You are on page 1of 85

Sudan University of Science and Technology

College of Graduate Studies

Efficiency Evaluation of Induction Motor Using


Synthetic Loading Technique

A Thesis Submitted in Partial Fulfillment of the Requirement


for the Degree of M.Sc. in Electrical Engineering (Power)

Submitted by:
Hassab Elrasoul Zain Elabdien Eltigani

Supervised by:
Dr. Abdelaziz Yousif Mohamed Abbas

September 2012

1
Dedication

To whom she dressed me the life,

Mother
To whom he gave me a beautiful pain,

Father
The suns that burn to light for us,

Teachers
To whom I share with them the sorrow and
sweet,

Friends
Whenever commit a fault we will find
forgiveness,

Brothers and Sisters

ii
Acknowledgement

All praise is to Allah, the lord of the world, the almighty, with whose
gracious help it is possible to accomplish this work, and my prayers and peace
be upon Mohammed the last of the messengers.
I would like to express my deep appreciation and gratitude to my supervisor,
Dr. Abdelaziz Yousif Mohamed Abbas, for his patient guidance and his generous
support and encouragement. Also I would like to thank all staff of Electrical
Engineering Department. Also I would like to thank my family and colleagues.

iii
Contents
Dedication... i
Acknowledgement.....ii
Contents ..iii
Abstract ...vi
.............................................................................................................. vii

Chapter One: Introduction


1.1 Introduction...1

1.2 Problem Statement.. ..2

1.3 Description of Motor Losses.4

1.4 Synthetic Loading.5

1.5 Thesis Out Lines...6

Chapter Two: Literature Review and Background study

iv
2.1 Introduction...8
2.2 Standard Efficiency Test Methods...8
2.3 Induction Motor Field Efficiency Evaluation Methods...11
2.4 Physical Nature of Basic Methods...13
2.4.1 Nameplate Method..14
2.4.2 Slip Method. 16
2.4.3 Current Method17
2.4.4 Statistical Method.18
2.4.5 Accuracies of Basic Methods19
2.5 Synthetic Loading Method.20
2.5.1 Dual Frequency Method..22
2.5.2 Sweep Frequency Method...23
2.5.3 Constant Speed of Rotating Magnetic Field Method..23
2.5.4 Current Control Method.24
2.6 Summary25

Chapter Three: Modeling of Induction Motor

3.1 Introduction .26


3.2 The Park Transformation Theory.....26
3.3 Mathematical Model....28
3.4 Calculation of Losses32
3.5 Description of Modeling...33
3.6 Summary33

Chapter Four: Results and Discussion

v
4.1 Introduction ....34
4.2 Results.34
4.3 Conventional Method Results.35
4.4 Synthetic Loading Technique.39
4.4.1 Dual Frequency Results.39
4.4.2 Constant Speed of Rotating Magnetic Field Results.....44
4.5 Discussion....48

4.6 Summary ...52

Chapter Five: Conclusions and Recommendations


5.1 Conclusions.53
5.2 Recommendations...54
Appendices................................................................................................55
References .......71

vi
Abstract

This thesis investigates the application of the synthetic loading technique for
efficiency evaluation of induction machines. The standard tests require
specialist test facilities, additional machines, and for large machines, linear
machines, or vertical mounted machine and floor space. Therefore, an
efficiency test method that avoids the need for an external mechanical load is
desirable. Synthetic loading can determine machine losses and eliminates the
need for a mechanical load connected to the test machine. The synthetic loading
technique forces the machine under test to accelerate and decelerate using
power electronics as power source so the machine alternating between motor-
generator actions. If configured correctly the machine, on average over each
synthetic loading cycle, operates at rated rms current, rated rms voltage and
rated speed, thus producing rated copper loss, iron loss and friction and windage
loss. The thesis considers how to properly configure synthetic loading for
induction machines. The simulation results show that the synthetic loading
technique is capable of evaluating the efficiency of the induction machines.

vii


.

.

- .


.

.

viii
ix
Chapter One

Introduction
1.1 Introduction:
Induction motor is an alternating current motor that the rotor does not receive
electric power by conduction but by induction. It consists of two parts; stator
and rotor. The stator is made up of a number of stampings, which are slotted to
receive the winding. It is wound for a definite number of poles, the exact
number of poles being determined by the requirement of speed. The rotor is
divided into two types; squirrel cage and slip ring. In squirrel cage the rotor has
a cylindrical laminated core with parallel slots for caring the rotor conductors
which consist of heavy bars of copper, aluminium or alloys. The slip ring is
distributed winding consisting of coils and it wound for as many poles as the
number of stator poles. But the squirrel cage is most used because of its
specification like its simplest design [1].
Induction motors are by far the most common consumers of generated
electricity in the developed world today. The three-phase induction motor has
been described as the workhorse of industry and more recently racehorse of
industry when used with power electronic controllers. The induction machine
is produced in a very wide range of outputs from fractional to large multi
megawatts units. The induction machine is used in a wide variety of
applications as a mean of converting electric power to mechanical power, pump,
steel mill and hoist drives are a few applications of large multiphase induction
motors. On smaller scale the two phase servomotor is used in position-follow-
up control systems, and single phase induction motors are widely used in house
hold appliances as well as in hand and bench tools. The induction machine is
extensively used for various kind industrial drives for the following main
advantages:

1
It is very simple, extremely rugged and unbreakable construction.

It is cost low and it very reliable.

It has sufficiently high efficiency. In normal running condition, no


brushes are needed, hence friction losses are reduced.
And there are also some disadvantages like:
Its speed cannot be varied without sacrificing some of its efficiency
It is speed decreases with increase in load.
When a motor is switched on, there is a high inrush current from the mains
which may appear, especially if the power line section is inadequate, cause a
drop in voltage likely to affect receptor operation. This drop may be severe
enough to be noticeable in lighting equipment. To overcome this, some sector
rules prohibit the use of motors with direct on-line starting systems beyond a
given power. There are several starting systems which differ according to the
motor and load specifications. The choice is governed by electrical, mechanical
and economic factors. The kind of load driven is also important in the choice of
starting system. Main starting modes
Direct on-line starting.

Star-delta starting.

Part winding motor starting.

Resistance stator starting.

Autotransformer starting.

Slip ring motor starting.

Soft starter starting/slackening.

Frequency converter starting.

2
1.2 Problem Statement
Testing of induction machines to determine the energy dissipated as heat and
the resulting temperature rise is important for both users and manufactures.
High temperature cause deterioration of insulation materials and high rates of
power dissipation implies low efficiency values. The power losses of
induction motors are categorised as electrical and mechanical
losses. Some of the electrical losses can be predicted reliably
from the machine design data but some components such as
stray load losses and mechanical losses are less predictable.
Similarly, there is still a substantial amount of uncertainty
attached to the prediction of the cooling properties of the
machine [3].
The efficient use of energy has found new prominence with
environmental agencies raising the awareness of greenhouse
gas emissions predominantly and finite resources as a backup
argument. Previously, it was well understood that an efficient
motor would reduce the running costs of the machine over its
working life, typically a hundred times greater than the
purchase price of the machine. Recently, however, lobbying
has seen government incentives and policies introduced to
encourage the use of energy efficient equipment and
particularly energy efficient motors. Motor manufacturers have
introduced a separate category referred to as energy efficient
motors".
Despite the attractiveness of such energy efficient motors
they are still only used in relatively small numbers. It has been
reported that one of the reasons for the low acceptance of

3
energy efficient motors is poor information on these machines.
Further to this there are no energy efficiency standards for
motors.
A motor labelled as "high efficiency" from one supplier might
be less efficient than the standard motor from another supplier.
To compound the problem, there is no single efficiency
standard test method for induction motors that is used
throughout the industry. And to make matters worse, a recent
article identifies the current problems with existing methods of
measuring induction motor efficiency and stresses the need for
a more accurate method of efficiency measurement.

1.3 Description of Motor Losses


Losses in an induction motor can be segregated into no load losses and load
losses.
No-Load Losses:

Windage and friction are mechanical losses due to bearing friction and
windage. Core loss constitutes hysteresis and eddy current losses in the iron at
no-load.
Load Losses:

Stator ( ) losses are losses in the stator winding, Rotor ( ) or slip loss
are losses in the rotor winding. (Note: R is a variable with temperature.)
Stray-load losses are additional fundamental and high frequency losses in the
iron, strand and circulating current losses in the stator winding, harmonic losses
in the rotor conductors under load. These losses are proportional to the rotor
current squared [2].

4
As the rotor and stray losses are essentially zero at no-load, the no-load

watts are equal to friction losses, core losses and stator losses. As and and
no-load watts can each be measured, these losses can be separated. The windage
and friction can be separated by taking several tests at reduced voltage so that a
curve can be plotted against voltage squared. If the data is taken from
approximately one third voltage to the point where the current reaches a
minimum value, then the plotted curve should be a straight line and the
intersection with the zero voltage axis will give the windage and friction loss.
For low-slip motors the windage and friction can be considered constant and
the no-load value used. The core loss is a function of the internal or secondary
voltage, although this refinement is included only when equivalent circuit
parameters are available to determine the secondary voltage as a function of
load. When equivalent circuit parameters are not available, the core loss is
assumed constant and the no-load value used. Both of these assumptions are
conservative in that they result in a slight decrease in tested efficiency.
The stator loss is determined by measuring the dc resistance of the
winding at a known temperature, adjusting the resistance for temperature, and
multiplying by either a tested or calculated primary current squared. The rotor

loss is determined by measuring the slip and input watts at the load point
of interest. The formula (input power- stator loss - core loss) times slip in per-
unit (adjusted for temperature as required) yields the rotor losses. Where it is
impractical or impossible to run the machine under load, the rotor losses can be

determined by calculation. In these cases, is determined from a locked-rotor

impedance test, and is determined by equivalent circuit or circle diagram


techniques [2].

5
A stray-load loss is the most difficult loss to measure and perhaps one of the
most variable losses between motors of identical designs. There are direct and
indirect methods of determining stray-load losses. In the indirect test methods,
the stray-load loss is determined as the leftover losses (test loss minus
conventional loss). Test losses are input minus output. Conventional losses are
the sum of stator losses, rotor losses, friction losses and core losses. The stray-
load losses measurement can be improved by forcing it to fit the equation stray

loss equals K( ).

1.4 Synthetic loading


The most common currently used method of carrying out full-
load performance testing of a three-phase induction motor is to
apply full-load torque to the machine's output shaft. To load a
large machine, equally large test equipment or a duplicate
machine is required. The cost of setting up such a test facility,
maintaining the equipment, and the time and setting up
procedures for mechanically coupling the load machine, may
make full load tests prohibitively expensive. Large vertically
mounted machines are extremely difficult to test by applying a
load to the shaft because of the difficulty in finding a suitable
vertical load.
In this thesis a novel method will be described for rapid full load efficiency
evaluation of three phase induction motors using a synthetic loading technique,
without the need to connect a load to the machines drive shaft. The method
proposed considerably reduces the testing time compared with conventional
methods of efficiency measurement and the accuracy of the result is maintained.
The essence of synthetic loading is that while the motor is
running at no-load, the rotor is then oscillated in a motoring-

6
generating cycle such that, on average, over one cycle of the
synthetic loading full load current is drawn from the supply. The
machine under this condition will be producing rated copper
losses. If the average applied voltage over one loading cycle is
also the rated voltage, then the rated iron loss will be present.
Since the machine is also running at close to its operating
speed, then the rated friction and windage losses will be
present. The test is also performed at rated temperature
(synthetic loading was originally designed to produce full load
temperature rise characteristics).
The synthetic loading technique has began first by applying
two voltages of different frequency to the stator winding causes
heating due to alternating motor - generator operation. After that the researcher
found a new way to make the machine alternating between motor-generator
operation, and these methods is using the power electronic device. Sweep
frequency method, dual frequency method and constant speed of rotating
magnetic field (CSORMF) method. But the new method is using the current
control to load the machine [4].

