J y2y2 ' 1
3
mb2 % l2

6 Spherical J zz ' 2
3
mr 2
Shell

r
7 Hemispherica x' J xx ' J yy ' J zz ' 23 mr 2
l Shell 2
J yy ' J zz ' 5
12
mr 2

8 Sphere J zz ' 2
5
mr 2

-111-
3r

March 1, 2001,
9 Hemisphere x' J xx ' J yy ' J zz ' 25 mr 2
8
J yy ' I zz ' 83
320
mr 2

1 Uniform J yy ' 1
12
ml 2
0 Slender Rod
J y1y1 ' 1
3
ml 2

2r
1 Quarter- x'y' J xx ' J yy ' 12 mr 2
1 Circular Rod B
J zz ' mr 2

1 Elliptical J xx ' 1
4
ma 2 % 1
12
ml 2
2 Cylinder
J yy ' 1
4
mb 2 % 1
12
ml 2

J zz ' 1
4
ma2 % b2

J y1y1 ' 1
4
mb 2 % 13 ml 2

2h
1 Conical Shell z' J yy ' 1
4
mr 2 % 12 mh 2
3 3
J y1y1 ' 1
4
mr 2 % 16 mh 2

J zz ' 1
2
mr 2

J yy ' 1
4
mr 2 % 1
18
mh 2

-112-
4r

March 1, 2001,
1 Half Conical x' J xx ' Jyy ' 14 mr 2 % 12 mh 2
4 Shell 3B
2h J x1x1 ' J y1y1 ' 14 mr 2 % 16 mh 2
z'
3 J zz ' 1
mr 2
2

J zz ' 1
2
& 16
mr 2
9B
2

3h
1 Right- z' J yy ' 20
3
mr 2 % 35 mh 2
5 Circular Cone 4
J y1y1 ' 20
3
mr 2 % 1
10
mh 2

J zz ' 10
3
mr 2

J yy ' 20
3
mr 2 % 3
80
mh 2

r
1 Half Cone x' J xx ' Jyy ' 203 mr 2 % 35 mh 2
6 B
3h J x1x1 ' J y1y1 ' 3
mr 2 % 1
mh 2
z' 20 10

4 J zz ' 3
mr 2
10

J zz ' 3
& 1 mr 2
B
10 2

3c
1 Semiellipsoid z' J xx ' 1
5
mb2 % c2
7 8
J yy ' 1
5
ma2 % c2

J zz ' 1
5
ma2 % b2

J xx ' 1
5
mb2 % 19
64
c2

J yy ' 1
5
ma2 % 19
64
c2

-113-
2c

March 1, 2001,
1 Elliptic z' J xx ' 1
6
mb 2 % 12 mc 2
8 Paraboloid 3
J yy ' 1
6
ma 2 % 12 mc 2

J zz ' 1
6
ma2 % b2

J xx ' 1
6
m b 2 % 13 c 2

J yy ' 1
6
m a 2 % 13 c 2

a
1 Rectangular x' J xx ' 1
10
mb2 % c2
9 Tetrahedron 4
b J yy ' 1
ma2 % c2
y' 10

4 J zz ' 1
ma2 % b2
c 10
z'
4 J xx ' 3
80
mb2 % c2

J yy ' 3
80
ma2 % c2

J zz ' 3
80
ma2 % b2

a 2 % 4R 2
2 Half Torus x' J xx ' J yy ' 12 mR 2 % 58 ma 2
0 2BR
J zz ' mR 2 % 34 ma 2

