You are on page 1of 145

THESIS FOR THE DEGREE OF LICENCIATE OF ENGINEERING

Vector Control of a Double-Sided PWM


Converter and Induction Machine Drive

ROLF OTTERSTEN

Department of Electric Power Engineering


CHALMERS UNIVERSITY OF TECHNOLOGY
Göteborg, Sweden 2000
Vector Control of a Double-Sided PWM Converter
and Induction Machine Drive

ROLF OTTERSTEN

c ROLF OTTERSTEN, 2000.


°

Technical Report no 368L


Department of Electric Power Engineering
Chalmers University of Technology
SE-412 96 Göteborg
Sweden
Telephone + 46 (0)31-772 1000

Chalmers Bibliotek, Reproservice


Göteborg, Sweden 2000
Abstract

This thesis describes high-performance control methods for a double-sided


PWM converter and induction machine drive. A control scheme is developed
that uses stator flux orientated vector control for the induction machine and
grid flux orientated vector control for the grid side PWM converter. The
derivation of PI-controllers for synchronous current control, direct voltage
control and stator flux control is described. The thesis also describes some
important properties of stator flux orientation and deals with decoupling the
stator flux from the torque during stator flux orientated vector control. Sev-
eral stator flux observers of various complexity are described and a Luenberger
observer with nearly constant poles is developed. Furthermore, a feed-forward
term that improves the direct voltage control during large torque transients
is developed and eight variants of pulse width modulation are studied.

iii
iv
Acknowledgements

This thesis is funded by the European Commission, in the framework of the


Non Nuclear Energy Programme Joule III, and the Swedish National En-
ergy Administration. The Joule project is entitled “Low Cost Variable Speed
System for Stall Wind Turbines”, Contract JOR3-CT98-0311, and is a co-
operative venture between Ecotècnia S.C.C.L. (Spain), Quest AB (Göteborg,
Sweden), and the Department of Electric Power Engineering at Chalmers
University of Technology.
I would like to thank my main supervisor Dr. Ola Carlsson for helping out
with EC bureaucracy.
I am very thankful Dr. Torbjörn Thiringer and Robert Karlsson for building
first-class PWM converters. I would especially like to thank Dr. Thiringer
for also providing me with measurement systems, technical discussions and
laboratory help.
I would like to acknowledge Ph.D. candidate Andreas Petersson, who con-
tributed with the theoretical work of what eventually became Paper I in this
thesis.
Thanks to Dr. Jan Svensson for discussions and proof reading.
Finally, I would also like to thank all my colleagues at the department who
have assisted me during the work on this thesis.

v
vi
Table of Contents

Abstract iii

Acknowledgements v

Table of Contents vii

1 Introduction 1
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Objective of the Thesis . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Related Research . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.4 Outline of the Thesis . . . . . . . . . . . . . . . . . . . . . . . 4
1.4.1 Report . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.4.2 Paper I . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.4.3 Paper II . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2 Double-Sided PWM Converter 7


2.1 Name Convention . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Characteristics of Double-Sided PWM Converter . . . . . . . . 7
2.2.1 Regenerative Braking . . . . . . . . . . . . . . . . . . . 7
2.2.2 Fast Speed Dynamics . . . . . . . . . . . . . . . . . . . 9
2.2.3 Boosted and Controlled DC-Voltage . . . . . . . . . . . 9
2.2.4 Arbitrary Power Factor on the Grid Side . . . . . . . . 9
2.2.5 Grid Current Harmonics . . . . . . . . . . . . . . . . . 10
2.2.6 Radio Frequency Interference . . . . . . . . . . . . . . 10
2.2.7 Disadvantages . . . . . . . . . . . . . . . . . . . . . . . 11
2.3 Semiconductor Switches . . . . . . . . . . . . . . . . . . . . . 11
2.4 Sensors and Control Components . . . . . . . . . . . . . . . . 12
2.5 Control of the Double-Sided PWM Converter . . . . . . . . . 14
2.5.1 Two Induction Machine Control Principles . . . . . . . 15
2.5.2 Two Control Strategies for a Grid Connected PWM
Converter . . . . . . . . . . . . . . . . . . . . . . . . . 18

3 Stator Flux Orientated Vector Control of an Induction Ma-


chine 21
3.1 Transient Model of an Induction Machine . . . . . . . . . . . . 21
3.1.1 Transient T-Model . . . . . . . . . . . . . . . . . . . . 22
3.1.2 Transient Inverse Γ-Model . . . . . . . . . . . . . . . . 23

vii
3.2 Internal Model Control . . . . . . . . . . . . . . . . . . . . . . 24
3.3 Some Characteristics of Stator Flux Orientation . . . . . . . . 26
3.3.1 Cross Coupling between Active Current and Stator Flux 26
3.3.2 Torque Pull-Out . . . . . . . . . . . . . . . . . . . . . 27
3.3.3 Operation in Field Weakening Region . . . . . . . . . . 30
3.3.4 Low Speed Operation . . . . . . . . . . . . . . . . . . . 31
3.4 Stator Flux Orientated Stator Flux Controller . . . . . . . . . 31
3.5 Generation of Active Current Reference . . . . . . . . . . . . . 33
3.6 Stator Flux Control Without Reactive Current Control . . . . 34
3.7 Stator Flux Orientated Current Controller . . . . . . . . . . . 34
4 Stator Flux Observers 37
4.1 Open-Loop Stator Flux Observers . . . . . . . . . . . . . . . . 37
4.1.1 Voltage Model . . . . . . . . . . . . . . . . . . . . . . . 37
4.1.2 Improved Integrator Algorithms for the Voltage Model 39
4.1.3 Current Model . . . . . . . . . . . . . . . . . . . . . . 41
4.2 Closed-Loop Stator Flux Observers . . . . . . . . . . . . . . . 42
4.2.1 Current and Voltage Model . . . . . . . . . . . . . . . 42
4.2.2 Luenberger Observer with Nearly Constant Poles . . . 44
4.3 Euler Forward Discretization . . . . . . . . . . . . . . . . . . . 50
5 Current Control and Control of the DC-Voltage for a Grid
Connected PWM Converter 51
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
5.2 Synchronization to the Grid . . . . . . . . . . . . . . . . . . . 51
5.3 DC-Voltage Controller . . . . . . . . . . . . . . . . . . . . . . 52
5.3.1 Direct Voltage Filtering . . . . . . . . . . . . . . . . . 55
5.4 Open-loop Reactive Power Control . . . . . . . . . . . . . . . 55
5.5 Design of a Current Controller for a Grid Connected PWM
Converter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
6 Carrier Based Modulation 57
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
6.2 Regular Sampling . . . . . . . . . . . . . . . . . . . . . . . . . 61
6.3 Modulation Index . . . . . . . . . . . . . . . . . . . . . . . . . 61
6.4 Triangular PWM with Zero Sequence Injection . . . . . . . . . 62
6.5 Some Zero Sequence Signals . . . . . . . . . . . . . . . . . . . 63
6.6 Avoiding Short Time Duration Switching States . . . . . . . . 72
6.7 Summary of Triangular PWM with Zero Sequence Injection . 72
7 Simulations 75
7.1 Software Setup . . . . . . . . . . . . . . . . . . . . . . . . . . 75
7.1.1 Changes in the Control System . . . . . . . . . . . . . 75
7.1.2 Simulation Environment . . . . . . . . . . . . . . . . . 77
7.2 Safety Limits . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
7.2.1 Speed Limit . . . . . . . . . . . . . . . . . . . . . . . . 77
7.2.2 Stator Current Limits . . . . . . . . . . . . . . . . . . 78
7.2.3 Voltage Limit . . . . . . . . . . . . . . . . . . . . . . . 80

viii
7.3 Simulation Results . . . . . . . . . . . . . . . . . . . . . . . . 81
8 Experiments 93
8.1 Hardware Description . . . . . . . . . . . . . . . . . . . . . . . 93
8.1.1 Double-Sided PWM Converter . . . . . . . . . . . . . . 93
8.1.2 Induction Machine . . . . . . . . . . . . . . . . . . . . 96
8.1.3 Control Computer . . . . . . . . . . . . . . . . . . . . 96
8.1.4 Pulse Width Modulation . . . . . . . . . . . . . . . . . 97
8.1.5 Sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
8.1.6 Measurement System . . . . . . . . . . . . . . . . . . . 98
8.2 Software Description . . . . . . . . . . . . . . . . . . . . . . . 98
8.3 Laboratory Experiences with Stator Flux Observers . . . . . . 99
8.4 Experimental Results . . . . . . . . . . . . . . . . . . . . . . . 100
9 Conclusions 109
10 Future Work 111
References 113
A Glossary of Symbols, Subscripts, Superscripts and Abbrevi-
ations 119
B Space Vectors 123
B.1 Space Vector Notation . . . . . . . . . . . . . . . . . . . . . . 123
B.2 Coordinate Transformations . . . . . . . . . . . . . . . . . . . 124
B.3 Torque in Terms of Space Vectors . . . . . . . . . . . . . . . . 125
C Base Values 127
D Parameter Sensitivity of Stator Flux Observers 129
D.1 Open-Loop Observers . . . . . . . . . . . . . . . . . . . . . . . 129
D.1.1 Voltage Model . . . . . . . . . . . . . . . . . . . . . . . 129
D.2 Closed-Loop Observers . . . . . . . . . . . . . . . . . . . . . . 130
D.2.1 Current and Voltage Model . . . . . . . . . . . . . . . 130
D.2.2 Luenberger Observer . . . . . . . . . . . . . . . . . . . 130
E Induction Machine Data 133
F Publications 135

ix
x
Chapter 1

Introduction

1.1 Background

Electrical machine systems consume approximately 50 % of the produced


electrical energy in the western world [1] and are used in all kinds of kinds
industrial processes, transportation systems and household appliances. The
market value of electric machine systems was $70.5B in 1997 [2] and the
value was growing at a rate of 7 % per year [2]. About 75-80 % of installed
electrical machine systems operate at constant speed while 20-25 % operate at
adjustable speed [1]. The use of adjustable speed machine systems is predicted
to increase further due to increased industrial needs for better torque, speed
and position control [2]. Furthermore, adjustable speed machine systems are
also motivated because they can reduce energy consumption for quadratic
torque loads, eliminate inrush currents and reduce mechanical wear.