1.5 Thesis Outlines


The thesis is organised into five chapters covering analysis and simulation
aspects of synthetic loading as a method of efficiency evaluation for induction
machines.
Chapter two presents background theory and a literature review of efficiency
evaluation of electrical machines. The literature in the areas of machine power
loss and efficiency, the standard efficiency test and the synthetic loading
technique of electrical machines are reviewed. Chapter three develops detailed
mathematical equations for the induction machines in d- and q- axis equivalent
circuits.

7
Chapter four presents simulation results of the synthetic loading technique and
the standard efficiency test for the induction machine under four different
synthetic loading frequencies. The individual losses from the simulation result
of the standard efficiency test are compared with the losses developed during
the simulation of synthetic loading. Chapter five draws general conclusions and
provides suggestions for further research work in this area.

8
Chapter Two
Literature Review and Background Study

2.1 Introduction
In this chapter details relevant background and a literature review of
efficiency evaluation for electrical machines are provided. The literature review
covers research related to power losses, efficiency, standard efficiency
evaluation methods, and synthetic loading for efficiency evaluation for
electrical machines.

2.2 Standard Efficiency Test Methods


Determination of temperature rise as well as power dissipated as heat inside
the electrical machines is a matter of interest for both consumers and
manufacturers. Heat generation affects the insulation materials and the cooling
systems of the machine. In this section, the established methods of full-load
testing for efficiency evaluation and temperature rise tests for induction
machines are reviewed.
Paul, et al. [2] reported that IEEE Standard 112-1977 is a revision of
IEEE Standard 112A-1964 and was prepared by the 112 Working Group of the
IEEE Rotating Machinery Committee. Following is a summary of some general
changes as compared to the 1964 edition that relate to efficiency determination.
The test procedure for the standard method is more specific relative to the

temperature correction for losses and slip for all the various calculation
procedures. For final performance evaluation, the following paragraph has been
added: "When the rated load temperature rise has not been measured, the
resistance of the winding shall be corrected to the temperature. This reference

temperature shall be used for determining losses at all loads. If the rated

8
temperature rise is specified as that of a lower class of insulation system, the
temperature for resistance correction shall be that of the lower insulation class."
The separation of core loss from the no-load losses is now mandatory rather
than optional. Concurrent with this change, the core loss is included in the
separation of stray-load loss under Method C and is added as another branch to
the equivalent circuit in Method F.
IEEE Standard 112, in addition to other types of tests, currently includes five
methods for the determination of motor efficiency. These may be divided into
two parts:
Direct Measurement of output
Method A (brake): normally associated with fractional HP
machines.
Method B (dynamometer)
Method C (duplicate machines)
Determination of Losses Without Output Measurement
Method E (input measurements)
Method F (equivalent circuit)
Grantham, et al. [3] shows the efficiency test standards available for
induction machines. The IEEE Rotating Machinery Committee has produced a
standard IEEE 112-1991 by which efficiency is tested in the United States. In
addition the International Electro-technical Commission (IEC) produced the
IEC-34-2 standard for use in Europe, which is similar to British Standard
BS269, and the Japanese Electro-technical Committee (JEC) produced standard
JEC-37. A comparison shows that these methods are similar differing only in
the details of procedure to be followed in each test [4]. In addition, the IEEE
112 test method B is the most rigorous and more detailed than those that
measure all the motors losses using a dynamometer such as specified in the IEC
or JEC standard.

9
Ghai, [5] shows the Electrical Machinery Committee (EMC) of the Power
Engineering Society (PES) of the Institute of Electrical and Electronics
Engineers (IEEE) has established a comparison of U.S. standards with
International Standards. It compares IEC 34 series of standards with National
Electrical Manufacturers Association (NEMA) standards for large induction
motors for general purpose applications. It concludes that a substantial amount
of correspondence exists between NEMA and IEC standards. NEMAs recent
practice of using IEC standards as the starting point for writing additional
standards has increased the degree of correspondence between NEMA and IEC
standards. Furthermore, requirements for any specific design parameter are not
always the same between NEMA and IEC, though the intent is identical. The
inclusion of tolerances on induction machine performance, and the lower stray
load loss allowance in IEC, causes a given design to have better indicated
performance per IEC 34 than for NEMA.
Boglietti, et al. [6, 7] reports motor efficiency measurements
calculated in accordance with international standards. The
standard efficiency tests IEEE 112-B, IEC 34-2 and JEC 37
recommend different measurement procedures, in particular
for the stray loss determination and the temperature
corrections of the copper losses. In these papers, a comparison
of the measurement procedures defined by these international
standards has been reported, together with some comments on
the prescribed methodologies. Furthermore, a comparison is
made based on experimental results obtained by tests on four
general purposes, three phase induction motors. However, the
stray load loss measurement is a critical key step for the
accurate evaluation of induction motor efficiency. Therefore, a
critical analysis of stray losses has been performed. The stray
load loss sensitivity to measurement errors has been analyzed,
10
in order to understand which are the most critical quantities
that influence their evaluation.
Alan K. Wallace, et al. [8] shows the IEEE 112 B as a laboratory
environment testing method, designed to have a maximum of accuracy and
repeatability. According to [8], a well calibrated laboratory, with a well
performed IEEE 112 -B may reach overall accuracies within 1%. As with any
loss segregation method, assumptions have to be made when separating the
different losses. These assumptions are an integral part of the testing standard,
ensuring portability. Since a laboratory is, by definition, a measurement
environment, it permits both, a complex instrumentation setup and also the
performance of tests at operating conditions other than the restricted options
available in an industrial environment. The method IEEE 112 B has been
designed to characterize the capability of the motor under test, at rated voltage
conditions and as a function of load for steady state operation. The standard
describes acceptable unbalance conditions, and acceptable levels of harmonic
distortions. These specifications demand a clean voltage supply, which
translates into both, a best case and also a repeatable scenario. In order to
achieve repeatability, all the calculations must be corrected to a base line
ambient temperature, which has been chosen to be 40C.

2.3 Induction Motor Field Efficiency Evaluation Methods


Motor-driven systems use two-thirds of the total electricity consumed by
industry. Historically, energy efficiency improvements in these systems have
been important for economic reasons only. However, these improvements have
now assumed an environmental role in meeting the U.S. commitment to reduce
greenhouse emissions. In general, an energy efficiency improvement program
includes development of a motor management plan that focuses on development
of a plant motor inventory and an evaluation of motor performance for large or

11
critical motors. The evaluation of motors focuses on the operating efficiency
and motor load to identify energy efficiency gains and possible reliability
improvements. This requires a reliable method for assessing motor performance
in the field.
John S. Hsu, et al. [9] shows that the majority of motors in the field are
induction motors for which IEEE Standard 112 [10] would be applicable.
However, field evaluation of operating efficiency introduces an environment for
which IEEE Standard 112 is not applicable. For example, IEEE Standard 112
requires that induction motor tests be performed with a voltage unbalance not
exceeding 0.5% (note that this is significantly smaller than the NEMA MG-1
permissible limit of 1% [11] for successful operation of motors). However, field
conditions may exceed this limit by a significant measure. Thus, when
evaluating motor performance in the field, it is important to use a technique that
can accommodate field conditions and yield results of sufficient accuracy for
the evaluation needs.

There are many methods pertinent to field efficiency evaluation in the


literature, and new methods are appearing every year. The reader is encouraged
to refer to a survey, Assessment of methods for estimating motor efficiency
and load under field conditions by Kueck et al. [12] for a rather complete list
of references of efficiency estimation available, either commercially or in the
literature. This survey was prepared for the U.S. Department of Energy,
Bonneville Power Administration, and the Pacific Gas and Electric Company
and is available from the Bonneville Power Administration Printing Office.

A field evaluation method can consist of a single basic method or can be built
using a combination of different basic methods. This may help field engineers
select or establish an efficiency evaluation method suited to their need. The
basic methods are as follows:
nameplate method;
12
slip method;
current method;
statistical method;
equivalent circuit method;
segregated loss method;
air-gap torque method;
Shaft torque method.
All methods calculate efficiency according to the definition of:

The shaft output power is the input power minus the losses. How to
assess losses and evaluate output power gives rise to fundamental differences
among the various methods. Consequently, the accuracies of methods are
different. The degree of intrusiveness of a field evaluation method is determined
by what data are required to be measured in the field and the difficulty of
performing the measurements. One or more of the following measurements may
be involved:
nameplate reading;
speed measured by opto-tachometer;
currents measured by clamp-on transducer;
voltages measurement;
input power measurement;
stator winding resistance reading;
winding temperature data;
no-load data measured with uncoupled shaft;
Shaft torque measurement.

2.4 Physical Natures of Basic Methods


13
In this section, the physical basis of some basic method for induction motor
field efficiency evaluation methods is described shortly in terms of how the
efficiency is obtained.

2.4.1 Nameplate Method


The least intrusive field evaluation method is to obtain motor information from
the nameplate. In this method, it is assumed that the efficiency of the motor is
constant and equal to the nameplate value. This works best when the efficiency
load curve is fairly flat, so that the full-load efficiency is applicable for most
load conditions. The typical load factor of industrial motors is around 75% [13].
Using typical efficiency-versus-load curves for motors having various poles and
horsepower ratings, it can evaluate the potential accuracy of the nameplate
method. Figure 2.1 the efficiency is not a strong function of load for a two-pole
motor between 50% and 100% of load. However, the efficiency of an eight-
pole, 1-HP motor shows a marked decline over that load range. Hence, the
nameplate method may be applicable for some motors, but could result in
substantial inaccuracies for other motor types [13].

14
Figure 2.1 Typical efficiency-versus-load
With this nameplate method, three additional problems may occur. First, the
nameplate data may be given according to a method other than IEEE Standard
112 Method B. Second, the motor may have been rewound. Third, the field
environment pertinent to the voltage unbalance and harmonics content may be
different from that for which the nameplate data is derived.
Nameplate efficiencies of a given motor can be evaluated according to
different standards. The three most frequently used standards are the National
Electrical Manufacturers Association (NEMA) that uses IEEE Standard 112, the
Japanese Electrotechnical Committee (JEC), and the International
Electrotechnical Commission (IEC). These three standards are not in agreement
[14], [2] and may result in a given motor being stamped with rather different
efficiencies. A typical example given in [15] illustrates the confusing
international nameplate data situation. Rewound motors introduce additional
uncertainty, since the nameplate data may no longer be valid. Core loss of a
rewound motor may or may not be increased, depending upon the lamination
insulation and the cleaning process of the stator. The copper loss depends on the
new coil extension and wire gauges. Certain engineers suggest that, after each
rewinding to the same horsepower and same number of poles, a two percentage
points reduction of efficiency should be considered. However, a different
opinion indicates that the efficiency should not be reduced if the rewinding
follows Electrical Apparatus Service Association (EASA) standards.
The field environment pertinent to the voltage unbalance and harmonics
content is commonly worse than that from which the nameplate data are
derived. The worst situation for a field efficiency evaluation using the
nameplate method is that having a less than 10-HP low speed rewound motor
that was not repaired according to EASA Standards. The motor has data
stamped on the nameplate that is not given according to IEEE Standard 112
Method B and is operated under a polluted supply (voltage unbalance or

15
harmonic distortion). The efficiency can easily be off ten percentage points
from the nameplate efficiency. However, the bottom line is that a nameplate
method is better than no field evaluation at all.