-114-
March 1, 2001,
This Page is Blank

-115-
March 1, 2001,
Appendix 2
Simulating differential equations using
Matlab/Simulink
Many of the exercises are assumed to be solved by Matlab/Simulink, because
it is a nice tool for solving differential equations. However, it may not be evident how
to get a solution. Using an example from chapter 1 it is shown how to do this.
The example is the mechanical equation
dTr,mec
Te ' J @ % B @ Tr,mec % Tc % T stic % Tl (153)
dt
The state in this case is the mechanical angular speed, Tr,mec. The equation in (153)
is written on form, which is how equations often appear from the theoretical analysis.
The equation must be re-written on integral form in order to be able to solve it. First
step is to gather all the constants on one side as
dTr,mec
T e & Tc & T stic & Tl ' J @ % B @ Tr,mec (154)
dt
because an iterative method is usually used to find the result, it is assumed that the
solution for the previous step is known. Therefore the state-variable may be moved
as
dTr,mec
T e & Tc & T stic & Tl & B @ Tr,mec ' J @ (155)
dt
the next step is to move the moment of inertia to the other side as well
dTr,mec
1
J
@ T e & T c & T stic & T l & B @ Tr,mec ' (156)
dt
and finally the state variable may be found by integration as
t
1
J
@ Te & T c & Tstic & T l & B @ Tr,mec ' Tr,mec (157)
0

Modelled in Matlab/Simulink this becomes as shown in Figure 129 .

-116-
March 1, 2001,
Figure 129 The mechanical system solved in Matlab/Simulink.

-117-
March 1, 2001,
This Page is Blank

-118-
March 1, 2001,
Appendix 3
Litterature

Here are listed the books which are used as an inspiration when writing this
text.

Name Title Publisher ISBN

Syed Nasar Electric Machines and Electromechanics Schaums Outlines 7045994


Second Edition

J.L. Meriam Engineering Mechanics John Wiley & 471849103


L.G. Kraige Volume 2, Dynamics Sons
Second Edition

Davis K. Cheng Field and Wave Electromagnetics Addison.Wesley 201528207


Second Edition

Chee-Mun Ong Dynamic Simulation of Electric Prentice Hall 137237855


Machinery using MATLAB/SIMULINK

Werner Leonhard Control of Electric Drives Springer Verlag 3540593802

-119-
March 1, 2001,
This Page is Blank

-120-
March 1, 2001,
Exercises
Relevant Exercises for the Chapters
E.1 Exercises for Chapter 1
E.1.1
A trailer with a mass of 8000 kg is attached to a car. Over one minute is the
speed increased from 30 km/h to 60 km/h. What is the acceleration? The friction and
the drag from the air are together 2000 N. By what force does the trailer pull on the
car?

E.1.2
Every minute flows 1200 l of water through a turbine. The water falls 12 m.
The efficiency of the turbine is 70%. Calculate the power of the turbine.

E.1.3
A car is driving with the speed 60 km/h. The mass of the car is 1200 kg. How
large is the energy stored as kinetic energy? The car uses 32 metres to come to a full
stop. What is the average force from the brakes during the deceleration? If the speed
was 120 km/h, how far does the car travel when the same force is applied from the
brakes?

E.1.4
An iron rod has the length 80 cm and the cross-section 1 cm2. The rod rotates
along an axis perpendicular to the centre-axis of the rod. The rod rotates with 8
rev/sec. The mass density of iron is 7.9 t/m3. Calculate the kinetic energy.

E.1.5
Show using (10) that the moment of inertia for a solid cylinder along its
centre-axis is J ' 12 mr . What is the radius of gyration?

E.1.6
Show using (10) that the moment of inertia for a solid sphere is J ' 2
5
mr .
What is the radius of gyration?

E.1.7
The swinging pendulum test is done on an object. The wires are 0.9m, the
observed frequency is 0.498 Hz and the mass is 50 g. What is the moment of inertia?

-121-
March 1, 2001,
E.1.8
A run-out test is done to determine the moment of inertia. In one test a time-
constant of 20 seconds is observed. In a second test with an attached extra mass with
the moment of inertia, Jtest=1 kgm2, the time-constant is 43 seconds. Calculate B and
J.