A modern adjustable speed ac machine system is equipped with an adjustable


frequency drive (AFD). The AFD is a power electronic device for torque/speed
control of an electric machine. The AFD controls the torque/speed of the
electric machine by converting the fixed voltage and frequency of the grid to
adjustable values on the machine side. An AFD converter for an electric ma-
chine can be designed with various power electronic principles. The six pulse
diode rectifier/pulse width modulated (PWM) inverter with an intermediate
voltage stiff direct voltage link is perhaps the most widespread AFD technol-
ogy in industrial applications. The six pulse diode rectifier/PWM inverter
is a cost-effective and high-efficiency AFD that is hard to beat. However,
alternative “rectifier”/“inverter” topologies can be attractive in certain ap-
plications since the six pulse diode rectifier/PWM inverter only permits one
power flow direction and injects low order current harmonics to the grid. The
double-sided PWM converter (Fig. 1.1) that uses PWM converters on both
the grid and machine side is one such alternative converter topology. The
double-sided PWM converter offers a bi-directional power flow, full control of
the direct voltage and the grid current, and reduced grid current distortion.

1
  
 !" #
$!
 DC-link


 


AC machine

Grid M

Figure 1.1: Double-sided PWM converter and ac machine.

1.2 Objective of the Thesis


The objective of this thesis is to study fast and accurate torque control of a
double-sided PWM converter and induction machine drive in the medium/high
speed range. Stator flux orientated vector control of the induction machine
and grid flux orientated vector control of the grid side PWM converter have
been chosen for this purpose.

1.3 Related Research


Numerous researchers have carried out extensive work on the double-sided
PWM converter. The following presents a selection of achievements that
have been of great inspiration for the author.
Eguchi and Imura were among the first who proposed vector current control
of the grid connected voltage source converter in 1986 [3]. It was shown that
vector current control is superior to amplitude-phase control. An experimen-
tal 24-pulse forced commutated bridge, consisting of four forced commutated
thyristor inverters, of 20 kW was used and the lowest order voltage harmonic
was the 11th. The output voltages of the four 6-pulse forced commutated
thyristor inverters were phase controlled in order to obtain the desired volt-
age vector. It was shown that the inverters could operate at lagging, leading,
unity power factors and good controllability of the active and the reactive
power were also shown.
Kohlmeier et al. studied high performance GTO-based double-sided PWM
converters in 1987 [4],[5]. The reference currents were generated in rotor flux
orientated and grid flux orientated reference frames for the machine side and
grid side PWM converters, respectively. Predictive current control, i.e. the
switching instants were determined by a circular boundary around the current
reference vector [6], was used in order to obtain a switching frequency below
1.5 kHz. Several digital signal processors were used to control the converter.
The converter had a high power factor on the grid side, low grid current
distortion and bi-directional power flow. Furthermore, it was found that a
feed-forward term that is proportional to the induction machine loading could

2
improve the direct voltage control.
Svensson studied a transistor-based grid connected PWM converter and pulse
width modulation in his Ph.D. thesis [7] from 1988, along with several other
topics. Very accurate dead-beat response and low current distortion was
obtained with a switching frequency of 1 kHz.
Blaabjerg et al. [8] presented a IGBT-based double-sided PWM converter
with 4.8 kHz switching frequency in 1993. The entire control system, mon-
itoring and communication were implemented in one microcontroller. Space
vector modulation was used for both PWM converters. The induction ma-
chine was Volts-per-Hertz controlled while the grid connected PWM converter
used a vector current controller. An outer direct voltage control loop gener-
ated the active current reference for the grid connected PWM converter while
the reactive current reference was set to zero. The power factor dropped
slightly at low loads, due to non-linearities in voltage transducers and in the
grid side PWM converter, but was otherwise near unity. Similar to Kohlmeier
et al., a feed-forward term was used to improve the direct voltage control.
Several research papers on the double-sided PWM converter during the 1990’s
were focused on minimizing the direct voltage link electrolyte capacitor bank
[9],[10]. Carlsson found in 1998 that the electrolytic capacitor bank could be
replaced by a small plastic capacitor bank and a braking chopper [11].
Hansen et. al [12] recently showed that it is possible to eliminate the grid
voltage of a double-sided PWM converter if the grid is balanced. However,
the grid current became distorted for an unbalanced and distorted grid volt-
age. Grid voltage sensors together with a distortion compensating algorithm
were therefore recommended for a standard industrial grid connected PWM
converter. Furthermore, [12] also recommended discontinuous pulse width
modulation for the grid connected PWM converter in order to reduce switch-
ing losses or, alternatively, reduce the size of the grid filter.
The historical development of vector control for induction machines is too
extensive to be fully covered here. Only the inventors of vector control along
with research that has made a significant impact on this thesis are mentioned
in the following.
Vector control is a popular high performance control principle for induction
machine regulation that was developed by Blaschke [13] and Hasse [14] in
the late 1960’s. Vector control has since then been constantly developed
and there are several variants of vector control today. This thesis focuses on
vector control that uses stator flux orientation and synchronous PI current
control. Xu et. al. have written several research papers, spanning from
1988 to 1992, on stator flux orientated vector control [15],[16],[17],[18]. These
papers describe the characteristics of stator flux orientation, operation in the
field weakening region, propose a current decoupling term and treat sensorless

3
operation. The papers are highly recommended reading for studies on stator
flux orientated vector control.
The excellent Ph.D. thesis [19] by Lennart Harnefors, presented in 1997, has
also been a huge source of inspiration for this thesis. The thesis represents
state-of-the-art in induction machine control and is highly recommended.
This report in this thesis has especially adopted internal model control for
the design of PI-controllers from [19].

1.4 Outline of the Thesis


This thesis consists of three parts. These three parts are a technical report
and two papers, Paper I and Paper II. Thus, the following make up the present
thesis

Report: Rolf Ottersten, “Vector Control of a Double-Sided PWM Converter


and Induction Machine Drive,” Department of Electric Power Engineering,
Chalmers University of Technology, Göteborg, Sweden, May, 2000.

Paper I: Jan Svensson, Rolf Ottersten, “Shunt Active Filtering of Vector


Current-Controlled Voltage Source Converter at a Moderate Switching Fre-
quency,” IEEE Transactions on Industry Applications, vol. 35, no. 5, pp.
1083-1090, Sep./Oct. 1999.

Paper II: Rolf Ottersten, Jan Svensson, “Vector Current Controlled Voltage
Source Converter - Deadbeat Control and Saturation Strategies,” in Proc.
IEEE Nordic Workshop on Power and Industrial Electronics, Espoo, Finland,
pp. 65-70, August 26-27, 1998.

The main part of the thesis consists of the technical report. Paper I and
Paper II are included in Appendix F of the report. The following subsections
provide a content description of the technical report and the two papers.

1.4.1 Report
The report describes vector control of the double-sided PWM converter and
induction machine drive and also investigates some different variants of pulse
width modulation. The report is divided into ten chapters.

Chapter 2 presents some important characteristics of the double-sided PWM


converter. The characteristics are in some cases compared with the char-
acteristics of the commonly used six pulse diode rectifier/PWM inverter.
Some common control schemes for induction machine control and the
control of a grid connected PWM converter are also described.

4
Chapter 3 derives controllers for stator flux regulation and for current con-
trol during stator flux orientation. Transient models of the induction
machine and some characteristics of stator flux orientation are also de-
scribed.
Chapter 4 describes some open-loop and closed loop stator flux observers.
One problem with a direct open-loop stator flux observer is its sensitivity
to measurement bias. Recently presented methods for handling measure-
ment bias are described. Two closed-loop observers are described. The
first closed-loop observer is the well known closed-loop observer [20] that
combines the best properties of the current and voltage models. A second
closed-loop observer, a Luenberger observer with nearly constant poles,
is developed in this thesis.
Chapter 5 develops a vector control scheme for the current control and the
direct voltage control of a grid connected PWM converter. A feed-
forward term is used to improve direct voltage regulation in order to
improve direct voltage control during heavy torque transients.
Chapter 6 deals with some features of carrier based modulation and espe-
cially triangular PWM with zero-sequence injection. Both conventional
and discontinuous zero-sequence signals are described. The differences
between the zero-sequence signals concerning the resulting current dis-
tortion and switching losses are theoretically analyzed.
Chapter 7 describes software implementation and presents simulation re-
sults of a double-sided PWM converter. In addition, some software safety
limits are implemented and described.
Chapter 8 presents experimental results of the developed control scheme for
stator flux orientated vector control of induction machines. The direct
voltage link is charged by a dc-machine during the experiments.
Chapter 9 contains the conclusions of this thesis.
Chapter 10 provides recommendations for future work on the double-sided
PWM converter.