2.4.2 Slip Method


This method presumes that the percentage of load is closely proportional to
the percentage of the ratio of measured slip to full-load slip. The shaft output
power is, thus, approximated using the following relationship:

Where, slip is a function of motor speed given by the ratio between the
difference of synchronous and rotor speed to the synchronous speed. Motor
speed can be measured by an optical tachometer, which has a low intrusion
level. Input power must also be measured, which has a higher degree of
intrusion. The slip method can be an improvement over a pure nameplate
method, especially when the motor efficiency versus load curve is not flat. Any
method that uses slip to estimate percentage of load is related to the slip
method. Once the shaft output power is known, one may use a typical
efficiency-versus-load curve specifically for standard-, high-, or in-between-
efficiency motors. Alternatively, one may combine the basic slip method with
other additional measurements, such as the input power, which has a higher
degree of intrusion, to obtain efficiency through (2.1).
NEMA MG-1 Section 12.46 [11] states that variation from the nameplate
speed of ac integral horsepower induction motors shall not exceed 20% of the
difference between synchronous speed and rated speed when measured at rated
voltage, frequency, and load and with an ambient temperature of 25 C. This
means that the nameplate slip can be 20% inaccurate when the motor is
operating in the field, thus introducing significant inaccuracies in (2.2). The no-
load speed of induction motors is always close to the synchronous speed.

16
Subsequently, the projection of a light load through the basic slip method is
relatively more accurate than the projection of a heavy load.

2.4.3 Current Method


This method presumes that the percentage of load is closely proportional to
the percentage of the ratio of measured current to full-load current. The shaft
output power is, thus, approximated using the following relationship:

Where, is the nameplate full load current and is the measured current. For
small integral horsepower motors the no-load current may not be greatly
reduced from the full-load current, the assumed load-versus-current curve used
by (2.3) is further apart from the actual curve at light loads. This is opposite to
what the basic slip method is relatively good for. The load is normally
overestimated. The expression of shaft output power defined by (2.4) requires

that the no-load current be known. This may increase the intrusiveness
substantially when a no-load test is required:

The average of the two approaches of (2.3) and (2.4) may give a more
accurate shaft output power. NEMA MG-1, Section 12.47, states that, when
operated at rated voltage, rated frequency, and rated horsepower output, the
input in amperes shall not vary from the nameplate value by more than 10%.
This is another source of errors for the current methods. Motor current
measured by a clamp-on probe corresponds to a relatively low level of
intrusiveness. The insulation of motor leads and terminals is not disturbed. The
measured current is used to estimate the load of a motor. The simple current

17
method does not require a no-load current value. Just as with the slip method, to
obtain efficiency one will have to either use typical efficiency-versus-load
curves or measure input power.

2.4.4 Statistical Method


Empirical equations are set up to use minimal numbers of measured data for
input power and efficiency estimations. Usually, application of this method is
restricted to the group of motors for which empirical equations were derived.
If the statistical results are not used for the group of motors that the empirical
equations are based on, significant errors in the efficiency estimation are likely.
The statistical results can be quite different for the same variable. A good
example is the stray load loss estimation. NEMA MG1 [11], paragraph 20.52,
states that, if stray load loss is not measured, the value of stray load loss at rated
load shall be assumed to be 1.2% of the rated output for motors rated less than
2500 hp and 0.9% for motors rated 2500 hp and greater. IEEE Standard 112
[10, Sec. 5.4.4], gives different assumed stray load loss values for motors rated
less than 2500 hp. They are as follows:
1) 1125 HP = 1.8%;
2) 126500 HP = 1.5%;
3) 5012499 HP = 1.2%.
Additionally, certain non-U.S. standards use 0.5% of rated output as the stray
load loss at rated load. The statistical approach is commonly used with other
basic methods. For instance, Ontario Hydro 1 has published a segregated loss
method that simplifies IEEE Standard 112 method E1 much further. As pointed
out in this study, it is not always possible to interrupt a process long enough to
decouple a motor from its load and conduct a no-load test. The study suggests
that one way around this obstacle is to assume a value for the combined
windage, friction, and core losses. This method suggests that these combined
18
losses be set to 3.5% of input rated power. The stray load losses are estimated
based on the IEEE Standard 112 standard assumed values.
This method can be simplified even further by using assumed values for rated
power factor. Approximations can also be made for the temperature rise of the
winding, and the winding resistance can even be estimated by measuring the
line-to-line resistance taken from the circuit breaker and subtracting the
estimated cable resistance. The only other measurements required are power
into the motor and motor speed.

2.4.5 Accuracies of Basic Methods


Although more rigid accuracy evaluation of field efficiency measurement is
needed, accuracy estimation of measured efficiency presently is based on
experience. For NEMA frame motors, the nameplate efficiency is commonly
stamped according to Method B of IEEE Standard 112. Bonnett [16] suggests
that with the use of existing technologies, it is unreasonable to expect accuracies
of efficiency for the following items better than the following:
accuracy of calculations: 0.5 points of efficiency;

manufacturing and material variations: 0.5 points of efficiency;

Test accuracy: 0.5 points of efficiency.

A total of 1.5 points of efficiency, in Bonnetts opinion, is the simple


summation of the three factors that he considered. It is not the square root of the
sum of accuracy squares, which is a commonly used method. For various basic
methods, the best accuracy is provided by the torque gauge method. It may have

an accuracy of 1%, which is in line with Bonnetts manufacturing and test


accuracies. The rationale is that we know the least accurate method is the

nameplate method; it has the worst accuracy of 10% for loads between half

19
and full. The best accuracy is provided by the torque gauge method. It has an

accuracy of 1%. All the other methods that is partially involved with either
nameplate data and/or statistical values fall in between these two extreme basic
methods. The lesser use of nameplate information and the greater reliance on
direct measurement without assumed values for efficiency evaluation, the better
the accuracy would be. The inaccuracies, in general, are worse below 50% load.
The statistical or empirical method can provide a wide range of accuracy. It
depends on the sample population and the application range. Intuition tells us
that a well-targeted sample range for a small application range normally gives
high accuracy. For example, empirical equations obtained from data of one
specific sample motor can be very accurate when they are applied to the same
specific motor. The accuracy becomes extremely poor when the one sample
conclusion is applied to a totally different motor.

2.5 Synthetic Loading Method


The essence of synthetic loading is that while the motor is running at no-load,
the rotor is then oscillated in a motoring-generating cycle such that, on average,
over one cycle of the synthetic loading full load rms current is drawn from the
supply. The machine under this condition will be producing rated copper losses.
If the average applied voltage over one loading cycle is also the rated rms
voltage, then the rated iron loss will be present. Since the machine is also
running at close to its operating speed, then the rated friction and windage
losses will be present. The test is also performed at rated temperature (synthetic
loading was originally designed to produce full load temperature rise
characteristics).
Each method of synthetic loading is designed to produce these conditions.
The assumption is that if these conditions are met, then the losses during
synthetic loading will be the same as the losses at the machines rated full load

20
operating condition. Therefore, if the input power can be measured accurately
during the loading cycle, then the total full load losses should be able to be
identified and hence, the efficiency of the motor evaluated.
Synthetic loading offers the advantage that the machine
under test is not coupled to a load, thereby negating the need
for a special test bed. Also, the machine could be tested on site
and is not restricted by the axis of mounting.
The only equipment required to do the test is an inverter rated
at the same power rating as the machine under test, current
and voltage sensors and a typical
PC fitted with a DSP card for measurement and control. The
costs associated with this test method would be significantly
reduced.
Soltani, et al. [17] reported that full-load testing of large induction
machines is constrained by the limitations in the power-supply and loading
equipment of the manufacturers facilities, resulting in costly set up time. A new
synthetic loading method is proposed based on a bang-bang phase control
strategy. The rated power oscillation created is routed to an auxiliary system
and the source hydro has to provide only the total losses of the system, without
seeing the excessive power swings observed in other synthetic loading
techniques. In this technique, only induction machines are used which would
enable motor manufacturers to build the test rig in-house. The control stage is
very simple to implement and requires only unregulated dc supplies for the
excitation windings. The method is suitable for any induction machine and does
not requires any set up time. It is possible to strictly maintain constant define
r.m.s voltage and current at rated values for the duration of the heat runs.
In 1921 Ytterberg introduced the two-frequency or mixed-frequency
synthetic loading method for insulation temperature rise test of induction
machines [4]. This method develops the equivalent of a full-load test but is
21
achieved without mechanically loading the induction machine. With this
method there are two possible connection techniques for the two-frequency
synthetic loading method. The two-frequency method connects the machine
under test to two different supplies connected in series. The main supply has
rated voltage and frequency Va and fa. The auxiliary supply has a voltage and
frequency Vb and fb which is lower than rated value and is adjusted until the
induction motor input voltage and the stator current equal the rated values.
The other connection technique for the two-frequency
method is to inject two voltages of two different frequencies
and amplitudes into the machine under test using an isolating
transformer [4]. The main voltage has rated amplitude and
frequency and the auxiliary voltage chosen to be slightly
different from the main voltage, normally less than it. The
auxiliary voltage is generated via a synchronous generator or
an inverter. The two supplies are at different frequencies and
are mixed using a specially designed isolating transformer.

2.5.1 Dual Frequency Method


Grantham, et al. [22] reported the essence of the (DF) method is to produce a
supply containing two distinct frequencies; this has the effect of producing two
fields rotating at different speeds. If all quantities are represented as per unit
values, for sinusoidal voltages the flux can be set equal to the supply voltage
divided by the frequency. With two voltages in series, the total flux is the sum
of two flux waves of different magnitude and frequency; the magnitude and
angular velocity given by Schwenk [19] are as follows:

(2.5)

(2.6)

Where:

22
These equations represent a flux wave that varies in magnitude and angular
velocity as a function of time, supply voltages and frequencies. From these
equations it can also be seen that the frequency of oscillation of the flux wave
angular velocity and the frequency of oscillation of the magnitude of the flux
wave are the same and equal to the difference of the main and auxiliary supply

frequencies, that is ( ). Because of the inertia of the rotor, the varying

frequency of the resultant flux wave causes alternating induction motor-


generating action, thus increasing the effective r.m.s current flowing in the
machine.
This method, which currently requires a system of electrical machines to
generate the two frequencies, can be very conveniently implemented using
microprocessor controlled power electronics to generate the necessary voltages
and frequencies for the two supplies.
The Dual-Frequency synthetic loading input voltage is:

2.5.2 Sweep Frequency Method


Another method of artificially increasing the motor current is to use a single
supply frequency, but to rapidly modulate this frequency over a small range
centred on the rated frequency. This causes alternating induction motor-
generating action. This method is termed the sweep frequency (SF) method.
Because it is not possible to produce the required change of frequency using
electrical machines, this method necessitates the use of microprocessor
controlled power electronics and the method is new. A pulse width modulated
(PWM) inverter is ideal for this purpose. The voltage equation for sweep
frequency is:

23
(2.8)

This voltage produces a flux wave which varies both in magnitude and angular
velocity.