E.1.9
An electrical machine produces a constant torque of 40 Nm and has a load
torque of 20.5 Nm. Using the results from E.1.7 and when Tc is measured to 1.5 Nm,
what is the angular speed in steady-state? What is the angular speed as a function
of time?

E.1.10
The load is a fan connected to the shaft. The fans load-characteristic may be
describes by the function T l ' K @ T2 where K ' 175 @ 10&6 and with all the other
parameters as in E.1.8, what is the angular speed as a function of time? Comment
on the differences between E.1.8 and E.1.9.

-122-
March 1, 2001,
E.2 Exercises for Chapter 2
E.2.1

Through a wire flows a DC-current of 50A. The wire is 100 cm long. The wire
is placed in a uniform magnetic field with a flux density of 1 Tesla. Calculate the
force acting on the wire. If the wire is moved at 5 m/s, calculate the power. Calculate
the induced voltage in the wire. Show that the mechanical power equals the electrical
power.

E.2.2
A system consists of two mutually coupled coils as shown below.

The inductances for the system are: L11=A, L22=B, L12=L21=C@cos1. Find the torque
when i1=I0 and i2=0. Find the torque when i1=i2=I0. Find the torque when i1=Im@sinTt
and i2=I0. Find the torque when i1=i2=Im@sinTt. Find the torque when coil 1 is short-
circuited and i2=I0.

E.2.3
A relay is constructed as shown below.

-123-
March 1, 2001,
Show that the force acting on the yoke is
@ N2 @ i2 @ A
F ' 0
2
4 @ l0

where A is the area of the air-gap and the magnetic material is linear. Determine an
expression for the force on the yoke when the air-gap is closed. If the force acting on
the yoke is 20N, the area of the air-gap is 8cm2, the number of turns on the coil is 200
and the length of the air-gap is 1cm, what is the current?

E.2.4
A relay is constructed as shown below.

With a coil places as shown on the figure, find the ampere-turns (N@I) for the coil, if
it has to produce a force of 60N and where the air-gap is 1cm. Determine the force on
the yoke when the air-gap is closed and where the relative permeability for the iron
is 450. Show the force acting on the yoke as a function of the air-gap-length.

-124-
March 1, 2001,
E.2.5
In an inductor or a capacitor flows the current shown in the figure.

The inductance is 5H and the capacitor is 2F. The voltage over the capacitor at t=0
is 0V. Show the curves for the voltage over the capacitor and inductor. Show the
curves for the power and energy in the inductor. If the inductance is 10mH, calculate
the energy in the inductance at t=2.

-125-
March 1, 2001,
E.3 Exercises for Chapter 3
E.3.1

Calculate the induced voltage in the armature of a four-pole lap-wound dc-


motor having 728 active conductors and running at 1800 rpm. The flux per pole is 30
mWb. What is the voltage if the motor is wave-wound? For the lap-wound motor, if
the armature is designed to carry a maximum current of 100A, what is the power
developed by the armature? If the motor is re-connected as a wave-wound motor,
what is the power? What is the developed torque for the two motors?

E.3.2
A shunt generator has the following data: PN=100kW, Va=230V, Ra=0.05S, Rf=57.5S

Neglecting the voltage drop over the brushes, calculate the voltage induced at full-
load and at half-load.

E.3.3
A compound generator has the following data: PN=50kW, Va=250V, Ra=0.06S,
Rse=0.04S, Rf=125S, Vbrush=2V

Calculate the voltage induced at full-load.

-126-
March 1, 2001,
E.3.4
A separately excited dc-generator has a constant loss of Pc in the field winding. The
motor has an armature voltage Va, an armature current Ia and an armature
resistance Ra. At what value of Ia does the generator operate at maximum efficiency?