The report also contains five appendices:


Appendix A lists the symbols, subscripts, superscripts and abbreviations
that are used in this thesis.
Appendix B describes some fundamentals of space vector theory.
Appendix C presents the per-unit system that is used in this thesis.

5
Appendix D provides the frequency response functions of the stator flux
observers that are studied in this thesis.
Appendix E lists the parameters of the induction machine that is used in
the theoretical and simulation parts of this thesis.
Appendix F contains Paper I and Paper II, which are described in more
detail below.

1.4.2 Paper I

The grid current of double-sided PWM converter is fully controlled. This


permits to use the grid connected PWM converter of the double-sided PWM
converter as a parallel active power filter. A parallel active power filter is
capable of cancelling low order current harmonics from non-linear loads. The
parallel active power must be able measure the load current and to detect
the current harmonics in real time/nearly real time. Paper I compares four
time delayed methods to detect current harmonics. Stationary conditions are
assumed so the detection methods are sufficient. A vector current controller
with deadbeat gain is used to track current harmonics.

1.4.3 Paper II

A vector current controller with deadbeat gain is suitable to use for an par-
allel active power filter. Deadbeat gain permits a high bandwidth and an
accurate cancellation of current harmonics. One drawback of using deadbeat
gain is that the current controller easily saturates due to limited dc-voltage.
Paper II presents control methods for reducing the negative influence from
voltage saturation. Furthermore, a one sample calculation delay may appear
in the current control due to the implementaion of the control computer. The
paper proposes the use of a Smith predictor in order to compensate for the
calculation delay.

6
Chapter 2

Double-Sided PWM Converter

2.1 Name Convention


The double-sided PWM converter and its two sub PWM converters are known
by various names in technical literature. The author prefers the names grid
side or grid connected PWM converter and machine side PWM converter for
the two sub PWM converters. These names clarify the locations of the two
sub PWM converters and indicate that the power flow can be reversed at
any time instant. Furthermore, two three-phase PWM converters with an
intermediate stiff direct voltage and with only two possible phase potential
levels are referred to as a double-sided PWM converter.

2.2 Characteristics of Double-Sided PWM Converter


Fig. 2.1 shows three important characteristics of the double-sided PWM con-
verter. These three characteristics are operation within all four quadrants of
the torque-speed domain, nearly sinusoidal and fully controlled currents on
both the machine side and the grid side, and a unity power factor is possible on
the grid side. The four quadrant operation permits bi-directional power flow
and hence four quadrant operation. The fully controlled grid current gives
full control of the direct voltage and full control of the grid power factor,
preferably set to unity. Finally, the fully controlled grid current also enables
the use of a the grid connected PWM converter as a parallel active power
filter. The following sections discuss the characteristics of the double-sided
PWM converter in more detail. Some differences/similarities between the
double-sided PWM converter and the six pulse diode rectifier/PWM inverter
are also discussed.

2.2.1 Regenerative Braking


An electric machine can normally operate both as a motor and as a generator.
When an adjustable speed machine operates as a generator, for instance dur-

7
ω . *,/#0132#--4 Max.
1.5

transient e
Max. g1
power 1
cont.
i
ωb power ig1 s1

%'& ()&*,+#-

Current, voltage (pu)


Generator Forward Max. Max. 0.5
motor cont. transient
operation 2 1 operation torque torque 0

Reverse 3 4 Generator
motor %"& ()&!*+#- Te
−0.5
operation operation
−ωb
−1

−1.5
0.02 0.025 0.03 0.035 0.04 0.045 0.05
Time (s)
(a) (b)

Figure 2.1: (a) Four-quadrant torque-speed diagram. (b) Simulated grid voltage eg1 , grid
current ig1 and induction machine current is1 with 4.8 kHz switching frequency and 10 %
grid filter impedance.

ing braking, the regenerated power must flow from the machine to the direct
voltage link which raises the voltage on the direct voltage link. In some cases,
the regenerated energy is small and can be absorbed by the direct voltage
link without problems. During braking though, overvoltage may occur on the
direct voltage link which causes the drive to trip and the capacitor bank is
then exposed to additional stress. Applications that require frequent accel-
eration and deacceleration may therefore require regenerative braking. The
braking power can be transferred back to the grid (regenerative braking) or
it can be dissipated by a braking chopper on the direct voltage link (dynamic
braking). The braking power can be calculated from the load cycle profile and
the load torque characteristic [21]. Regenerative or dynamic braking should
be considered if Pbr /Pnom > 0.1, where Pbr is the regenerated power at top
speed and Pnom is the nominal power of the drive [21].

The double-sided PWM converter, with its grid connected PWM converter, is
a four quadrant converter and have regenerative braking built in. The power
flow between the grid and direct voltage link can therefore be reversed at any
time instant. However, a standard six pulse diode rectifier/PWM inverter
is non-generative converter and cannot reverse the power flow. The direct
voltage link is therefore normally equipped with a braking chopper if dynamic
braking is required. The braking chopper transfers the regenerated power into
heat. Additional cooling may be required due to the dissipated heat, which in
that case introduces additional cost. As an alternative to the braking chopper,
dc-braking [1] can also be used to obtain some, low-performance, electric
braking. To sum up, the double-sided PWM converter can be very cost-
competitive for regenerative applications. The double-sided PWM converter
also places less stress on the dc-capacitor for such applications.

8
The braking capability of an electric machine can also be obtained with a
mechanical brake. However, a mechanical brake leads to wear on mechanical
parts, possibly less smooth braking and a waste of energy. Regenerative
braking is therefore often a better choice. Note, though, that a mechanical
brake is sometimes necessary as an emergency backup to the electrical brake.
Some AFD applications which are likely to require regenerative braking are
lifts, ramped stopping of flywheels or large mechanical arms, and certain
material handling and packaging lines [21].

2.2.2 Fast Speed Dynamics


The four-quadrant operation permits fast speed dynamics. The fast speed
dynamics are closely related to the regenerative braking capability, which
permits faster acceleration and deacceleration compared to a six pulse diode
rectifier/PWM inverter without braking chopper.

2.2.3 Boosted and Controlled DC-Voltage


The controlled direct voltage of a double-sided PWM converter permits a low
dependency on machine loading and grid voltage sags [23]. Furthermore, the
operation of the grid side PWM converter is similar to a dc/dc boost converter
and ac inductors are therefore required. The direct voltage must be at least
equal to the peak of the grid phase-to-phase voltage for successful operation
and in practice it must be higher. The boosted direct voltage permits fast
torque response. Alternatively, the boosted direct voltage also permits the
use of a machine with higher nominal voltage. Unfortunately, the boosted
direct voltage may also increase dv/dt problems [24].
The direct voltage is charged with short-duration current pulses during normal
operation. The short-duration pulses permit the use of a smaller capacitor
bank compared to a six pulse diode rectifier/PWM inverter, which requires a
rather bulky dc-capacitor in order obtain a smooth direct voltage.

2.2.4 Arbitrary Power Factor on the Grid Side


The grid connected PWM converter permits arbitrary power factor on the
grid side. This permits the use of the grid connected PWM converter as a
reactive power compensator. Often, though, unity power factor is desired.
Unity power factor minimizes the grid current RMS value. A unity power
factor can also be obtained with a six pulse diode rectifier/PWM inverter.
However, the low order current harmonics of the diode rectifier increase the
RMS value of the current compared to when using a grid connected PWM
converter. Both the six pulse diode rectifier and the grid connected PWM
converter reduce the current RMS value compared to a directly connected

9
induction machine, without phase compensation though, since the grid no
longer has to supply reactive power.

2.2.5 Grid Current Harmonics


No stand-alone grid connected converter/rectifier emits a purely sinusoidal
grid current. The grid current is always more or less distorted. The grid
current of a grid connected PWM converter is nearly sinusoidal but contains
high order current harmonics at essentially the multiples of the switching fre-
quency. The six pulse diode rectifier on the other hand emits low order cur-
rent harmonics such as 5th, 7th, 11th... The low order harmonics may cause
transformer overheating, motor failure, fuse blowing, capacitor failures [25].
Relevant standards for grid harmonics are [26]:
IEC 1000-3-2: Sets harmonic current limits for smaller equipment, typical
household appliances or similar below 16 A/phase.
IEC 1000-3-4: Sets harmonic current limits for larger customers, i.e. indus-
try, for loads between 16-75 A. Deals with both individual equipment
and the whole system installation and gives a consideration of the grid
short circuit ratio.
IEEE 519: Limits current and voltage harmonics at the point of common
connection. Applicable mainly to larger consumers and for continuous
operation and in USA.
It is often claimed that the use of double-sided PWM converters are motivated
due to better power quality than the six pulse diode rectifier/PWM inverter.
However, a dilemma for the double-sided PWM converter is that the diode
rectifier/PWM inverter offers cost-effective solutions to reduce the grid cur-
rent harmonics [27],[28],[29]. One interesting solution is to mix groups of six
pulse diode rectifier/PWM inverters with one double-sided PWM converter
that has active power filtering capability [29]. It must also be mentioned that
the grid current of the grid connected PWM converter is not purely sinusoidal
but contains high order current harmonics. The high order current harmonics
can introduce voltage harmonics that disturb senstive electronics [30]. Filter-
ing of the high order current harmonics is somehow simpler than low order
harmonics, though. The high order current harmonics can be filtered with
a rather small notch- or LCL-filter likely combined with common mode and
radio frequency interference filters.