2.5.3 Constant Speed of Rotating Magnetic


Field Method
M. Sheng, et al. [20] shows the constant speed (CSORMF) method is to
produce a rotating magnetic field whose speed is constant, while its magnitude
varies sinusoidally. This concept departs somewhat from the principles of
synthetic loading using the DF and SF methods, in which the speed of the
rotating magnetic field oscillates about the synchronous speed.
If it is assumed that the rotor winding is running at the same speed as the
rotating magnetic field, the current in the rotor winding would be zero, provided
that the magnitude of the rotating magnetic field is constant. This is because
there is no changing flux linking the rotor winding. Now, if the magnitude of
the rotating magnetic field varies sinusoidally with time, an e.m.f will be
induced and therefore a current will flow in the rotor winding due to the

changing flux magnitude seen by the rotor winding, i.e. 0.

The direction of the e.m.f is such as to oppose the change of the flux linkage
with the rotor winding. An oscillating torque will therefore be produced due to
the interaction of the rotor current and the resultant rotating magnetic field. This
oscillating torque will accelerate and slow down the rotor of the test machine,
causing the rotor speed to oscillate about the synchronous speed.
This method works in much the same way as a transformer in which the
primary and secondary windings, although stationary, are magnetically
coupled .by varying flux magnitude. The rotor rotates very close to synchronous
speed, thus ensuring that an equivalent cooling condition close to that for

24
conventional loading is achieved. The CSORMF synthetic loading input voltage
per-phase in a three-phase system is:

Where,
V = centre voltage amplitude in each phase,
= main frequency of 50Hz,

= sweep frequency in Hz,

= sweep magnitude in Hz, and

2.5.4 Current Control Method


Abbas, et al. [21] states the new method used in synthetic
loading is using the current control to mange the current in the
machine to give the rated losses. During this method the
equipment required to the efficiency test is a vector controller
(inverter, controller, appropriate voltage and current sensors,
and position sensor or estimator) and a power analyzer. The
cost and time associated with performing a synthetic loading
test is significantly reduced as the test equipment is portable.

2.6 Summary
This chapter shows the literature review and background study, it describe at
the first the standard efficiency test methods that used in the past and it is still
used. Then the induction motor field efficiency evaluation methods are
presented, and the description of some methods that the electrical engineer is

25
used to evaluate the induction machines efficiency is provided. At last the
synthetic loading technique is described.

Chapter Three
Modeling of Induction Motor

3.1 Introduction
The voltage equations that describe the performance of a three phase
induction machines have six equations, and in these equations it is found that
some of the machine inductances are function of the rotor speed, whereupon the
coefficients of the differential equations that describe the behavior of these
machines are time varying except when the rotor is stalled. Thus a change of
variables is often used to reduce the complexity of these differential equations.

26
This change (transformation) refers machine variables to a frame of reference
that rotates at an arbitrary angular velocity. All known real transformations are
obtained from this transformation by simply assigning the speed of the rotation
of the reference frame.

3.2 The Park Transformation Theory


The substitutions which replace the variables (currents, voltages, and flux
linkages) associated with the stator windings of a synchronous machine with
variables related to fictitious windings rotating with the rotor was first
investigated by Park .This method was further extended by Keyhani and Lipo to
the dynamic analysis of induction machines. By these methods, a machine
winding can be reduced to a set of two phase-windings with their magnetic axes
aligned in quadrature as shown in Figure 3.1. The d-q axis transformation
eliminates the mutual magnetic coupling of the phase-windings, thereby making
the magnetic flux linkage of one winding independent of the current in the other
winding. This system of transformation allows both machine windings in the
stator and rotor of an induction machine to be viewed from a common reference
frame, which may rotate at any angular speed or remain fixed to the stator.
Generally, the reference frame can also be considered to be rotating at any
arbitrary angular speed. The transformation from a three-phase system to a two-
phase system and with the zero-sequence included is shown in the next
equations.
The Park transformation matrix transforms three-phase components, , ,

and , which can be either currents or voltages, from the stationary reference

frame abc, into three-variable quantities, , , and , which is called arbitrary

reference frame dqo.


Park transformation equation

27
(3.1)

Inverse Park transformation equation

(3.2)

In the above equation, f can represent voltage, current, flux linkage or electric
charge.

Figure 3.1 The transformation of the abc stationary reference


frame to the dq arbitrary reference frame
3.3 Mathematical Model
A three phase machine with symmetrical windings is assumed. The three
phase abc model is then transformed to a two phase d-q model, which reduces
the complexity in the machine model.
Generally, low order mathematical models are used to represent the induction
machines in simulating and analyzing dynamic behaviour. Four nonlinear
differential equations, based on Parks d-q axis representation are used to

28
represent induction machines. These equations are arranged in a set of first
order differential equations which are derived from the d- and q-axis.
So the mathematical description of an ideal symmetrical squirrel-cage type
induction motor in an arbitrary reference frame is given by the following
matrix. The voltage equations of the induction machine in the d-q frame can be
expressed as:

q
D
d d

Figure 3.2 The d-q axis arbitrary reference frame

= (3.3)

The stator voltage refer to the d- axis of reference frame.

: The stator voltage refer to the q- axis of reference frame.

The stator currents refer to d- and q- axis of reference frame.

The rotor currents refer to d- and q- axis of reference frame.

The rotor voltage refer to the q- axis of reference frame.

The rotor voltage refer to the d- axis of reference frame.

29
: Resistance of stator winding. : Resistance of rotor conductor.

: Stator self inductance. : Rotor self inductance.

: Magnetizing inductance. : Electrical angle velocity of rotor.

: Angle velocity of arbitrary reference frame

P : Number of pole pairs

: Mechanical angle velocity of rotor

In this modelling the arbitrary reference frame fixed to the


stator, on the other hand the arbitrary reference frame rotate

with electrical angle velocity of rotor ( ). In this method of analysis,


it is assumed that the effect of saturation due to either the magnetizing
inductance or the leakage inductances is negligible. Using this assumption, the
values of the magnetizing inductance, stator leakage, and rotor leakage
inductances are constant and thus do not vary with the magnetizing current. Re-
arranging the matrix :

30
= + +

(3.5)

And the differential equations are:

(3.6)

31
=

(3.7)

= (3.9)

Where:
32

The electromagnetic torque, developed by the rotor is described by

(3.14)

(3.15)

The dynamic equation of the speed is

33
= (

(3.17)

Where, D is Frictional coefficient and J is rotor moment of inertia.

3.4 Calculation of the Losses


After we make the modeling of the induction motor, we will use the MATLAB
program to show the result for the stator copper losses, rotor copper losses and
friction losses. We will show the result in the next chapter.

(3.18)

(3.19)

(3.20)

The induction motor to be simulated undergoing synthetic


loading is an industrial three-phase, four poles, 50Hz, 7.5kW,
415V, delta connected induction motor, rated at 13.8A with
single cage rotor. The following constant parameters were
determined from the usual locked rotor and synchronous speed
tests:
Rs = 2 ; Rr = 2 ;
ls = lr = 6.37 mH;
Rm = 1000 ; Lm = 318 mH;

D = 0.0055 Nm.s/rad; J = 0.05 kg. ;

The dynamic model was used to get the performance of the machine, the four
differential equations and the dynamic equation are the main equations. The

inputs are . The input torque

34
for the conventional method will be the full load to get the full load current for
the machine, and when we used the synthetic loading the torque will be zero.

3.5 Description of the modeling


In this thesis the first method that used is the conventional method, was used
the machine data and their parameters that shown above and the input voltage
value:
(3.21)

(3.22)

And the torque is the full load value.


The second method used is the dual frequency method. Constant speed of
rotating magnetic field method is the third method used in this thesis. The three
methods is used and the results is shown in the next chapter, the comparison
was made to show how is synthetic loading technique is effective in efficiency
evaluation.

3.6 Summary
This chapter in the beginning shows the benefit of using the Park
transformation for the induction machine; this transformation reduces the
complexity in the machine model and that by making four equations instead of
six equations. Then the mathematical model in the arbitrary reference frame is
used to obtain the results. The losses calculated are the stator, rotor and friction
losses and their equations are shown above.

Chapter Four
Results and Discussion

35
4.1 Introduction
Simulation of the synthetic loading technique, with careful selection of the
input conditions is essential in order to evaluate the efficiency of the induction
machines experimentally. Therefore, simulation and computer modeling is used,
based on accurate system parameters, to evaluate the individual loss
components and the input power to the induction machine model. In this chapter
simulation results of the synthetic loading method for efficiency evaluation at
rated speed are presented for constant load torque conditions of the induction
machine. The synthetic loading and the standard efficiency tests, as methods of
efficiency evaluation, are simulated using MATLAB. As the aim of this
research is to determine the suitability of synthetic loading for efficiency
evaluation of induction machines, the individual losses from the simulation
results of the standard efficiency test are compared with the losses developed
during the simulation of the synthetic loading. The main losses in induction
machines are stator copper loss, rotor copper loss, iron loss, and friction and
windage loss.

4.2 Results
The nonlinear differential equations that describe the induction machines are
simulated using MATLAB and the efficiency evaluation of induction motor is
performed. The parameters of the machines are shown in the appendix A. The
machine is a 4-pole, 50-Hz, 3-phase induction motor. The parameters are
expressed in ohms using the 50Hz value of the reactance. The program that
used to evaluate the induction motor efficiency is provided in the appendix B.
The function m-file and the function ode23 is used to solve the differential
equations and the results are discussed.

36
4.3 Conventional Method Results
It is essential to understand that the synthetic loading method is an accurate
method to evaluate the efficiency of induction motor. Therefore, it is vital to
perform the conventional method to calculate the losses and the efficiency of
induction motor and then perform the synthetic loading methods.The
conventional method is used to fully load the three-phase induction machine has
the following rated 7.5kW, 13.9Amps and 415V. The input power and the total
loss of the induction machine at nominal operating conditions will be provided.
It is essential to consider the type of winding connection on that machine if
its delta or star because it will make some different in the calculations. Figure
4.1 shows the current waveforms during the full load condition. The three
phases peak current is 7.9Amps which confirms the machine is fully loaded.
The duration that taken is after a time to make the induction machine reach the
steady state. Figure 4.2 shows the maximum rated voltages waveform during
the full load condition. The peak line to line voltage is 586V.
Figure 4.3 shows the rotor speed. In the program the load torque applied to
the machine is the full load torque so the speed at the steady state is 1448r.p.m.
It can be noted that the rotor speed begins from zero and reach its rated speed.

10

5
Current (A)

-5

-10
1.9 1.91 1.92 1.93 1.94 1.95 1.96 1.97 1.98 1.99 2
Time (s)

Figure 4.1 The three phase currents of the induction motor under full load
torque at rated speed

37
600

400

200
Voltage (V)

-200

-400

-600
1.9 1.91 1.92 1.93 1.94 1.95 1.96 1.97 1.98 1.99 2
Time (s)
Figure 4.2 The three phase line voltages of the induction motor under full-load
torque at rated speed

X: 1.103
Y: 151.6

150
Rotor Speed (rad/s)

100

50

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2


Time (s)
Figure 4.3The rotor speed of induction motor under full-load torque

Figure 4.4 shows the stator copper losses. The induction motor at starting
draw a current about 5 times the full load current. Accordingly, the losses at
starting are much higher than the steady state. Therefore, the induction motor
needs a period of time to reach its steady state condition and that value of the
stator copper loss is about 384.4W. Figure 4.5 depicts the rotor copper losses.
The losses during starting period are varying until the motor reach the steady
state and it is about 280.2W. In the simulation of the induction machine the load
torque applied to the machine is the full load so the speed at the steady state is
the rated value and it is 151.6 rad/s. From equation (3.20) the friction and
windage loss is 126.4W.