E.3.5
This is a Simulink-exercise.
A permanent magnet DC-motor has the following characteristics: Va=60V,
La=0.64e-3H, Ra=0.25S, kM=0.13, J=2.6@10-4Nm, B=1.7189@10-4, Tc=0.0670Nm,
Tl=5Nm. At some time instances the following happens: t=0.01s 6 the voltage is
applied to the motor, t=0.03s 6 the voltage is reversed, t=0.05s 6 the voltage is
reversed again, t=0.10s 6 the load is engaged. Show the angular speed of the motor
in the time-interval t=0.0s 6 0.20s.

E.3.6
This is a Simulink-exercise.
A separately excited DC-motor has the following characteristics: Va=180V,
La=3.5e-3H, Ra=1.43S, J=3.693Nm, B=0.0145, Tc=0.0793Nm,Vf=190V, La=29.6e-3H,
Ra=399S, Tl=5Nm. The relationship between the field current and field constant, kM,
is measured as If= [0 0.144 0.194 0.250 0.300 0.351 0.400 0.441], kM=[0 0.388 0.498
0.596 0.669 0.729 0.771 0.808]. At t=0s is the voltage applied to the field winding, at
t=0.1s is the voltage applied to the armature, at t=100 is the load engaged, at t=150s
is the armature voltage reduced by 10%, at t=200s is the armature voltage reduced
by 20%. Show the angular speed of the motor in the time-interval t=0s 6 300s.

E.3.7
This is a Simulink-exercise.
Using the data and solution from the previous exercise. The voltage source for the
armature winding has a maximum long term current capability of 9 A. The
maximum current is therefore set to 8A. Modify the model so that this is achieved.
Show the angular speed, voltage and current.

-127-
March 1, 2001,
E.4 Exercises for Chapter 4
E.4.1
This is a Matlab-exercise.
Using the single-phase equivalent circuit, calculate the mechanical power and torque
of an induction motor with the following parameters:
Rs = 0.14S Stator resistants
Rr = 0.20S Rotor resistants
Xs = 0.45S Stator reactants
Xr = 0.40S Rotor reactants
Rfe = 300S Iron-core resistants
Xh = 13S Magnetizing reactants
U = 230V Supply voltage
f_n = 50Hz Frequency

What happens if the rotor resistance is increased by 20%?

-128-
March 1, 2001,
This Page is Blank

-129-
March 1, 2001,
Solutions
Solutions to the problems (do not cheat!!)
S.1 Solutions for Chapter 1
S.1.1
m=8000kg, V1=30 km/t, V2=60 km/t, t=1 min, Fg=2000 N

V2 & V1
a '
)t
60 & 30 @ 1000
'
1 @ 60 @ 3600
' 0.138 m 2
s

F ' Fg % F a
' Fg % m @ a
' 2000 % 8000 @ 1.38
' 3.11 kN
S.1.2
V=1200l, h=12m, =70%

)Epot ' m @ g @ h
' 1200 @ 9.82 @ 12
' 141.4kJ
)E
P ' @0
)t
141408
' @ 0.70
60
' 1.65 kW

-130-
March 1, 2001,
S.1.3
m=1200kg, v=60km/t

a)
2
1 1 1000
Ekin ' @ m @ v2 ' @ 1200 @ 60 @ ' 166.7 kJ
2 2 3600
b)
s=32m

v2 ' 2 @ a @ s
\
2
1000
60 @
v2 3600 m
a ' ' ' 4.34
2@s 2 @ 32 s2

F ' a @ m ' 4.34 @ 1200 ' 5.2 kN

c)
v=120 km/t
v2
F ' m@a ' m@
2@s
\
2
1000
2
120 @
v 3600
s ' m@a ' m@ ' 1200 @ ' 128 m
2@F 2 @ 5208

-131-
March 1, 2001,
S.1.4
l=80cm, q=1cm^2, D=7.910^3kg/m^3, n=8 rev/sec

m ' l@q@D
' 0.8 @ 1 @ 10&4 @ 7.9 @ 103
' 0.632 kg
1
Ekin ' @ J @ T2
2
1 1
' @ @ m @ l 2 @ T2
2 12
1 1
' @ @ 0.632 @ 0.82 @ 8 @ 2B 2
2 12
' 42.58 J
S.1.5