2.2.6 Radio Frequency Interference


The switchings of a PWM converter in conjunction with various stray capaci-
tances and conductor inductances cause charging/discharging currents during

10
each switching transient. The charging/discharging currents can be divided
into common mode and differential mode components and represent damped
oscillations with very high frequencies. The damped oscillations create a wide-
band radio frequency interference (RFI) that may interfere with the control
equipment and other electric equipment. The European standard EEF 82/499
sets tolerable levels for RFI emission over an AC grid. The common mode
current creates an RFI that seems to be especially troublesome since the com-
mon mode current may take various long paths back to the PWM converter.
Special RFI filters [31] possibly on both the grid side and machine side con-
verter are probably the only successful solution for reducing the RFI and to
comply with standards. Shielded and properly grounded power cables and
control cables are also useful for reducing RFI problems.

2.2.7 Disadvantages
Two main disadvantages of the double-sided PWM converter are a higher
price and higher losses as compared to a six pulse diode rectifier/PWM in-
verter. The grid connected PWM converter is rather rare on the market today
and the converter require more semiconductor components, more sensors and
a more complicated control system. An increased cost of about 40-50 % com-
pared to the six pulse diode rectifier/PWM inverter is possible [29]. The
higher losses of a PWM converter are because both switching and conduc-
tion losses are present and the conduction losses of an IGBT are higher than
that of a diode. A reasonable assumption is also that a double-sided PWM
converter produces more RFI than a six pulse diode rectifier/PWM inverter
since a double-sided PWM converter has twice as many RFI sources.

2.3 Semiconductor Switches


PWM converters require semiconductor switches that are capable of switch-
ing on/off at arbitrary time instants. Typical switching frequencies are 2-
8 kHz. Fig. 2.2 shows ratings of single semiconductor switches as they were
in 1997 [6]. Note that this figure is likely to be outdated since new and im-
proved devices are entering the market continuously. The MOSFET is the
preferred semiconductor valve for low power/low voltage applications. The
MOSFET has low conduction losses and low switching losses. The MOS-
FET also has a parasitic anti-parallel diode that may be used as a built-in
free-wheeling diode if its recovery time is acceptable. However, the on-state
resistance of a normal MOSFET has a strong dependency on the blocking
voltage. This makes the IGBT the preferred choice for applications in the
mid/high voltage range. The IGBT was introduced in the 1980’s and has,
today, nearly replaced the bipolar junction transistor (BJT) as the preferred
semiconductor switch in adjustable frequency drives. Both the MOSFET and

11
6

4
Imax (kA)

0 Mosfet
BJT
0.2 IGBT
GTO
1
3 6
fsw (kHz) 4 5
10 3
20 1 2
0 Umax (kV)

Figure 2.2: Ratings of semiconductor switches as they were in 1997.

the IGBT are voltage controlled which allows simple driver circuits. The
GTO has high switching losses and requires a complicated driver since it is
current controlled. The GTO is therefore mainly used for high power appli-
cations beyond 2-3 MW [6]. The material in this thesis is not focused on any
specific semiconductor switch although IGBT:s are assumed and have been
used during experiments.

2.4 Sensors and Control Components


A double-sided PWM converter consists of two PWM converters with an
intermediate voltage stiff direct voltage link. The grid side PWM converter
maintains constant direct voltage and handles the reactive power exchange
with the grid while the machine side PWM converter controls the torque and
flux of the induction machine.
A traditional implementation of a double-sided PWM converter requires a
large amount of sensors. All electrical sensors should normally have galvanic
insulation between the power electronics and the low voltage electronics. It
may not be required to anti-alias filter the sampled variables but at least
low-pass filters are useful for suppressing switching noise. Fig. 2.3 shows a
possible sensor implementation which consists of
• Current sensors on the grid side, required for current control.
• Current sensors on the machine side, required for current control.
• Voltage sensors on the grid side. These sensors are required for synchro-
nization to the grid.

12
HG A"IJK;!L< 9=;=M>?5 7!A":B ;C!D ;C
E C$9=F< 9=;>@?57!A':B ;CD ;C DC-link 56!7$89 : ; < 9=;>@?5 7A':B ;CD ;C
Grid Induction machine

eg ig
is θr
Sample&hold u dc

Driver Driver Sample&hold


3 3
Modulator   Modulator
6 3 ug us 3 2
Control

Figure 2.3: Double-sided PWM converter with sensors and control components.

• Voltage sensor on the direct voltage link, required for control of the direct
voltage.
• Shaft sensor for rotor position/speed, required for position/speed control.

It is desirable to reduce the amount of sensors since sensors represent an in-


creased risk of system failure and an increased cost. Speed sensorless systems
are common today for less critical applications. Furthermore, the current sen-
sors on the ac side can be moved to the direct voltage link [32] since two of the
phase currents appear twice on the direct voltage link during each switching
period. There are also recent research efforts to eliminate the three voltage
sensors on the grid side [12].
The control algorithms of the back-to-back system are often implemented
in a digital signal processor (DSP). The DSP operates with a sampling fre-
quency of some kHz and the double-sided PWM converter is therefore a crit-
ical real-time application. The reference voltages of the controller are sent
to modulators which determine the switching instants of the semiconductor
switches. The realization of the switching instants are handled by separate
drivers. Blanking time must be added to the driver signals in order to avoid
short-circuits in the phase-legs. The drivers should be galvanically separated
from the control system. Optical fibers are often used for this purpose.
It is sound to let a diagnostic program monitor the operation of the double-
sided PWM converter. The diagnostics program can for instance detect sensor
failure, grid disturbances, overspeed, overcurrents and overvoltage on the di-
rect voltage link. The monitor program can take suitable actions in case of
overloading or faults or it can trip the converter. The diagnostic program can
also check the condition of the sensors before starting the system. Smaller
measurement offsets can be stored and compensated by the control program.
In addition to the diagnostic program, fast backup hardware for overvoltage

13
and overcurrent protection is also useful. The monitor program may also have
a user-friendly self-commissioning feature. With the self-commissioning fea-
ture, the control system automatically identifies the induction machine and
grid filter parameters before actual operation starts.
In addition to sensors and control components, several passive components
stand for a significant part and cost of the double-sided PWM converter. The
passive components are, for instance filters, fuses and overvoltage protection.
The passive components are treated very briefly, or not at all, in this thesis.
Instead, this thesis focuses on the control of the double-sided PWM converter.

2.5 Control of the Double-Sided PWM Converter


Fig. 2.4 shows the block diagram of the vector control system for a double-
sided PWM converter driving an induction machine, which is the main theme
of this thesis. This figure is greatly simplified and does not display all details
of the control system. Both the grid connected PWM converter and the
induction machine are vector controlled and have multi loop cascaded control
systems of a very similar structure. Table 2.1 summarizes the similarities
of the two control systems. The current control loop is the inner control
loop for both systems. The main difference is how the active and reactive
current references are generated. The induction machine control may have an
additional outer loop for position control while, of course, the speed control
loop is not used during torque control.
A cascaded control system, such as vector control, is a form of state feedback.
One important advantage of state feedback is that the inner control loop can
be made very fast. For vector control, current control is the inner control
loop. The fast inner current control nearly eliminates the influence from
parameter variations, cross coupling, disturbances and minor non-linearities
in the control process [33]. A cascaded control system such as vector control
is therefore in many cases superior to simpler control principles without an
inner current control loop. Two such control methods are Volts-per-Hertz
control for the induction machine and amplitude-phase control for the grid
connected PWM converter. These two control methods do not have an inner
loop for current control.

Table 2.1: Similarities of vector control of an induction machine and vector control of a grid
connected PWM converter.
Control loop Induction machine Grid connected PWM converter
Inner loop Current control ⇔ Current control
Active loop Speed control ⇔ Direct voltage control
Reactive loop Flux control ⇔ Reactive power control

14
pc)deORYRNxV h$O]_TORORY[kgV `^RhW` V `_s h_hWw X Ql]_ORYRYR`]'V `_Q v ]_V X g[`MkWX T V `_h
software
DC-voltage PN ORQRSRhWX T,QZUWV Xa ORX QRYZ` UWY[Q
c)desNf]_ORYRgR`_hWV `_h

u dc
u dc −
 
control ig Current u g
Grid
 Open-loop − control
Q
θ p UWhWV `)a X,UWY
control ig Grid
Possible
gf
V O r O[T,UWh voltage
t pFu gRORT V,U s `M]_ORYRV hWORTy
co-ordination: Current
feedback Coordinate
modulation
t pFu gRORT V,U s `MkW`'`QRmnUW]_o transform.
DC-link

p OY[V hWORT[ORkV ^R`MX Y[QS[]_V X ORY \ UW]_^RX YR`)w subsystem

PN ORQRUW]_SR^RT,XUWYV X `bORYZa X QRUWY[` Q


 
is
n Te Open-loop

c)de\ Nf]_ORYRgR`_hWV `_h


PI
− control is
Current u s

Induction Mechanical
n − control
Flux machine subsystem
ψ 

ψ
− control is
p SRhWhW`_YRVk3``_QmUW]_o Coordinate
transform.
)i T SRjlkW`_`_QRmnUW]_o
θ
Flux
q_r `'`QlkW`_`_QRmnUW]_o observer

Figure 2.4: Block diagram of a possible vector control system for a double-sided PWM
converter.