38
1000

800
Losses (W)

600
X: 1.082
Y: 384.3
400

200

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Time (s)
Figure 4.4The stator copper losses of induction motor under full-load torque

1000

800
Losses (W)

600

400 X: 0.9421
Y: 280.3

200

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Time(s)

Figure 4.5The rotor copper losses of induction motor under full-load torque

150

X: 1.092
Y: 126.4

100
Losses (W)

50

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Time (s)

Figure 4.6The friction and losses of induction motor under full-load torque

The electrical torque can be expressed as:


(4.1)

39
the peak oscillating torque with the field oscillating frequency can be calculated,
standardized with respect to the rated value. Figure 4.7 shows the electrical
torque of the induction machine. Figure 4.8 illustrates the input power to the
induction machine and the sum of the losses that calculated. It can be noticed
that the input power have a maximum value in the starting and that is due to the
higher current that drawn and when the current reach the steady state value the
input power also reach the steady state value. Also the losses are the same; it is
higher in the starting period and reaches its rated value at steady state.
200

150
Torque (N.m)

100
X: 1.162
Y: 51.26

50

-50
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Time (s)
Figure 4.7The electrical torque of induction motor
4
x 10
8
Input Power & Total Losses (W)

Input Power
Total Losses
6

2 X: 1.129
Y: 8436
X: 1.298
Y: 791

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2


Time (s)
Figure 4.8The input power with the total losses

4.4 Synthetic Loading Technique Results


Based on the dynamic model of induction machine the synthetic loading
process can be modeled and evaluated for use in determining the efficiency of

40
an induction machine. The input voltages, and , are modified in each

method so as to produce the synthetic loading effect. It will be noted that only
one set of permutations has been presented for each method of synthetic
loading.
4.4.1 Dual Frequency Results
The Dual Frequency (DF) method, as the name suggests, relies upon the input
voltage being made up of two sinusoids per phase, each with its own frequency
and peak amplitude. The summation of these two sinusoids per phase is applied
to the machine undergoing synthetic loading. During DF synthetic loading, as
the motor is running, the rotor accelerates and decelerates over a cycle of the
beat frequency. The beat frequency is the difference between the two sinusoidal
frequencies applied per phase.
The main and auxiliary peak line to line voltage amplitudes and frequencies

in the dual frequency method are V and V and 50Hz and


35Hz, respectively. The resultant r.m.s line to line voltage and current are given
the rated speed of the rotor, rated stator copper losses, rotor copper losses and
friction losses, and this is for the dual frequency method and particular to this
set of machine parameters.
Figure 4.9 shows the three phase current waveforms during the synthetic
loading technique with a main frequency of 50Hz and auxiliary frequency of
35 Hz, and load torque is zero. Figure 4.10 shows the variation of the rated
maximum voltage during synthetic loading with a synthetic main frequency of
50Hz and auxiliary frequency of 35Hz. The phase voltage magnitude during
synthetic loading depends on, main and auxiliary voltages that used. Equation
(2.7) is used as the input voltage in DF method. To calculate the value of the
two voltages that used in the dual frequency the following equation can be used:

41
Any value for the voltage ( can be chosen and then (4.2) is used calculate the

voltage .

20

10
Current (A)

-10

-20
1.8 1.82 1.84 1.86 1.88 1.9 1.92 1.94 1.96 1.98 2
Time (s)
Figure 4.9 The three phase currents of the induction motor for (DF) method

500
Voltage (V)

-500

1.8 1.82 1.84 1.86 1.88 1.9 1.92 1.94 1.96 1.98 2
Time (s)
Figure 4.10 The maximum rated voltages of the induction motor (DF) method
Figure 4.11 shows the variation of speed versus time. It can be noted that
although the machine speed is actually varying with time, the average speed is
1500 rpm and is very close to the rated nominal speed of the machine. The
speed variation illustrated in Figure 4.11 clearly show that during synthetic
loading the machine accelerates as a motor and decelerates as an induction
generator. Consequently the machine can be fully loaded via the repeated
acceleration and deceleration and without the need to connect a mechanical load
to the machine's drive shaft.

42
180

Rotor Speed (rad/s) 170

160

150

140

130
1.5 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2
Time (s)
Figure 4.11The rotor speed of induction motor for (DF) method

Figure 4.12 shows the stator copper losses for induction machine when using
(DF) method. The time taken is from 1.5s to 2s and this period is lies in the
steady state condition. In synthetic loading technique the average is taken for
one cycle and this value must be the rated value. The average for the stator
copper loss is 385.43W. Figure 4.13 shows the variation of rotor copper loss
with time .The loss during synthetic loading is varying the average value for one
synthetic loading cycle is 295.26W. Figure 4.14 shows the friction loss when
DF method is used, the average for one synthetic loading cycle is 136.14W.
1500

1000
Losses (W)

500

0
1.5 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2
Time (s)
Figure 4.12The stator copper losses of induction motor for (DF) method

43
800

600
Losses (W)

400

200

0
1.5 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2
Time (s)
Figure 4.13The rotor copper losses of induction motor for (DF) method
170
160
150
Losses (W)

140
130
120
110
100
1.5 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2
Time (s)
Figure 4.14The friction losses of induction motor for (DF) method

Figure 4.15 shows the variation of electrical torque with respect to time.
Figure 4.15 clearly indicates that the induction machine is heavily loaded by
synthetic loading technique. The machine is acting as a motor for half a cycle of
torque variation and as a generator for the second half cycle. The average value
of torque over a cycle is negative value proportional to the rotational losses and
any other fixed mechanical torque effective on the rotor. This torque cycle is
exactly repeated in the settled down operating condition and can be used to
separate the mechanical losses from the total losses if required.
Figure 4.16 shows the power measured at the input of the motor during the
simulation and the sum of the power losses while undergoing synthetic loading.
The average power at the motor terminals is positive when the rotor is
accelerating and negative when the rotor is decelerating. The sum of power

44
losses is always positive for the input frequency combination shown (main and
auxiliary frequencies at 50Hz and 35Hz respectively).
The results of one such simulation for the DF method of synthetic loading are
shown in Figure 4.16. It can be noted that the peak input average power is less
than the full load rated output power for the machine being modeled, namely a
peak input average power of approximately 6kW, while the rated output power
is 7.5kW.
To confirm the synthetic loading as a technique for evaluating efficiency, the
measured power at the input terminals must equal the sum of the losses during
the synthetic loading. This is a difficult task to establish experimentally since
the segregated losses cannot be measured directly during synthetic loading.
Therefore, simulation and computer modelling is used, based on accurate
parameters to evaluate the individual losses as well as the power that can be
measured at the input of the induction machine.
100

50
Torque (N.m)

-50

-100
1.5 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2
Time (s)
Figure 4.15The electrical torque of induction motor for (DF) method

4
x 10
1 Input Power
Input Power & Total Loss (W)

Total Losses

0.5

-0.5

-1
1.5 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2
Time (s)

45
Figure 4.16The input power with the total losses for (DF) method

4.4.2 Constant Speed of Rotating Magnetic Field Results


The constant speed of rotating magnetic field (CSORMF) method of
synthetic loading produces a magnetic field that is rotating at a constant speed.
If the rotor is rotating at the same speed as the magnetic field, no voltage or
current is induced in the rotor because there is no changing flux in the rotor
winding. However, if the magnitude of the flux wave is varying with time, then
an e.m.f will be induced in the rotor conductors and current will flow. The e.m.f
will be established such that it opposes the change of flux in the rotor winding.
The interaction of the induced rotor current and the rotating magnetic field will
produce a torque that causes the rotor to rotate. Since the excitation produces a
magnetic field of constant speed and oscillating magnitude, then the induced
current will also be oscillating, hence producing an oscillating torque, resulting
in synthetic loading of the induction machine.
The frequency is the same in both the primary and secondary windings. In the
same way, the amplitude of the applied per-phase voltage waveform is varied
during the CSORMF method of synthetic loading to produce a stator magnetic

field rotating at a constant speed ( ) with alternating amplitude. The amplitude

of the stator flux is varied according to the sweep magnitude, , and at a

sweep frequency of . The interaction between the stators constant speed


magnetic field and the induced rotor magnetic field produces torque. The torque
reaction causes the rotor to periodically accelerate and decelerate and the
machine becomes synthetically loaded. The modelled CSORMF synthetic
loading results for a main frequency of 50Hz with sweep frequency of 20Hz and
sweep magnitude of 5.5Hz. Figure 4.17 shows the three phase current

46
waveforms during the synthetic loading technique with a main frequency of
50Hz and sweep frequency of 20Hz, and sweep magnitude of 5.5Hz and the
load torque is zero.
Figure 4.18 shows the variation of the maximum rated voltage during
synthetic loading with a synthetic main frequency of 50Hz and sweep frequency
of 20Hz, and sweep magnitude of 5.5Hz and the load torque is zero. The phase
voltage magnitude during synthetic loading depends on, main voltages that used
main frequency, sweep frequency and sweep magnitude. Equation (2.9) is used
as the input voltage in (CSORMF) method.
15

10

5
Current (A)

-5

-10

-15
1.7 1.75 1.8 1.85 1.9 1.95 2
Time (s)
Figure 4.17 The three phase currents of the induction motor for (CSORMF)

500
Voltage (V)

-500

1.8 1.82 1.84 1.86 1.88 1.9 1.92 1.94 1.96 1.98 2
Time (s)
Figure 4.18 The maximum rated voltage of the induction motor for (CSORMF)

Figure 4.19 shows the variation of speed versus time. It can be note that
although the machine speed is actually varying with time; the average speed of
156.6 rad/sec (1495 rpm) is very close to the rated nominal speed of the
machine. The speed variation illustrates clearly that during synthetic loading the

47
machine accelerates as a motor and decelerates as an induction generator.
Consequently the machine can be fully loaded via the repeated acceleration and
deceleration and without the need to connect a mechanical load to the machine's
drive shaft.
180

170
Rotor Speed (rad / s)

160

150

140

130
1.5 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2
Time (s)

Figure 4.19The rotor speed of induction motor for (CSORMF) method

Figure 4.20 shows the stator copper losses for induction machine when
CSORMF method is used. The time taken to calculate the average is between
1.5s to 2s. This period is lies in the steady state condition. In synthetic loading
technique the average taken for one synthetic loading cycle and this value must
be the rated value. The average for the stator copper loss is 381.18W. Figure
4.20 shows the variation of rotor copper loss with time .The loss during
synthetic loading is varying and the average value for one synthetic loading
cycle is 283.16W. Figure 4.22 shows the friction loss when CSORMF method is
used, the average for one synthetic loading cycle is 134.9W.