2
J ' r0 dm

2B r
3
' Dt r0 dr0 d2
0 0

Br 4
' Dt , m'B @ r 2 @ t @ D
2
' 1
2
mr 2

J
k '
m
r
'
2

-132-
March 1, 2001,
S.1.6
Seeing that the moment of inertia of a slice of the sphere equals the solution
in S.1.5 and then integrating over the volume of the sphere the moment of inertia
becomes
dJxx ' 12 dm y 2

' 1
2
B D y 2 dx y 2

BD
' 2
r & x 2 dx

\
r
BD
Jxx ' 2
r & x 2 dx
&r
4
' 8
15
B D r5 , m' @ B @ r 3@ D
3
' 2
5
m r2

Jxx
k '
m

2
' r
5

-133-
March 1, 2001,
S.1.7
f=0.498Hz, r =0.9m. M=50g
1 gr
f ' @
2B k2
\

gr
k '
f2B 2

9.82 @ 0.9
'
0.498 @ 2B 2
' 0.950 m
J ' k2 @ m
' 0.9502 @ 0.050
' 45.134 @ 10&3 kgm 2
J ' J0 % r 2 @ m
\
J0 ' J & r 2 @ m

' 45.134 @ 10&3 & 0.92 @ 0.05


' 4.6341 @ 10&3 kgm 2

-134-
March 1, 2001,
S.1.8
J1=20 sec, J2=43 sec, Jtest=1 kgm2
J
J1 '
B
J % Jtest
J2 '
B
\
J ' J1 @ B
J1 @ B % Jtest
B '
J2
\
B ' 43.48 @ 10&3 Nms/rad
J ' 0.870 kgm 2

-135-
March 1, 2001,
S.1.9
Te=40Nm, Tl=20.5Nm, Tc=1.5 Nm
dT
Te ' J @ % B @ T % T c % Tl
dt
dT
0 ' J@ \
dt
Te % T c % Tl
T '
B
40 % 1.5 % 20.5
'
43.48 @ 10&3
' 413.98 rad/s

40
Te
1
1.5 1/0.870
s
Tc
20.5 1/J
Tl
43.48e-3

S.1.10
T l ' K @ T2 and K ' 175 @ 10&6

40
Te
1
1.5 1/0.870
s
Tc
1/J

43.48e-3

175e-6

-136-
March 1, 2001,
S.2 Solutions for Chapter 2
S.2.1
i=50 A, l=100 cm, B=1 Tesla

Using the Lorenz-force, the force on the wire becomes


F ' B @ i @ l ' 1 @ 50 @ 1 ' 50 N

the power used to move the wire is


W ' F @ )x

W F @ )x
P ' ' ' F @ v ' 50 @ 5 ' 250 W
)t )t
using Faradays law, the induced voltage becomes
d8
e ' & ' B @ l @ v ' 1 @ 100 @ 10&2 @ 5 ' 5 V
dt

The mechanical power is 250 W and the electrical power is P ' e @ i ' 5 @ 50 ' 250W .the
two powers match and the energy is in balance.

S.2.2
i1 ' I0 i2 ' 0

Wm ' 1 @ A @ I02
2
dWm ' Fe @ dx ' Fe @ r @ d1 ' Te @ d1
\
MW m
Te ' '0
M1
i1 ' i2 ' I0