Vector control often uses PI-controllers with various feed-forward terms in


order to improve dynamic response and to reduce cross coupling between
flux/torque and active power/reactive power. These feed-forward terms are
often in the current control and flux control and are referred to as voltage
and current decoupling, respectively. Feed-forward terms may also be used
to obtain a better co-ordinated control of the double-sided PWM converter.
Especially the direct voltage control of the grid connected PWM converter can
be greatly improved during large torque transients by using a feed-forward
term that is proportional to the induction machine loading [4],[8].

2.5.1 Two Induction Machine Control Principles


Volts-per-Hertz Control
Volts-per-Hertz control, or V/f-control, is a widely used induction machine
control principle. The control performance of Volts-per-Hertz control is sat-
isfactory for many applications. The underlying principle of Volts-per-Hertz
control is that the stator frequency is varied close to the motor speed. The
stator flux can be maintained at its rated value by varying the stator voltage
amplitude linearly with the stator frequency since the stator flux is propor-
tional to the stator voltage divided with the stator frequency in steady state.
A speed sensor is seldom required for applications with moderate performance
demands. Slip frequency compensation is often used to obtain the desired

15
q β
us
is
ωk
ψ sd d
isq
isd
θk
α (fixed)

Figure 2.5: Rotating two-axis flux orientated reference frame.

speed regardless of the loading torque.


The torque is proportional to the slip if the flux is maintained constant.
However, the slip frequency must not exceed the pull-out slip frequency in
order to maintain stable operation. The induction machine equations were
linearized in [34] and it was shown that there are regions in the magnetizing
current-stator frequency plane and in the stator voltage-stator frequency plane
where Volts-per-Hertz control becomes unstable if [34] τm /τr < 0.5, where τm
is the mechanical time constant and τr is the open circuit rotor time constant.
Due to the instability problems, Volts-per-Hertz control is mainly suitable for
applications with slow dynamic loads, although a higher value of the leakage
inductance Lσ tends to reduce the instability problem [34]. Furthermore,
Volts-per-Hertz control is likely to provide a reduced peak torque for speeds
less than 5 Hz [1]. This is due to the difficulty in maintaining constant stator
flux at low speeds due to stator resistance voltage drop. A voltage boost is
normally added in order to attempt to compensate for the stator resistance
voltage drop.

Vector Control
The principle of vector control was presented by Blasche [13] and Hasse [14]
in the late 1960’s. Vector control is a relatively complex control method
and requires more computational power and more and better sensors than
open-loop Volts-per-Hertz control. Vector control exists in many different
variants. Generally though, vector control implies that the torque and the
flux of the induction machine are controlled in a manner that is similar to a
dc-machine with a separate field winding. Vector control is implemented in
a rotating two-axis reference frame, see Fig. 2.5, that is synchronized with
one of the induction machine flux linkages. The rotating two-axis reference
frame is often referred to as a flux orientated reference frame. Vector control
regulates the length of the flux vector and the length and position of the stator
current vector in the flux orientated reference frame. The stator current can
be divided into two orthogonal components which correspond to currents
that produce flux and torque. The two current components are similar to
the field current and the armature current of a separately magnetized dc-

16
machine. The orthogonal current components allow independent, or nearly
independent, control of torque and flux.
The current controller in a vector control scheme is often implemented in the
flux orientated reference frame as well. This is because the fundamental AC
variables in the stationary reference frame become DC variables in the flux
orientated reference frame in steady state. The DC variables allow accurate
tracking of the current reference signals and the performance of the current
controller becomes independent of frequency. Fig. 2.6 shows the typical block
diagram of a PI current controller implemented in a flux orientated reference
frame, i.e. a synchronous PI-controller. The current control is the fast inner
loop in a cascaded vector control system. Flux control and speed control are
slower outer loops that generate the reactive and active current references,
respectively.
Flux orientation can be accomplished by using either direct or indirect flux
orientation. Direct flux orientation methods use an observer to estimate the
flux and directly track the flux position. Indirect rotor flux orientation re-
quires, at least theoretically, no flux observer and tracks the rotor flux position
indirectly from the rotor position and the slip angle.
Rotor flux orientation and stator flux orientation are two common flux ori-
entation methods. Only rotor flux orientation has the advantage of having
fully decoupled control of the torque and rotor flux during perfect rotor flux
orientation, i.e., a torque step does not affect the rotor flux and the rotor flux
can be controlled via an open loop. For stator flux orientation, there exists
a cross coupling between the torque and the stator flux even during perfect
flux orientation. The cross coupling in stator flux orientation can be handled
with a decoupler. Stator flux orientation has a slightly higher complexity
than rotor flux orientation due to the cross coupling and decoupled control
of torque and flux cannot be guaranteed for stator flux orientation. Further-
more, stator flux orientation with simple stator flux feedback often has poor
low speed properties due to high sensitivity to the stator resistance estimate

 + ud '

ud 
uα

id u1

PI
Voltage dq αβ 

u '   u2

iq + q decoupling uq αβ u β 123 
PI u3

cos(θ)
sin(θ)
id iα i1
dq αβ i2
iq αβ iβ
z|{_}"~}  €‚ƒ}„{†… ‡)ˆ‰~Š_‹ŒŠ3‚Š3}"‹Ž 123 i3
,‘ ’)“ {”"‹€)}~,}"‹R•ƒ”‹  ‹”‹,€‰–‹  ”~Ž—‹R˜ Two-phase Three-phase
system system
(fixed) (fixed)

Figure 2.6: PI current controller implemented in a flux orientated reference frame.

17
and sensitivity to measurement offsets. However, stator flux orientation can
still be desirable due to low parameter sensitivity at medium to high speeds
and good torque capability in the field weakening region [18].

The success of vector control depends on the accuracy of the feedback flux sig-
nal. The feedback flux signal consists of the flux position and the feedback can
also contain the flux amplitude. Errors in the estimated flux amplitude de-
grade the accuracy of torque control and affect the losses of the machine. The
losses are affected since an incorrect flux amplitude estimate either increases
the core losses or the resistive losses in the machine due to higher slip [35].
An incorrect flux position feedback decreases the dynamic performance of the
torque control and flux and torque are then no longer decoupled.

Very early vector control schemes [13] obtained the flux feedback by measur-
ing the flux linkage with Hall sensors in the airgap or by using separate flux
coils in the induction machine. However, these two methods resulted in a
more complicated machine design and reduced reliability. The flux feedback,
today, is therefore estimated from machine equations instead. The accuracy
of the estimation depends on the accuracy of the induction machine param-
eters. The parameters are likely to vary due to saturation and temperature
variations and some kind of on-line parameter correction may be required,
at least for accurate torque control. Indirect rotor flux orientation is, above
all, dependent on the open circuit rotor time constant. On the other hand,
the traditional direct stator flux orientation method has a dependency on
the stator resistance estimate. The dependency on the stator resistance is
high at low frequencies but low at high frequencies. Therefore, direct stator
flux orientation usually does not operate well at low frequencies. To overcome
these problems, indirect and direct methods can be combined by using special
algorithms to obtain better accuracy [20], [19] in a wide operating region. If
the flux feedback signal is sufficiently accurate, though, vector control offers
faster and more accurate torque control, higher peak torque at low speeds,
wider speed range and greater stability than Volts-per-Hertz control.

2.5.2 Two Control Strategies for a Grid Connected PWM Con-


verter

Amplitude-Phase Control

Amplitude-phase control is based on the stationary circuit in Fig. 2.7. Con-


trolling the converter voltage amplitude Ug and phase δ provides a method
for controlling the active and reactive power since

18
I g jX in Ug E g = grid™†š†› œ"Rž†Ÿ
+ +
Ig = grid current
Ug Eg jX in I g
− − δ Ig U g = converter voltage
Eg δ = power angle
(a) (b)

Figure 2.7: Amplitude-phase control. The resistance in the system is neglected. (a) Single-
phase diagram. (b) Phasor diagram.

U g Eg δ≈0 Ug Eg
P = 3 sin(δ) ≈ 3 δ (2.1)
Xin Xin
Eg δ≈0 Eg
Q = 3 (Ug − Eg ) cos(δ) ≈ 3 (Ug − Eg ) (2.2)
Xin Xin
Amplitude-phase control has a low dynamic performance [3] and the active
and reactive power are not decoupled since amplitude-phase control is based
on a stationary model and the the grid resistance is not necessarily negligible.