48
1200

1000

800
Losses (W)

600

400

200

0
1.5 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2
Time (s)
Figure 4.20The stator copper losses of induction motor for (CSORMF) method
600

500

400
Losses (W)

300

200

100

0
1.5 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2
Time (s)
Figure 4.21The rotor copper losses of induction motor for (CSORMF) method
170
160
150
Losses (W)

140
130
120
110
100
1.5 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2
Time (s)
Figure 4.22The friction and losses of induction motor for (CSORMF) method

Figure 4.23 shows the variation of electrical torque with respect to time. It
clearly indicates that the machine is heavily loaded by synthetic loading
technique, the machine acting as a motor for half a cycle of torque variation and
as a generator for the other half cycle. Figure 4.24 shows the average power
measured at the input of the induction motor and the sum of average power
losses while undergoing synthetic loading. The average power at the motor
49
terminals is positive when the rotor is accelerating and negative when the rotor
is decelerating. The sum of average power losses is always positive for the input
frequency combination.
50
Torque (N.m)

-50
1.5 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2
Time (s)
Figure 4.23The electrical torque of induction motor for (CSORMF) method
Input Power & Total Loss (W)

Input Power
6000
Total Losses
4000
2000
0
-2000
-4000
-6000
1.5 1.55 1.6 1.65 1.7 1.75 1.8 1.85 1.9 1.95 2
Time (s)
Figure 4.24 The input power with the total losses for (CSORMF) method

4.5 Discussions
The synthetic loading technique gives accurate results for efficiency
evaluation of induction motor. The result obtained using conventional method is
compared with that obtained using synthetic loading methods. The comparison
shows that the conventional method results are in consistent with the synthetic
loading results. The efficiency for the Dual-Frequency and CSORMF methods
are calculated from the nameplate output power (7.5kW) divided by the
nameplate output po-wer plus the power losses determined during synthetic
loading (7.5kW + total losses). The efficiency is determined in this way since it
may not be possible to measure either the input or output power at rated load

50
on-site. If the rated steady-state input power is known, then the efficiency is
calculated using the input power minus the power losses divided by the input
power.

Consequently the efficiency for the conventional method:

And the efficiency using the synthetic loading technique for DF method:

And the efficiency for the synthetic loading technique for CSORMF method:

It can note that the results are very close to each other, there is a small
difference between the power losses but it does not affect result. Therefore, it
can be concluded that the synthetic loading technique is a best method that can
be used for efficiency evaluation of induction motor.

Table 4.1 Steady-state performance with Dual-Frequency synthetic loading


method for efficiency evaluation

51
Synthetic Loading; Voltage & Frequency (Hz)
Standard = 413.2 = 410 = 400 = 390
Efficiency = 38.14 = 64.22 = 110.56 = 141.8
Test = 35 = 39.1 = 42.7 = 44.1
Rms line voltage
415 414.95 414.99 414.98 414.97
(V)
Rms line current
13.83 13.88 13.85 13.85 13.87
(A)
Speed (rpm) 1448 1500.8 1495.7 1495.1 1490.3
Output power(W) 7500 _ _ _ _

Input power(W) 8436 816 812.4 812.3 807.3


Stator copper loss
384.3 385.4 384.01 384.1 384.8
(W)
Rotor copper loss
280.3 295.2 293.28 293.13 288.3
(W)
Friction and
126.4 136.1 135.18 135.1 134.2
windage loss (W)
Total losses 791 816.7 812.4 812.33 807.3
Efficiency % 90.4 90.17 90.22 90.22 90.28

Four synthetic loading frequencies that establish full-load torque conditions


are summarized in Table I for dual frequency method. In each case, we change

, , and . is constant and equal 50 Hz. Efficiency is calculated using the

rated output power divided by the output power plus total losses measured using
synthetic loading. The stator copper loss, rotor copper loss, friction and windage
loss are the same approximately as the standard efficiency test method. This is
expected because the stator rms currents, the rms voltage and the speed are at
rated values. The synthetic loading technique produces efficiencies between
90.17% and 90.28% when the rated output power of the induction motor is used
to calculate the efficiency. Therefore, synthetic loading underestimates the
efficiency by less than 0.12% at best, 0.23% at worst, and 0.175% on average.
The simulation results show that the losses, rms current, and average speed

52
using the synthetic loading technique are consistent with the standard efficiency
test method, which indicates that synthetic loading could be used for the
evaluation efficiency of the induction motor.

Table 4.2 Steady-state performance with (CSORMF) synthetic loading


method for efficiency evaluation
Synthetic Loading; Frequency (Hz)
Standard
Fm = 15 Fm = 18 Fm = 19 fm = 20
Efficiency
Sm = 8.7 Sm = 6.75 Sm = 6.15 Sm = 5.5
Test
Rms line voltage
415 415 415 415 415
(V)
Rms line current
13.83 13.9 13.86 13.85 13.81
(A)
Speed (rpm) 1448 1497.13 1496.66 1496.18 1495.3

Output power(W) 7500 _ _ _ _

Input power (W) 8436 827.6 807 804.9 799.7


Stator copper loss
384.3 395.1 384.6 383.7 381.2
(W)
Rotor copper loss
280.3 297.3 287.3 286.1 283.6
(W)
Friction and
126.4 135.2 135.1 135.1 134.9
windage loss (W)
Total loss 791 827.6 807 804.9 799.7

Efficiency % 90.4 90.06 90.28 90.3 90.36

Four synthetic loading frequencies that establish full-load torque


conditions are summarized in Table 4.2 for (CSORMF) method. In the
(CSORMF) method the main voltage and main frequency (equal 50 Hz) are

constant, we have to choose the correct value for and to get the rated

value for the current, voltage and average speed. Efficiency is calculated using
the rated output power divided by the output power plus total losses measured

53
using synthetic loading. As we said the stator copper loss, rotor copper loss,
friction and windage loss are the same approximately as the standard efficiency
test method. The synthetic loading technique produces efficiencies between
90.06% and 90.36% when the output power of the standard efficiency test is
used to calculate the efficiency. Therefore, synthetic loading underestimates the
efficiency by less than 0.04% at best, 0.34% at worst, and 0.19% on average.
The results in table 2 shown that the values for the (CSORMF) method was
found in good agreement for each results.

4.6 Summary
In this chapter the results for the conventional efficiency test methodis
obtained. The three phase currrents, voltages, speed, stator losses, rotor losses,
friction losses, torque and the input power are presented. Then the synthetic
loading technique is performed using the dual frequency and constant speed of
rotating magnetic field methods. The three phase currrents, voltages, speed,
stator losses, rotor losses, friction losses, torque and the input power under
synthetic loading are presented. The induction machine efficiency is calculated
using the total losses and the input power. Synthetic loading simulation results
give excellent agreement with the standard efficiency test method.

Chapter Five
Conclusions and Recommendations
54
5.1 Conclusions
Synthetic loading as a method for evaluating the efficiency of three-phase
induction motors has been confirmed, using computer modeling and simulation
techniques. It is accurate and able to identify the total losses in the machine
under test. The total losses supplied during synthetic loading were shown to be
comparable to the sum of the individual losses of the induction machine being
modeled. As a consequence, when the technique is applied to the real machine,
there can be confidence in the efficiency evaluation result. Previously, there was
uncertainty about the performance of synthetic loading for temperature-rise tests
and efficiency evaluation. However, with accurate computer modeling based on
real parameters, the synthetic loading method has been shown to be a valuable
technique for efficiency evaluation. The new method will enable motors to be
tested virtually at any location, including on site, and with any mounting
arrangement, horizontal or vertical. Efficiency can be established easily and
quickly at normal operating temperature. The required equipment will be easy
to set up, requiring the connection of a three-phase inverter output to the
machines terminals. Synthetic loading requires no external load to be
connected to the test machines drive and thus reducing set-up time and costs.
The equipment will draw only sufficient average power from the mains to
supply its own losses and the losses of the test machine, this total power being
small compared with the full load power of the machine under test, thus
reducing the overall power consumption. One piece of test equipment can
replace a number of large heavy and expensive electrical machines, for the
testing of a complete range of machine frame sizes.
Microprocessor controlled power electronic techniques can be used to
produce the two supply frequencies of the existing dual frequency synthetic
load method. The power electronic method obviates the need for separate MG

55
sets and an isolating transformer. An alternative method of achieving synthetic
loading, the constant speed of rotating magnetic field method, is to directly vary
the supply frequency using microprocessor controlled power electronics.
Measured temperature rises when using both new methods are in good
agreement with those using the conventional load method with a 50 Hz inverter
supply.

5.2 Recommendations
The synthetic loading technique in this thesis is used for an induction
machine, in calculation of the efficiency the iron losses has been neglected so
in more research we have to put the iron losses in consideration. The technique
is used for a small machine which the rated output power is 7.5 kw, it can be
used to evaluate the efficiency of higher rated machine output power.
The microprocessor control power electronic is the basic equipment that can
produce the two supply frequency for the dual frequency method and also can
give the variation of the supply frequency for (CSORMF). Accordingly, to
make synthetic loading technique a useful method for efficiency evaluation
experimentally the inverter should be used to drive the induction machine. With
hardly any additional cost, standard commercial speed control inverters can be
modified to offer the extra features of the new test methods.

Appendix A

56
Induction Machine Data

Rated output power 7500 Watt


Rated voltage 415 V
Rated current 13.9 A
Frequency 50Hz
stator resistance, Rs 2
Rotor resistance, Rr 2
Stator leakage inductance, ls 0.00637 H
rotor leakage inductance, lr 0.00637 H
Magnetizing inductance, Lm 0.318 H
Rotor moment of inertia, J 0.05 kg.
Frictional coefficient, D 0.005 Nm.s/rad
Number of poles 4

Appendix B

57
B.1- Program One

function xprime=indmatrix(t,x)
Rs=2; Rr=2;
LM=0.318; Ls=(0.00637)+LM;
D=0.005; Lr=(0.00637)+LM;
F=50; W=2*pi*F;
J=0.05; P=2;
Vd=0; Vq=0;
VM=sqrt(2)*415;
VD=VM*sin(W*t);
VQ=-VM*cos(W*t);
TL=50.5;
B=[Ls LM 0 0;LM Lr 0 0;0 0 Ls LM;0 0 LM Lr];
C=inv(B);
y1=VD-Rs*x(1);
y2=Vd-Rr*x(2)-P*x(5)*(LM*x(3)+Lr*x(4));
y3=VQ-Rs*x(3);
y4=Vq-Rr*x(4)+P*x(5)*(LM*x(1)+Lr*x(2));
y=[y1;y2;y3;y4];
px=C*y;
piD=px(1); pid=px(2);
piQ=px(3);
piq=px(4);
Te=(3/2)*P*(x(2)*LM*x(3)-x(4)*LM*x(1));
pwr=(Te-D*x(5)-TL)/J;
xprime=[piD;pid;piQ;piq;pwr];