Wm ' 1 @ A % B @ I02 % C @ I02 @ cos1


2
Te ' &C @ I02 @ sin1
i1 ' ImsinTt i2 ' I0

Wm ' 1 @ A @ Im2 @ sin2 Tt % 1 @ B @ I0 % C @ I0 @ Im @ sinTt @ cos1


2 2
Te ' C @ I0 @ Im @ sinTt @ sin1

-137-
i1 ' I msinTt i2 ' I msinTt

March 1, 2001,
2
Te ' C @ I m @ sin2Tt @ sin1

Last question is a little tougher


i1 ' short i2 ' I0

d L11 @ i1 % L12 @ i2
' 0
dt
\
L11 @ i1 % L12 @ i2 ' k
\
k & L12 @ I0
' i1
A
2
1 k & L12 @ I0 1 2 k & L12 @ I0
Wm ' @A@ % @ B @ I0 % L12 @ Io @
2 A 2 A
2 2
1 k 2 1 L12 @ I0 1 2
' @ & @ % @ B @ I0
2 A 2 A 2
2
1 k2 1 2 1 I
' @ % @ B @ I0 & @ 0 @ C 2 @ cos21
2 A 2 2 A
MWm
Te '
M1
2
I0
' @ C 2 @ cos1 @ sin1
A

-138-
March 1, 2001,
S.2.3
The problem is solved by first examining the reluctance network as shown below.

The reluctance of one of the air-gaps is


l0
U0 '
0 @ A
and because the two air-gaps is in series, the combined reluctance for the air becomes
2 @ l0
U0 '
0 @ A
the reluctance of the iron is very small compared with the reluctance of the air-gap,
especially when the permeability is significantly larger than one. From the reluctance
network the flux is found as
N @ i @ 0 @ A
M ' '
U0 2 @ l0
the force acting on the yoke is then found as
2 2
1 dU0 1 0 @ N 2 @ i 2 @ A 2 2 0 @ r @ N 2 @ i 2 @ A
Fm ' & @M @
2
' @ @ '
2 dx 2 4@l
2 0 @ A 4@l
2
0 0

When the air-gap is closed is the solution much the same, only now is the reluctance
of the air-gap zero and the reluctance giving the flux is only that of the iron core, but
the change in reluctance is that of the air-gap. The result then becomes

-139-
March 1, 2001,
dU dU0 dUj dU0 &2
U ' U 0 % Uj Y ' % / U ' '
dx 0 '0
d
dx dx j
dx 0 @ A
dx

0 @ r @ N @ i @ A
M ' ' '
U /0U0'0 Uj lj

2 2 2
1 dU 1 0 @ r @ N 2 @ i 2 @ A 2 &2 r @ 0 @ N 2 @ i 2 @ A
Fm ' & @ M2 @ ' & @ @ '
2 dx 2 l
2 0 @ A 2
lj
j

Fm=20N, A=8cm2, N=200, l0=1 cm


0 @ N 2 @ i 2 @ A
Fm '
2
4 @ l0

2
F m @ 4 @ l0 20 @ 4 @ 1 @ 10&2 2
i ' ' ' 14.10 A
0 @ N 2 @ A 4 @ B @ 10&7 @ 2002 @ 8 @ 10&4

-140-
March 1, 2001,
S2.4
The core is cut through the middle leg, and the two parts are placed on top of each
other as shown in the figure.

The problem is then similar to the previous one.


2
0 @ r @ N 2 @ i 2 @ A
Fm '
2
4 @ l0
\
2
F m @ 4 @ l0 60 @ 4 @ 1 @ 10&2 2
N@i ' ' ' 1954 A @ vnd
0 @ A 4 @ B @ 10&7 @ 50 @ 10&4

The same way for the closed air-gap


N2 @ i2 @ A 19542 @ 50 @ 10&4
' 4502 @ 4 @ B @ 10&7 @
2
F m ' r @ 0 @ ' 7.59 kN
lj
2
80 @ 10&2 2
The force as a function of the air-gap

-141-
March 1, 2001,
S2.5
L= 5H, C=2F
di 1
Vl ' L @ , Vc ' @ i dt
dt C

Remember that at t=0 there is already a current flowing through the inductor,
meaning that there already is energy in it.