Vector Control
Vector control of a grid connected PWM converter has many similarities to
vector control of an electric machine. In fact, the grid can be modeled as a
synchronous machine with constant frequency and constant magnetization.
A non-measurable grid flux can be introduced in order to fully acknowledge
the similarities between an electric machine and the grid. In space vector
theory, the non-measurable grid flux becomes a space vector that defines a
rotating grid flux orientated reference frame, see Fig. 2.8. The grid flux vector
is aligned with the d-axis in this reference frame, and the grid voltage vector
is aligned with the q-axis. Finding the position of the grid flux vector is
equivalent to finding the position of the grid voltage vector, since there is
only a 90◦ phase angle between them. An accurate field orientation can be
expected since the grid voltage can be measured.
Vector control regulates the length and position of the grid current vector
in the grid flux orientated reference frame. The active and reactive power
are independently controlled with the active current component i gq and the
reactive current component igd since
3 gf gf
pg = e i (2.3)
2 gq gq
3 gf
qg = egf gq igd (2.4)
2
where egq is the grid voltage vector amplitude. The active current component
is generated by an outer direct voltage control loop and the reactive current
reference can be set to zero for a unity power factor. Vector control provides

19
a faster and less oscillative dynamic response than power angle control since
active power and reactive power are controlled independently [3].

q β
ug
e gq ig
ωgf
" ψ gd " d
i gq
i
θgf gd
α (fixed)

Figure 2.8: Rotating two-axis grid flux orientated reference frame.

20
Chapter 3

Stator Flux Orientated Vector Control


of an Induction Machine

The objective of this chapter is to derive controllers for stator flux regulation
and for current control during stator flux orientation. The stator flux con-
troller and the current controller become PI-controllers that are implemented
in the stator flux orientated reference frame. Current and voltage decoupling
terms for the controllers are also described. Transient models of the induction
machine, internal model control [19] and characteristics of stator flux orienta-
tion are described in order to prepare for the design of the controllers. In the
following, all variables refer to the stator side of the induction machine and
motor references are used. The parameters and base values of the induction
machine in Appendix E are used in all graphs. Appendix C describes the per
unit system that is used in this thesis.

3.1 Transient Model of an Induction Machine


The induction machine is the most common of electric machines, mainly ow-
ing to the fact that machines with cage rotor windings are simple and rugged,
and therefore require less maintenance. The induction machine has also a
high torque per inertia ratio and permits fast torque control. Induction ma-
chines used in demanding adjustable speed applications can be expected to
be started, stopped, braked and possibly reversed frequently. A transient
model of the induction machine is therefore required in order to obtain good
modeling accuracy. In the following, two equivalent transient models will be
described. The transient performance is described by using the space vector
theory [36]. It is assumed that the following simplifications are valid

1. Space harmonics of the flux linkage distribution can be neglected.


2. Linear magnetization characteristic or constant saturation.
3. No iron losses in the machine.

21
4. No temperature nor frequency dependency of the resistances and induc-
tances.
Appendix B describes the construction of a space vector as well as coordinate
transformations between stationary and rotating reference frames.

3.1.1 Transient T-Model


Fig. 3.1 shows a transient equivalent circuit of an induction machine. This
model is commonly referred to as the transient T-model since the three induc-
tances form a “T”. The electrical equations of the T-model in space vector
form and in an arbitrary reference frame k are given by
dψks
uks = Rs iks + + jωk ψks (3.1)
dt
k dψkr
0 = R r ir + + j(ωk − ω)ψkr (3.2)
dt
ψks = Ls is + Lm ikr
k
(3.3)
ψkr = Lm iks + Lr ikr (3.4)
In addition, the mechanical subsystem can be written as
3 3
Te = pIm{ψ∗s is } = − pIm{ψ∗r ir } (3.5)
2 2
J dω
= Te − T l (3.6)
p dt
where the latter equation assumes a stiff rotor shaft. The electro-mechanical
torque can be written in several other ways. Only the two expressions that
can be derived directly from (3.1) and (3.2) are included here. Appendix B
derives (3.5) analytically.
The T-model is more complex than necessary for modelling purposes. Since
the induction machine can be represented by two complex state variables,

d d
Rs jωk ψ sk L sσ L rσ j(ωk − ω)ψ rk Rr
i sk + −
dt dt
− + i kr
∼ ∼
dψ sk d dψ kr
usk Lm
dt dt dt
i km

Figure 3.1: T-form transient equivalent circuit in an arbitrary reference frame.

22
three inductances (derivatives) is one too many. It is therefore suitable to
transform the T-model into equivalent but less complex models that only
use two inductances. In the opinion of the author, the equivalent models
with only two inductances are easier to work with analytically and provide a
simpler representation of the induction machine. Two commonly used models
are the Γ-model and the inverse Γ-model [37]. Only the inverse Γ-model will
be described in the following since this model seems to be especially suitable
to work with.

3.1.2 Transient Inverse Γ-Model


The underlying idea of the inverse Γ-model, shown in Fig. 3.2, is that actual
rotor variables are not required. The non-actual rotor variables can therefore
be defined by choosing an arbitrarily stator-to-rotor turns ratio. When using
the inverse Γ-model, a stator-to-rotor turns ratio is chosen so that the rotor
and stator leakage inductances are combined into one equivalent stator leakage
inductance. The model is named for the fact that the two inductances form
an “inverse Γ”. The following turns ratio is introduced in order to derive the
inverse Γ-model
Lm
γ= (3.7)
Lr
This choice of turns ratio gives the following rotor flux linkage for the inverse
Γ-model
ψkR = γ ψkr (3.8)
Although actual rotor variables are not required, the torque of the T- and
the inverse Γ-models must be the same. This implies that the rotor current
of the inverse Γ-model is
1
ikR = ikr (3.9)
γ
The mutual inductance of the inverse Γ-model is defined by
ψkR = γ(Lm iks + Lr ikr ) = γLm (iks + ikR ) = LM (iks + ikR ) (3.10)

d
Rs jωk ψ sk Lσ k
j(ωk − ω)ψ R RR
i sk +
dt
− + i kR
∼ −

dψ sk d dψ kR
u sk LM
dt dt dt
i kM

Figure 3.2: Inverse Γ transient equivalent circuit in an arbitrary reference frame.

23
The stator inductance should not change between the two models, hence, the
leakage inductance becomes
Lσ = Ls − LM = σLs = Lsσ + γLrσ ≈ Lsσ + Lrσ (3.11)
The rotor resistance RR of the inverse Γ-model remains to be determined.
Starting with the rotor voltage equation of the T-form circuit, the rotor re-
sistance is defined by
dψkr dψkR
0 = Rr ikr + + j(ωk − ω)ψkr = γ 2 Rr ikR + + j(ωk − ω)ψkR
dt dt
k
dψ R
= RR ikR + + j(ωk − ω)ψkR (3.12)
dt
In summary, the transformation from the T-model to the inverse Γ-model is
Lm
γ = (3.13)
Lr
ψkR = γ ψkr (3.14)
ikr
ikR = (3.15)
γ
Lσ = σLs = Lsσ + γLrσ (3.16)
LM = γLm (3.17)
RR = γ 2 Rr (3.18)
The stator voltage, stator current, stator resistance and stator inductance
remain the same for both models. This gives the inverse Γ-model in the space
vector form and in an arbitrary reference frame as
dψks
uks = Rs iks + + jωk ψks (3.19)
dt
k dψkR
0 = R R iR + + j(ωk − ω)ψkR (3.20)
dt
ψks = Ls is + LM ikR
k
(3.21)
ψkR = LM (iks + ikR ) = ψks − Lσ iks (3.22)
3 3
Te = pIm{ψ∗s is } = − pIm{ψ∗R iR } (3.23)
2 2

3.2 Internal Model Control


Harnefors [19, pages 47–52] has proposed using internal model control for
current regulation of induction machines. The following text provides a very
short description of internal model control. In the following sections, internal

24
model control will be used for designing current and stator flux controllers.
The structure of internal model control is shown in Fig. 3.3. In this figure, F
is a “classical controller”, C is the so called internal model controller while G
and G b are the plant and plant models, respectively. It is assumed that G is
a 2×2 matrix with no zeros in the right half plane. The closed-loop transfer
function for a perfect plant model, G b = G, becomes

Gc |G=G
b = GC (3.24)
An idea that is not far-fetched would be to choose the optimal internal model
controller C = G−1 , resulting in a closed-loop transfer function of Gc =
I. However, this choice has little practical use since it introduces a high
sensitivity to plant model errors, and the control signals u will be very large.
In addition, implementing the optimal controller C = G−1 requires that the
order of the G−1 denominator is higher than the order of the G−1 nominator,
i.e., G−1 must be proper. Luckily, a proper controller can be obtained by
detuning the optimal controller with a low-pass filter Glp as
 
α1
0
b −1 b −1  p + α1 
C = G Glp = G  α2  (3.25)
0
p + α2
The following controller is obtained if α1 and α2 are chosen equal
µ ¶−1
b −1 C = 1 − α α b−1 α b−1
F = [I − C G] G = G (3.26)
p+α p+α p
b = G, but
The closed-loop transfer function for a perfect plant model, G
detuned optimal controller now becomes
· ¸
b Glp | b =
−1 α 1 0
Gc = G G (3.27)
G=G
p+α 0 1
Eqs. (3.26) and (3.27) are highly interesting since they imply that it is possible
to design a current controller that is directly given by machine parameters

F

y + e + u y
C G
− +

Figure 3.3: Structure of internal model control.