B.2 -Program Two


58
function xprime=dual(t,x)
Rs=2; Rr=2;
P=2; LM=0.318;
Ls=(0.00637)+LM; Lr=(0.00637)+LM;
D=0.005; J=0.05;
Va=sqrt(2)*413.2; Vb=sqrt(2)*38.14;
wa =2*pi*50; wb=2*pi*35;
Vd=0; Vq=0;
VD=Va*sin(wa*t)+Vb*sin(wb*t);
VQ=-Va*cos(wa*t)-Vb*cos(wb*t);
TL=0;
B=[Ls LM 0 0;LM Lr 0 0;0 0 Ls LM;0 0 LM Lr];
C=inv(B);
y1=VD-Rs*x(1);
y2=Vd-Rr*x(2)-P*x(5)*(LM*x(3)+Lr*x(4));
y3=VQ-Rs*x(3);
y4=Vq-Rr*x(4)+P*x(5)*(LM*x(1)+Lr*x(2));
y=[y1;y2;y3;y4];
px=C*y;
piD=px(1);
pid=px(2);
piQ=px(3);
piq=px(4);
Te=(3/2)*P*(x(2)*LM*x(3)-x(4)*LM*x(1));
pwr=(Te-D*x(5)-TL)/J;
xprime=[piD;pid;piQ;piq;pwr];

B.3- Program Three

59
function xprime=csormf(t,x)
Rs=2; Rr=2;
P=2; LM=0.318;
Ls=(0.00637)+LM; Lr=(0.00637)+LM;
F=50; W=2*pi*F;
D=0.005; J=0.05;
VM=sqrt(2)*415;
Vd=0; Vq=0;
Fm=20; Sm=5.5;
VD=VM*(sin(2*pi*F*t)+(Sm/(2*F*F))*(F-Fm)*sin((2*pi)*(F-Fm)*t-pi)+(Sm/
(2*F*F))*(F+Fm)*sin((2*pi)*(F+Fm)*t-pi));
VQ=-VM*(cos(2*pi*F*t)+(Sm/(2*F*F))*(F-Fm)*cos((2*pi)*(F-Fm)*t-pi)+
(Sm/(2*F*F))*(F+Fm)*cos((2*pi)*(F+Fm)*t-pi));
TL=0;
B=[Ls LM 0 0;LM Lr 0 0;0 0 Ls LM;0 0 LM Lr];
C=inv(B);
y1=VD-Rs*x(1);
y2=Vd-Rr*x(2)-P*x(5)*(LM*x(3)+Lr*x(4));
y3=VQ-Rs*x(3);
y4=Vq-Rr*x(4)+P*x(5)*(LM*x(1)+Lr*x(2));
y=[y1;y2;y3;y4];
px=C*y;
piD=px(1); pid=px(2);
piQ=px(3); piq=px(4);
Te=(3/2)*P*(x(2)*LM*x(3)-x(4)*LM*x(1));
pwr=(Te-D*x(5)-TL)/J;
xprime=[piD;pid;piQ;piq;pwr;Te];
t0=0;
tfinal=2;

60
tspan=[t0 tfinal];
x0=[0 0 0 0 0];
[t,x]=ode45('indmatrix',tspan,x0);
ID=x(:,1);
Id=x(:,2);
IQ=x(:,3);
Iq=x(:,4);
wr=x(:,5);
VM=sqrt(2)*415;
F=50;
W=2*pi*F;
LM=0.318;
Rs=2;
Rr=2;
P=2;
D=0.0055;
VD=-VM*sin(W*t);
VQ=VM*cos(W*t);
Te=(3/2)*P*(LM*x(:,2).*x(:,3)-LM*x(:,4).*x(:,1));
figure(1),plot(t,ID),
title('Stator current ID');
xlabel('time(s)');
ylabel('current(A)');
figure(2),plot(t,Id),
title('Rotor current Id');
xlabel('time(s)');
ylabel('current(A)');
figure(3),plot(t,IQ),
title('Stator current IQ');
xlabel('time(s)');
ylabel('current(A)');
figure(4),plot(t,Iq),
title('Rotor current Iq');
xlabel('time(s)');

61
ylabel('current(A)');
figure(5),plot(t,wr),
title('Rotor Speed');
xlabel('time(s)');
ylabel('speed r.p.s');
figure(6),plot(t,Te),
title('Torque');
xlabel('time(s)');
ylabel('torque N.m)');
figure(7),plot(wr,Te),
title('Torque-Speed');
xlabel('Speed,r/min');
ylabel('Torque N.m');
Loss1=(3/2)*Rs*((ID.*ID)+(IQ.*IQ));
Loss2=(3/2)*Rr*((Id.*Id)+(Iq.*Iq));
Loss3=D*(wr.*wr);
figure(8),plot(t,Loss1),
title('Stator Losses');
xlabel('time(s)');
ylabel('losses(W)');
figure(9),plot(t,Loss2),
title('rotor losses');
xlabel('time(s)');
ylabel('losses(w)');
figure(10),plot(t,Loss3),
title('friction losses');
xlabel('time(s)');
ylabel('losses(w)');
IAs=(x(:,3).*cos(W*t)+x(:,1).*sin(W*t))/sqrt(2);
IBs=(x(:,3).*cos(W*t-2*pi/3)+x(:,1).*sin(W*t-2*pi/3))/sqrt(2);
ICs=(x(:,3).*cos(W*t+2*pi/3)+x(:,1).*sin(W*t+2*pi/3))/sqrt(2);
figure(11),plot(t,IAs),
title('IAs');
xlabel('time(s)');

62
ylabel('current(A)');
figure(12),plot(t,IBs),
title('IBs');
xlabel('time(s)');
ylabel('current(A)');
figure(13),plot(t,IAs),
title('ICs');
xlabel('time(s)');
ylabel('current(A)');
figure(14),plot(t,IAs,'r',t,IBs,'b',t,ICs,'k'),
title('Is');
xlabel('time(s)');
ylabel('current(A)');
VA=(VQ.*cos(W*t)+VD.*sin(W*t));
VB=(VQ.*cos(W*t-2*pi/3)+VD.*sin(W*t-2*pi/3));
VC=(VQ.*cos(W*t+2*pi/3)+VD.*sin(W*t+2*pi/3));
figure(15),plot(t,VA),
title('VA');
xlabel('time(s)');
ylabel('volt(V)');
figure(16),plot(t,VB),
title('VB');
xlabel('time(s)');
ylabel('volt(V)');
figure(17),plot(t,VC),
title('VC');
xlabel('time(s)');
ylabel('volt(V)');
figure(18),plot(t,VA,'r',t,VB,'b',t,VC,'k'),
title('Vs');
xlabel('time(s)');
ylabel('volt(V)');
IAr=(x(:,4).*cos(W*t)+x(:,2).*sin(W*t))/sqrt(2);
IBr=(x(:,4).*cos(W*t-2*pi/3)+x(:,2).*sin(W*t-2*pi/3))/sqrt(2);

63
ICr=(x(:,4).*cos(W*t+2*pi/3)+x(:,2).*sin(W*t+2*pi/3))/sqrt(2);
figure(19),plot(t,IAr),
title('IAr');
xlabel('time(s)');
ylabel('current(A)');
figure(20),plot(t,IBr),
title('IBr');
xlabel('time(s)');
ylabel('current(A)');
figure(21),plot(t,ICr),
title('ICr');
xlabel('time(s)');
ylabel('current(A)');
tloss=Loss1+Loss2+Loss3;
Pin=(3/2)*(VD.*ID+VQ.*IQ);
figure(22),plot(t,Pin,'r',t,tloss,'b'),
title('ICr');
xlabel('time(s)');
ylabel('current(A)');
IDD=(ID.*ID);
IQQ=(IQ.*IQ);
Iss=sqrt((IDD)+(IQQ));
figure(23),plot(t,Iss),
title('Iss');
xlabel('time(s)');
ylabel('current(A)');

t0=0;
tfinal=2;

64
tspan=[t0 tfinal];
x0=[0 0 0 0 0];
[t,x]=ode45('indmatrix',tspan,x0);
ID=x(:,1);
Id=x(:,2);
IQ=x(:,3);
Iq=x(:,4);
wr=x(:,5);
VM=sqrt(2)*415;
F=50;
W=2*pi*F;
LM=0.318;
Rs=2;
Rr=2;
P=2;
D=0.0055;
VD=-VM*sin(W*t);
VQ=VM*cos(W*t);
Te=(3/2)*P*(LM*x(:,2).*x(:,3)-LM*x(:,4).*x(:,1));
figure(1),plot(t,ID),
title('Stator current ID');
xlabel('time(s)');
ylabel('current(A)');
figure(2),plot(t,Id),
title('Rotor current Id');
xlabel('time(s)');
ylabel('current(A)');
figure(3),plot(t,IQ),
title('Stator current IQ');
xlabel('time(s)');
ylabel('current(A)');
figure(4),plot(t,Iq),
title('Rotor current Iq');
xlabel('time(s)');

65
ylabel('current(A)');
figure(5),plot(t,wr),
title('Rotor Speed');
xlabel('time(s)');
ylabel('speed r.p.s');
figure(6),plot(t,Te),
title('Torque');
xlabel('time(s)');
ylabel('torque N.m)');
figure(7),plot(wr,Te),
title('Torque-Speed');
xlabel('Speed,r/min');
ylabel('Torque N.m');
Loss1=(3/2)*Rs*((ID.*ID)+(IQ.*IQ));
Loss2=(3/2)*Rr*((Id.*Id)+(Iq.*Iq));
Loss3=D*(wr.*wr);
figure(8),plot(t,Loss1),
title('Stator Losses');
xlabel('time(s)');
ylabel('losses(W)');
figure(9),plot(t,Loss2),
title('rotor losses');
xlabel('time(s)');
ylabel('losses(w)');
figure(10),plot(t,Loss3),
title('friction losses');
xlabel('time(s)');
ylabel('losses(w)');
IAs=(x(:,3).*cos(W*t)+x(:,1).*sin(W*t))/sqrt(2);
IBs=(x(:,3).*cos(W*t-2*pi/3)+x(:,1).*sin(W*t-2*pi/3))/sqrt(2);
ICs=(x(:,3).*cos(W*t+2*pi/3)+x(:,1).*sin(W*t+2*pi/3))/sqrt(2);
figure(11),plot(t,IAs),
title('IAs');
xlabel('time(s)');

66
ylabel('current(A)');
figure(12),plot(t,IBs),
title('IBs');
xlabel('time(s)');
ylabel('current(A)');
figure(13),plot(t,IAs),
title('ICs');
xlabel('time(s)');
ylabel('current(A)');
figure(14),plot(t,IAs,'r',t,IBs,'b',t,ICs,'k'),
title('Is');
xlabel('time(s)');
ylabel('current(A)');
VA=(VQ.*cos(W*t)+VD.*sin(W*t));
VB=(VQ.*cos(W*t-2*pi/3)+VD.*sin(W*t-2*pi/3));
VC=(VQ.*cos(W*t+2*pi/3)+VD.*sin(W*t+2*pi/3));
figure(15),plot(t,VA),
title('VA');
xlabel('time(s)');
ylabel('volt(V)');
figure(16),plot(t,VB),
title('VB');
xlabel('time(s)');
ylabel('volt(V)');
figure(17),plot(t,VC),
title('VC');
xlabel('time(s)');
ylabel('volt(V)');
figure(18),plot(t,VA,'r',t,VB,'b',t,VC,'k'),
title('Vs');
xlabel('time(s)');
ylabel('volt(V)');
IAr=(x(:,4).*cos(W*t)+x(:,2).*sin(W*t))/sqrt(2);
IBr=(x(:,4).*cos(W*t-2*pi/3)+x(:,2).*sin(W*t-2*pi/3))/sqrt(2);