-142-
March 1, 2001,

-143-
March 1, 2001,
S.3 Solutions for Chapter 3
S.3.1
Because it is a lap-wound motor there are as many parallel paths for the current as
there are poles.
N=728, n=1800, Mf=30mWb
d8
EMF lap ' &
dt
' N @ Mf @ T
n
' N @ Mf @ @2@B
60
1800
' 728 @ 30 @ 10&3 @ @2@B
60
' 4.117 kV
In a wave-wound motor there are always two paths for the current. This is twice that
of the lap-wound motor and the induced EMF therefore is
EMF wave ' 2 @ EMFlap ' 8.233 kV

There are 4 paths for the current in the lap-wound motor. The current in each of the
windings of the armature will therefore be
Ia 100
' ' 25 A
4 4
The power in each of these windings is
)
Pa ' 25 @ 4.117 @ 103 ' 102.91 kW

This then gives the total power as


P a ' 4 @ 102.91 kW ' 411.67 kW

For the wave-winding there are only two paths for the current. Each of the may carry
25A, giving a total current of 50A. The voltage is double that of the lap-wound motor.
This means that the power must be exactly the same as for the lap-winding.
The torque is

-144-
P a ' T @ Te

March 1, 2001,
\
Pa 411.67 @ 103
Te ' ' ' 2184 Nm
T 1800
2@B@
60
S.3.2
First calculate the field current
Va 230
If ' ' ' 4A
Rf 57.5
Then calculate the current flowing from the generator
Pa 100 @ 103
Il ' ' ' 434.8 A
Va 230
Looking at the figure it is seen that the sum-current must be
Ia ' I l % If ' 434.8 % 4 ' 438.8 A

The voltage drop over the armature resistor is


VR ' Ia @ Ra ' 438.5 @ 0.05 ' 22 V
a

The induced voltage is therefore


ea ' VR % Va ' 230 % 22 ' 252 V
a

The field current and the armature voltage is the at half-load, but the armature is
only the half of full-load
I 434.8
Ia ' l % If ' % 4 ' 221.4 A
2 2
The voltage drop over the armature resistor becomes
VR ' Ia @ Ra ' 221.4 @ 0.05 ' 11 V
a

The induced voltage becomes


ea ' VR % Va ' 230 % 11 ' 241 V
a

-145-
March 1, 2001,
S.3.3
First calculate the load current
P 50 @ 103
Il ' ' ' 200 A
V 250

Then calculate the voltage drop over the series winding


VR ' Il @ Rse ' 200 @ 0.04 ' 8 V
se

The field voltage is then


Vf ' VR % V ' 250 % 8 ' 258 V
se

The field current is then found as


Vf 258
If ' ' ' 2.06 A
Rf 125
The armature current is then
I a ' Il % I f ' 200 % 2.06 ' 202.6 A

The voltage drop over the armature resistance then becomes


VR ' Ra @ I a ' 0.06 @ 202.6 ' 12.12 V
a

The induced voltage is then


ea ' V % VR % VR % Vbrush ' 250 % 12.12 % 8 % 2 ' 272.12 V
a se

S.3.4
The output power is
P out ' ea @ I a

The input power is


2
Pin ' ea @ Ia % I a @ Ra % Pc

the efficiency is
P out ea @ I a
0 ' '
Pin 2
ea @ I a % Ia @ Ra % P c
The maximum efficiency is found by differentiating with regard to Ia

-146-
March 1, 2001,
d0
' 0
dI a
\
2
0 ' E @ E @ I a % Ia @ Ra % P c % E @ I a @ E % 2 @ Ia @ R a
\

Pc
Ia '
Ra
which means that the maximum efficiency is achieved when the armature loss equals
the constant loss in the field winding.