25
and the desired closed-loop system bandwidth α. It is therefore believed that
one of the main advantages of internal model control is the ease with which a
controller can be designed. However, manual tuning or control modifications
may be required since internal model control is dependent on plant model
accuracy.
For a system with no complex pole-pairs, a diagonal controller is obtained
when implementing internal model control. For a system with complex pole-
pairs, such as an induction machine modeled in a flux orientated reference
frame, a non-diagonal controller is obtained. This means that the controller
is attempting to cancel the complex pole pair. It was shown in [19] that
exact cancellation is not possible and the controller can introduce current
oscillations. Decoupled internal model control was proposed as a remedy
for this. Decoupled internal model control implies in essence that the cross
coupling between axes is decoupled by a separate feedback loop. The cross
coupling can therefore be neglected when applying internal model control and
a diagonal controller is obtained.

3.3 Some Characteristics of Stator Flux Orientation


This section describes four important characteristics of stator flux orienta-
tion. These four characteristics are cross coupling between active current and
stator flux, torque pull-out, field weakening operation and low speed opera-
tion. The characteristics of stator flux orientation are also compared with the
characteristics of rotor flux orientation.

3.3.1 Cross Coupling between Active Current and Stator Flux


When performing rotor flux orientated vector control, the torque and rotor
flux are ideally truly decoupled. For stator flux orientated vector control,
though, a cross coupling between the active stator current and stator flux is
present. A decoupler can be used to minimize the cross coupling [16]. Even-
tually, this section derives a stator flux PI-controller and a current decoupler
of slightly simpler design than in [16].
The rotor flux and the rotor current are expressed in terms of stator current
and stator flux in order to investigate the cross coupling between torque and
stator flux

ψ R = ψ s − L σ is (3.28)
µ ¶
ψR ψs Lσ ψs Ls
iR = − is = − 1+ is = − is (3.29)
LM LM LM LM LM
26
Inserting these two equations into the rotor voltage equation (3.20) gives
dψsf
s 1 sf disf Ls sf
+ jωsl ψs + ψs = Lσ s + jωsl isf
sf
s + i (3.30)
dt τr dt τr s
The slip speed ωsl and the open circuit rotor time constant τr are defined by
ωsl = ωsf − ω (3.31)
Lr LM
τr = = (3.32)
Rr RR
Solving (3.30) in steady state gives the following expressions for the reactive
and active stator currents
sf
1 + σ(ωsl τr )2 ψsd
isf
sd = (3.33)
1 + (ωsl στr )2 Ls
sf
(1 − σ)ωsl τr ψsd
isf
sq = (3.34)
1 + (ωsl στr )2 Ls
where the total leakage factor σ is equal to
L2m LM
σ =1− =1− (3.35)
Ls Lr Ls
It can be shown that when combining (3.33) and (3.34), and observing that
s/sp = στr ωsl , the following stator current will be obtained in steady state [38]
sf µ ³ ´¶
sf sf sf ψsd 1 + σ 1 + σ −j2atan ssp
is = isd + jisq = − e (3.36)
Ls 2σ 2σ
This is the equation for the circle diagram of an induction machine. A con-
siderable amount of information can be extracted from the circle [39]. The
diagram should be used with some care though since saturation is not consid-
ered. The reactive current, required to maintain a constant stator flux linkage,
is plotted against the active current for different slip speeds in Fig. 3.4. The
sf
parameters in Appendix E are used and it is assumed that ψsd =1 p.u. The
nominal slip speed of this machine is ωsl,n = −0.012 p.u. It can be observed
that the cross coupling is rather strong and a stator flux controller is recom-
mended to maintain constant stator flux. Open-loop regulation of the flux,
which is possible for rotor orientation [40], is probably not suitable for sta-
tor flux orientation. Stator flux orientation therefore has a slightly higher
complexity than rotor flux orientation.

3.3.2 Torque Pull-Out


It was pointed out in [16] that stator flux orientation has two torque break-
down points. The following derives these two breakdown points.

27
In steady state, or for a case in which the rotor flux is constant during rotor
flux orientation, the rotor voltage equation (3.20) simplifies to

0 = RR irf rf
R + j(ωs − ω)ψR (3.37)
Identifying the imaginary part of (3.37) yields the following relation between
rotor flux and active rotor current.
sωs rf
irf
Rq = − ψ (3.38)
RR Rd
where s is the slip, defined by
ωs − ω
s= (3.39)
ωs
Eqs. (3.23) and (3.38) imply that the steady state electro-mechanical torque
in a rotor flux orientated reference frame can be written as
3 rf rf 3 sωs rf 2 3 sωs L2M rf 2
Te = − pψRd iRq = [steady state] = p (ψ ) = p (isd ) (3.40)
2 2 RR Rd 2 RR
The electro-mechanical torque for rotor flux orientation at a given stator
frequency has hence, ideally, a linear dependency on the slip. For stator flux
orientation, though, it is required to eliminate the rotor flux from (3.40) in
order to derive a similar expression to (3.40). Equation (3.37) can then be
rewritten as
RR Ls rf RR rf
0 = ψR − ψs + jsωs ψrf
R
Lσ LM Lσ
LM rf ¡ ¢
= − ψs + 1 + jsωs στr ψrf
R (3.41)
Ls
Eq. (3.41) gives the following relation between rotor flux and stator flux in
steady state

rf 2 L2M L2M
(ψRd ) = |ψrf
R|
2
= 2 rf 2
|ψ | = 2 ¡ ¢ |ψrf 2
s | (3.42)
Ls |1 + jsωs στr |2 s Ls 1 + (sωs στr ) 2

The steady state torque for stator flux orientation becomes


3 sωs L2M
Te = p ¡ ¢ |ψ s | 2 (3.43)
2 RR L2s 1 + (sωs στr )2

Note that (3.42) has been derived based on the assumption of constant rotor
flux, which is not necessarily true for stator flux orientation. However, (3.43)
implies that stator flux orientation has two extreme values (pull-out torques)

28
2.5
ωslp
2
ωsl=0.04 p.u.
1.5
0.03
1 0.02
Active current (p.u.)
0.5 0.01

0 0

−0.5 −0.01

−1 −0.02
−0.03
−1.5 −0.04

−2 −ωslp

−2.5
0 1 2 3 4 5
Reactive current (p.u.)

Figure 3.4: The circle diagram of the induction machine given in Appendix E, plotted for
stator flux orientation in steady state.

4
Stator flux reg.
3 Rotor flux reg.

1
Torque (p.u.)

−1

−2

−3

−4
−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1
Rotor speed (p.u.)

Figure 3.5: Torque as a function of rotor speed curves for stator flux and rotor flux orientation,
plotted in steady state and below base speed.

29
that are not present in rotor flux orientation. The pull-out slip speeds are
given by
1
sp ω s = ± (3.44)
στr
This implies that the pull-out torques for stator flux orientation are
3 LM
Tep = ± p 2
|ψ s | 2 (3.45)
2 2σLs
The pull-out torque has no dependency on the frequency, thus, peak torque
is provided from zero speed up to base speed. Using numerical values for the
machine given in Appendix E and |ψs |=1 p.u., pull-out torques of ±1.86 p.u.
are obtained. For this stator flux, the peak torque must be set below the
torque pull-out in order to avoid torque breakdown for stator flux orientation.
Fig. 3.5 shows the electro-mechanical torque versus rotor speed curves for
rotor flux and stator flux orientation. The curves are plotted below base
speed. Parameters of the machine given in Appendix E are used and it is
assumed that |ψs |=1 p.u. and |ψR |=LM /Ls p.u.
Note that (3.43)-(3.45) are the same equations that can be derived for an
induction machine directly connected to the grid if the stator resistance can
be neglected. The difference is that the stator flux can be selected arbitrarily
in an adjustable frequency drive.

3.3.3 Operation in Field Weakening Region


The induction machine is both current limited and voltage limited when oper-
ating in the field weakening region. Maximum torque capability requires that
the current limit and voltage limit are properly balanced. It was shown in [18]
that stator flux orientation has good torque capability, i.e., peak torque, when
operating in the field weakening region. Varying the stator flux reference pro-
portional to 1/ω gives nearly optimal torque capability [18]. Furthermore,
stator flux orientation is parameter insensitive in the field weakening region
since the back-emf dominates and the stator resistance voltage drop can nearly
be neglected. For rotor flux orientation, the torque capability is reduced when
selecting the rotor flux reference proportional to 1/ω in the field weakening re-
gion [18]. This is because field weakening operation is governed by the stator
flux limit
|u|max
s
|ψ|max
s ≈ (3.46)
ωbase
The rotor flux becomes too low when selecting the rotor flux reference propor-
tional to 1/ω which gives reduced torque capability at low speeds in the field
weakening region. However, the optimal rotor flux can be calculated from the
optimal stator flux but this calculation can be complicated [18].

30
3.3.4 Low Speed Operation
Stator flux orientation with the conventional open-loop voltage model for sta-
tor flux feedback has poor low speed properties. This is because the stator
resistance voltage drop dominates over the back-emf at low speeds and the
sensitivity to an incorrect stator resistance estimate becomes high. Further-
more, the conventional open-loop voltage model is sensitive to measurement
offset and is often replaced with a delay element. The stator flux feedback
becomes highly erroneous when operating at frequencies that are below the
cut-off frequency of the filter. It seems that the minimum frequency of the
conventional voltage model, and variants thereof, is about 5 Hz [40].