67
ICr=(x(:,4).*cos(W*t+2*pi/3)+x(:,2).*sin(W*t+2*pi/3))/sqrt(2);
figure(19),plot(t,IAr),
title('IAr');
xlabel('time(s)');
ylabel('current(A)');
figure(20),plot(t,IBr),
title('IBr');
xlabel('time(s)');
ylabel('current(A)');
figure(21),plot(t,ICr),
title('ICr');
xlabel('time(s)');
ylabel('current(A)');
tloss=Loss1+Loss2+Loss3;
Pin=(3/2)*(VD.*ID+VQ.*IQ);
figure(22),plot(t,Pin,'r',t,tloss,'b'),
title('ICr');
xlabel('time(s)');
ylabel('current(A)');
IDD=(ID.*ID);
IQQ=(IQ.*IQ);
Iss=sqrt((IDD)+(IQQ));
figure(23),plot(t,Iss),
title('Iss');
xlabel('time(s)');
ylabel('current(A)');

t0=0;
tfinal=2;

68
tspan=[t0 tfinal];
x0=[0 0 0 0 0 0];
[t,x]=ode45('csormf',tspan,x0);
ID=x(:,1);
Id=x(:,2);
IQ=x(:,3);
Iq=x(:,4);
wr=x(:,5);
VM=sqrt(2)*415;
F=50;
W=2*pi*F;
LM=0.318;
Rs=2;
Rr=2;
P=2;
D=0.0055;
Sm=5.5;
Fm=20;
VD=VM*(sin(2*pi*F*t)+(Sm/(2*F*F))*(F-Fm)*sin((2*pi)*(F-Fm)*t-
pi)+(Sm/(2*F*F))*(F+Fm)*sin((2*pi)*(F+Fm)*t-pi));
VQ=-VM*(cos(2*pi*F*t)+(Sm/(2*F*F))*(F-Fm)*cos((2*pi)*(F-Fm)*t-
pi)+(Sm/(2*F*F))*(F+Fm)*cos((2*pi)*(F+Fm)*t-pi));
%VD=-VM*sin(W*t);
%VQ=VM*cos(W*t);
Te=(3/2)*P*(LM*x(:,2).*x(:,3)-LM*x(:,4).*x(:,1));
figure(1),plot(t,ID),
title('Stator current ID');
xlabel('time(s)');
ylabel('current(A)');
figure(2),plot(t,Id),
title('Rotor current Id');
xlabel('time(s)');
ylabel('current(A)');
figure(3),plot(t,IQ),

69
title('Stator current IQ');
xlabel('time(s)');
ylabel('current(A)');
figure(4),plot(t,Iq),
title('Rotor current Iq');
xlabel('time(s)');
ylabel('current(A)');
figure(5),plot(t,wr),
title('Rotor Speed');
xlabel('time(s)');
ylabel('speed r.p.s');
figure(6),plot(t,Te),
title('Torque');
xlabel('time(s)');
ylabel('torque N.m)');
figure(7),plot(wr,Te),
title('Torque-Speed');
xlabel('Speed,r/min');
ylabel('Torque N.m');
Loss1=(3/2)*Rs*((ID.*ID)+(IQ.*IQ));
Loss2=(3/2)*Rr*((Id.*Id)+(Iq.*Iq));
Loss3=D*(wr.*wr);
figure(8),plot(t,Loss1),
title('Stator Losses');
xlabel('time(s)');
ylabel('losses(W)');
figure(9),plot(t,Loss2),
title('rotor losses');
xlabel('time(s)');
ylabel('losses(w)');
figure(10),plot(t,Loss3),
title('friction losses');
xlabel('time(s)');
ylabel('losses(w)');

70
IAs=(x(:,3).*cos(W*t)+x(:,1).*sin(W*t))/sqrt(2);
IBs=(x(:,3).*cos(W*t-2*pi/3)+x(:,1).*sin(W*t-2*pi/3))/sqrt(2);
ICs=(x(:,3).*cos(W*t+2*pi/3)+x(:,1).*sin(W*t+2*pi/3))/sqrt(2);
figure(11),plot(t,IAs),
title('IAs');
xlabel('time(s)');
ylabel('current(A)');
figure(12),plot(t,IBs),
title('IBs');
xlabel('time(s)');
ylabel('current(A)');
figure(13),plot(t,IAs),
title('ICs');
xlabel('time(s)');
ylabel('current(A)');
figure(14),plot(t,IAs,'r',t,IBs,'b',t,ICs,'k'),
title('Is');
xlabel('time(s)');
ylabel('current(A)');
VA=(VQ.*cos(W*t)+VD.*sin(W*t));
VB=(VQ.*cos(W*t-2*pi/3)+VD.*sin(W*t-2*pi/3));
VC=(VQ.*cos(W*t+2*pi/3)+VD.*sin(W*t+2*pi/3));
figure(15),plot(t,VA),
title('VA');
xlabel('time(s)');
ylabel('volt(V)');
figure(16),plot(t,VB),
title('VB');
xlabel('time(s)');
ylabel('volt(V)');
figure(17),plot(t,VC),
title('VC');
xlabel('time(s)');
ylabel('volt(V)');

71
figure(18),plot(t,VA,'r',t,VB,'b',t,VC,'k'),
title('Vs');
xlabel('time(s)');
ylabel('volt(V)');
IAr=(x(:,4).*cos(W*t)+x(:,2).*sin(W*t))/sqrt(2);
IBr=(x(:,4).*cos(W*t-2*pi/3)+x(:,2).*sin(W*t-2*pi/3))/sqrt(2);
ICr=(x(:,4).*cos(W*t+2*pi/3)+x(:,2).*sin(W*t+2*pi/3))/sqrt(2);
figure(19),plot(t,IAr),
title('IAr');
xlabel('time(s)');
ylabel('current(A)');
figure(20),plot(t,IBr),
title('IBr');
xlabel('time(s)');
ylabel('current(A)');
figure(21),plot(t,ICr),
title('ICr');
xlabel('time(s)');
ylabel('current(A)');
tloss=Loss1+Loss2+Loss3;
Pin=(3/2)*(VD.*ID+VQ.*IQ);
figure(22),plot(t,Pin,'r',t,tloss,'b'),
title('ICr');
xlabel('time(s)');
ylabel('current(A)');
IDD=(ID.*ID);
IQQ=(IQ.*IQ);
Iss=sqrt((IDD)+(IQQ));
figure(23),plot(t,Iss),
title('Iss');
xlabel('time(s)');
ylabel('current(A)');

References

72
[1] B.L. Theraja and A.K. Theraja, A Textbook of Electrical Technology,
First Multi Color Edition 2005, S. Chand & Company Ltd.
[2] P.G. Cummings, W. D. Bowers and W.J. Martiny, Induction Motor
Efficiency Test Methods, IEEE Transactions on Industry Applications,
Vol. IA-17, No. 3, May/June1981.
[3] C. Grantham, H. Tabatabaei-Yazdi and M. F. Rahman, A Novel Method
for Rapid Efficiency Measurement of Three Phase Induction Motors,
IEEE Transactions on Energy Conversion, Vol. 14, No. 4, December
1999, pp. 1236-1240.
[4] D. J. McKinnon and C. Grantham, Improved Efficiency Test Methods
for Three Phase Induction Machines, Industry Application Conference,
2005. Fourtieth IAS Anual Meeting. Conference Record of the 2005,
Vol. 1, 2-6 October 2005, pp. 466-473.
[5] N. K. Ghai, IEC And NEMA Standards for Large Squirrel Cage Induc-
tion Motors A Comparison, IEEE Transactions on Energy Conversion,
Vol. 14, No. 3, September 1999, pp.545-552.
[6] A. Boglietti, A. Cavagnino, M. Lazzari and M. Pastorelli, International
Standards for the IM Efficiency Evaluation: A Critical Analysis of the
Stray Load Loss Determination, IEEE Transactions on Industry
Applications, Vol. 40, No. 5, September/October 2004, pp. 1294- 1301.
[7] A. Boglietti, A. Cavagnino, M. Lazzari and M. Pastorelli, IM Efficiency
Measurements in accordance to IEEE 112-B, IEC 34-2 and JEC 37
International Standards, Electric Machines and Drives Conference,
2003, EMDC03, IEEE International, Vol. 3, 1-4 June 2003, pp. 1599-
1605.

[8] A. Wallace, A. von Jouane, E. Widenbrg, J. Douglass, C. Wohlgemuth


G. Wainwright, A Laboratory Assessment of In- Service Motor Efficie-

73
ncy Testing Methods, IEEE IEMDC Proceedings 1997, WC 1.7-1.
[9] J.S. Hsu, J.D. Kueck, M. Olszewski, D.A. Casada, P.J. Otaduy, and L.
M. Tolbrt, Comparison of Induction Motor Field Efficiency Evaluation
Methods, IEEE Transactions on Industry Applications, Vol. 34, No. 1,
January/February 1998.
[10] IEEE Standard Test Procedure for Polyphase Induction Motors and
Generators, IEEE Standard 112-1991, 1991.
[11] NEMA Standard Publication No. MG1, National Electrical Manufacture
Association (NEMA), Washington, D.C., 1993.
[12] J. D. Kueck, M. Olszewski, D. A. Casada, J. S. Hsu, P. J. Otaduy, and
L. M. Tolbert, Assessment of methods for estimating motor efficiency
and load under field conditions, Lockheed Martin Energy Research
Corp., Oak Ridge, TN, Rep. ORNL/TM-13165, Jan. 1996.
[13] Understanding AC motor efficiency, Electrical Apparatus Service
Association, St. Louis, MO.
[14] R. E. Oesterlei, Motor efficiency test methodsApple and oranges,
Power Transmission Design, vol. 22, no. 5, pp. 4143, May 1980.
[15] H. Jordan and A. Gatozzi, Efficiency Testing of Induction Machines,
in Conf. Rec. IEEE-IAS Annu. Meeting, Cleveland, OH, Sept. 30-
Oct. 5, 1979, pp. 284289.
[16] A. H. Bonnett, An update on ac induction motor efficiency, IEEE
Trans. Ind. Applicat., vol. 30, pp. 13621372, Sept./Oct. 1994.
[17] J. Soltani, B. Szabados and J. Hoolboom, A New Synthetic Loading
for Large Induction Machines with No Feedback into the Power
System, IEEE Transactions on Energy Conversion, Vol. 17, No. 3,
September 2002, pp. 319- 324.

[18] C. Grantham and H. Tabatabaei-Yazdi, A novel machineless dynamo-


meter for load testing three-phase induction motors, in Proc. IEEE

74
1999 Int. Conf. Power Electron. Drive Syst. (PEDS ), Jul., Hong
Kong, pp. 579584.
[19] Schwenk, H. R., Equivalent loading of induction machines for temper-
atures tests, Trans IEEE, Vol. PAS-96, No 4, July/August 1977
pp. 1126-1 131.
[20] M. Sheng and C. Grantham, Synthetic loading of three-phase induction
motors by magnetic field magnitude modulation, IEE Proc.-Electronic
Power Appl., Vol. 141, No. 2, March 1994.
[21] Abdelaziz Y. M. Abbas and John E. Fletcher, The Synthetic Loading
Technique Applied to the PM Synchronous Machine, IEEE
Transactions On Energy Conversion, Vol. 26, No. 1, March 2011.
[22] P.C. Krause, O. Wasynczuk, S.D. Sudhoff , Analysis of Electric Mac-
hinery And Drive Systems, second edition, IEE Press 445 Hoes Lane,
P.O.Box 1331 Piscataway, NJ 08855-1331.

75

You might also like