-147-
March 1, 2001,

-148-
S.3.5
March 1, 2001,

-149-
S.3.6
March 1, 2001,

-150-
S.3.7
March 1, 2001,
S.4 Solutions for Chapter 4
S.4.1
% Script-file for calculating the developed torque of an induction
motor as a function % of the slip
%
% Rasmus Post, 10/8-2000
%
echo on
% Rs Xs Xr Rr
% _____ _____ _____ _____
% ---|_____|----|_____|-------|_____|----|_____|---
% | |
% ------- |
% | | |
% - - -
% | | | | | |
% Rfe | | | | Xh | | Rr(1-s)
% | | | | | | -------
% - - - s
% | | |
% -------------------------------------------------
%
% The shaft power is dissipated in Rr(1-s)
% -------
% s
echo off
clear all
close all
format compact

Rs = 0.14 %Stator resistans


Rr = 0.20 %Rotor resistans
Xs = 0.45 %Stator reactance
Xr = 0.40 %Rotor reactance
Rfe = 300 %iron core resistans
Xh = 13 %magnetizing reactance
U = 230; %Supply voltage
f_n = 50; %Supply frequency

dummy = 'Parameters are found at 50 Hz';


disp(dummy);
dummy = 'At which frequency should the calculations be done at?';
disp(dummy);
f = input('Frequency? = ');
omega_s = 2*pi*f;
s = 1:-0.001:-1;

-151-
s(find(s==0))=nan;

March 1, 2001,
omega = omega_s*(1-s);
Xs = Xs/f_n*f;
Xr = Xr/f_n*f;
Zs = Rs+i*Xs;
Zh = Rfe*i*Xh/(Rfe+i*Xh);
Zr = Rr./s+i*Xs;
Z = Zs+Zh.*Zr./(Zh+Zr);
I = U./Z;
Ir = I.*Zh./(Zh+Zr);
P = abs(3*Ir.^2.*Rr.*(1-s)./s).*sign(s);
M = P./omega;
Ps = abs(3*I.^2.*Rs).*sign(s);
Pr = abs(3*Ir.^2.*Rr).*sign(s);
Pfe = abs(3*(I-Ir).^2.*Rfe).*sign(s);
subplot(2,2,1)
plot(omega,abs(I).*sign(s))
xlabel('\omega [^r^a^s_s]')
ylabel('I_s [A]')
title('Stator current')
text(2*pi*f,0,'\leftarrow \omega_s')
axis tight
grid on
subplot(2,2,2)
plot(omega,abs(Ir).*sign(s))
xlabel('\omega [^r^a^d_s]')
ylabel('I_r [A]')
title('Rotor current')
text(2*pi*f,0,'\leftarrow \omega_s')
axis tight
grid on
subplot(2,2,3)
plot(omega,P)
xlabel('\omega [^r^a^d_s]')
ylabel('P_s_h_a_f_t [W]')
title('Shaft power')
text(2*pi*f,0,'\leftarrow \omega_s')
axis tight
grid on
subplot(2,2,4)
plot(omega,M)
xlabel('\omega [^r^a^d_s]')
ylabel('T_s_h_a_f_t [M]')
title('Shaft torque')
text(2*pi*f,0,'\leftarrow \omega_s')
axis tight
grid on
figure

-152-
subplot(2,2,1)

March 1, 2001,
plot(omega,Ps)
xlabel('\omega [^r^a^s_s]')
ylabel('P_s_t_a_t_o_r [W]')
title('Stator loss')
text(2*pi*f,0,'\leftarrow \omega_s')
axis tight
grid on
subplot(2,2,2)
plot(omega,Pr)
xlabel('\omega [^r^a^s_s]')
ylabel('P_r_o_t_o_r [W]')
title('Rotor loss')
text(2*pi*f,0,'\leftarrow \omega_s')
axis tight
grid on
subplot(2,2,3)
plot(omega,Pr)
xlabel('\omega [^r^a^s_s]')
ylabel('P_fe [W]')
title('Iron loss')
text(2*pi*f,0,'\leftarrow \omega_s')
axis tight
grid on

By increasing the resistance of the rotor the losses are increased, but the starting
torque is also increased!!

-153-

Вам также может понравиться