3.4 Stator Flux Orientated Stator Flux Controller


This section derives a stator flux PI-controller with a decoupling term. Eq. (3.30)
is suitable to work with when designing the stator flux controller. Rearrang-
ing (3.30) yields ³ ´
(p + jωsl )στr + 1 Ls
sf
ψs = isf
s (3.47)
τr (p + jωsl ) + 1
For perfect stator flux orientation, the stator flux linkage is aligned with the
sf
d-axis and ψsq = 0. Taking the real part of (3.47) gives

sf Ls ³ sf sf
´
ψsd = (στr p + 1)isd − ωsl στr isq (3.48)
τr p + 1
The influence of the derivative operator in the numerator of (3.48) is probably
minor since the total leakage factor typically varies between 0.05 and up to
0.2. Therefore, the following approximation to (3.48) should be reasonable
sf Ls ³ sf sf
´
ψsd ≈ i − ωsl στr isq (3.49)
τr p + 1 sd
Fig. 3.6 shows the step response of the exact and approximate transfer func-
tions, with parameters from Appendix E and σ = 0.073. The approximative
transfer function is quite accurate, at least for a step response.
Eq. (3.49) shows that the stator flux response time to a step in reactive current
isd is τr , and for no-load conditions the steady state value of the stator flux is
Ls isd . Unfortunately, there is a cross coupling between the active current i sq
and the stator flux ψsd under loaded conditions, which was also shown in
Section 3.3.1. It is, therefore, not trivial to obtain a decoupled control of
the torque and stator flux during stator flux orientation since the stator flux
responds to steps in active and reactive currents with the same time constant.
A current decoupler is normally used to decouple the active current from the

31
−4
x 10
a) 0.025 b) 1

0.02 0.8

Amplitude
0.015 0.6
Amplitude

0.01 0.4

0.005 0.2

0 0
0 1 2 3 4 5 0 0.5 1 1.5 2 2.5
Time (s) Time (s)

Figure 3.6: Comparison of exact, (στr s + 1)/(τr s + 1) (solid), and approximative, 1/(τr s + 1)
(dashed), transfer functions. (a) Step response. (b) Impulse response.

stator flux [16]. The slip speed is rewritten from the imaginary part (3.47) in
order to derive a suitable decoupler expression as
(στr p + 1)Ls isf
sq
ωsl = sf
(3.50)
τr (ψsd − Lσ isf
sd )
It is assumed that the derivative operator of the numerator can be neglected,
i.e., the assumption that was applied to (3.48). Combining (3.49) and (3.50)
gives
à ! à !
sf 2 sf 2
sf Ls Lσ (isq ) Lσ (isq )
ψsd ≈ isf − = G i sf
− (3.51)
τr p + 1 sd ψsd sf
− Lσ isf
sd
sd
ψ sf
sd − L σ i sf
sd
The second term of (3.51) represents the cross coupling between active current
and stator flux. The cross coupling can be compensated with a separate loop,
i.e., a so called current decoupler. The stator flux controller is now derived
by placing the closed-loop pole in the desired bandwidth α, hence
FG α α/p
= = (3.52)
1 + FG p + α 1 + α/p
α α
F = G−1 = (b
τ p + 1) (3.53)
p bsp r
L
The controller becomes a conventional PI-controller directly parameterized in
machine parameters. The proportional gain and the integration time of the
controller are
αb
τr
kp = Ti = τbr (3.54)
bs
L
32
Combining the PI-controller with the current decoupler gives the following
reactive current reference
µ ¶ bσ (isf
1 L sq )2
sf ?
isd = kp 1 + (ψsd − ψbsd ) + sf
sf ? sf
(3.55)
Ti p ψbsd − Lbσ isf
sd

The current decoupler is similar to the one in [16] but the current decoupler
that is developed here is of a slightly simpler design. It should be observed
that the decoupler is sensitive to the leakage inductance Lσ . This is a draw-
back since the magnetic saturation may vary during steps in active current,
i.e., when the decoupler is needed most. The decoupler may also introduce
additional noise from the current sensors.
The selection of the stator flux reference depends on the application and
there are several possibilities for selecting the reference [41],[42], although
a common strategy is to select the stator flux reference equal to its rated
value for speeds lower than base speed and proportional to 1/ω in the field
weakening region [18], where ω is the rotor speed.

3.5 Generation of Active Current Reference


The active current reference is generated from an outer control loop, which
often is a speed control loop. The design of a speed controller will not be
covered here, since it would depend on the load characteristics. See [43, pages
89–90], [19, pages 76–78] for details concerning speed controllers. Instead, this
section discusses how to generate correct torque during varying stator flux.
If a current decoupler is not used, or if a considerably detuned current decou-
pler is used, one can expect undesirable stator flux variations. Furthermore,
it is desirable to vary the flux in the field weaking region and also when max-
imizing the efficiency [35, pages 388-396]. The following is often applied in
order to account for stator flux variations and to obtain correct torque
2 Te?
isf ?
sq = (3.56)
3p ψbsd
sf

Thus, an outer control loop gives a torque reference and the active current
reference is given by dividing the torque reference by the estimated stator
flux. This method ideally ensures that the requested torque is obtained, but
of course only when the stator flux feedback is correct. The importance of
(3.56) should not be exaggerated, however. There is no need to introduce the
additional, possibly a bit computational expensive, division in (3.56) if the
stator flux can be considered to be fairly constant.

33
3.6 Stator Flux Control Without Reactive Current Con-
trol
A stator flux controller without reactive current control was presented in [44].
The controller is mentioned here since its design differs significantly from the
stator flux controller in the previous section. In the paper, the stator flux
controller was derived directly from the stator voltage equation as
µ ¶
sf ? 1 sf ? sf
usd = kp 1 + (ψsd − ψsd ) (3.57)
Ti p
This controller does not suffer if the magnetizing curve of the machine is
strongly non-linear, which may be troublesome for the stator flux controller
described above. In the paper [44], dead-beat gains were used and the ex-
perimental results were excellent. However, over-current protection is more
complicated to implement since the controller has no reactive current control.

3.7 Stator Flux Orientated Current Controller


The objective of this section is to derive a synchronous PI-controller for sta-
tor flux orientation. A suitable model of the induction machine is required
when designing a current controller. Combining (3.19) and (3.20) yields the
following equation
dis
−usf sf sf
s + Rs is + jωsf ψs + Lσ − RR isf sf
R − jωsl ψR = 0 (3.58)
dt
The rotor flux and rotor current in (3.58) should be eliminated when designing
a stator flux orientated current controller for an induction machine with cage
rotor windings. These variables can be substituted by using (3.28) and (3.29).
Equation (3.58) can now be written, with some minor modifications, in the
following way that is useful for the design of a stator flux orientated current
controller
disf Ls sf 1 sf
Lσ s + jωsl Lσ isf
s + i = u sf
− jω ψ sf
+ ψ (3.59)
dt τr s s s
τr s
It is convenient to exclude some terms from (3.59) before continuing the design
of the current controller. It is possible to exclude three terms from (3.59). The
first term that can be neglected is ψsfs /τr . This term is ideally constant during
stator flux orientation so it can be compensated by intregator action. The
second term that can be neglected is jωsl Lσ isf s . This term has been observed
to be weak because the slip speed is normally low. However, some schemes
use a separate decoupler for this term, which may be relevant for applications

34
where the slip speed is high and changes frequently. The third term that can
be neglected is the induced emf jω ψsfs . This term can be compensated by
the controller integrator based on an assumption of slowly varying stator flux
and slowly varying rotor speed, seen from the electric time perspective. On
the other hand, it is easy to design a voltage decoupler for the induced emf
if the assumption of slowly varying rotor speed does not apply. Nevertheless,
the following is possible to write when excluding these three terms
1
isf
s ≈ usf sf
s = Gus (3.60)
Ls
pLσ +
τr
Internal model control can be applied for the design of the current controller.
Placing the closed-loop pole in the desired bandwidth α yields
FG α α/p
= = (3.61)
1 + FG p + α 1 + α/p
α −1 b
F = b σ + α Ls
G = αL (3.62)
p τbr p
The controller F is an ordinary PI-controller. The reference voltage vector
becomes
µ ¶
sf ? 1
us = kp 1 + (isf ? sf
s − is ) (3.63)
Ti p
kp = α Lb σ Ti = σbτbr (3.64)

The term jω ψsf


s can be added to the reference voltage if one wishes to use a
voltage decoupler for the induced emf.

35
36
Chapter 4

Stator Flux Observers

This chapter describes some different stator flux observers. A stator flux ob-
server is required for vector control in order to derive stator flux feedback
without measuring the flux directly. The stringent meaning of the word ob-
server is an estimator that uses both inputs to integration models and feedback
control for error correction. However, the word observer is also frequently used
for an “observer” that is based on integration models only, although a more
correct term for such an “observer” would be a real-time simulation. For
convenience, observers with feedback are in this chapter referred to as closed-
loop observers while observers that only employ process models are referred
to as open-loop observers. Only directly parametrized closed-loop observers
are described in this thesis. A directly parametrized stator flux observer has
either a constant gain matrix or else the gain matrix is given by a simple
function of rotor speed.

4.1 Open-Loop Stator Flux Observers


4.1.1 Voltage Model
The voltage model is a common open-loop stator flux observer. The model
is based on direct integration of the stator equation in the stator orientated
reference frame, given by
Z t
s
bs =
ψ bs iss )dt + ψ
(uss − R b ss (0) (4.1)
0
q
|ψ s
b s| = (ψbsα
s )2 + ( ψbs )2
sβ (4.2)
ψbsα
s ψbsβ
s
b
cos(θsf ) = b
, sin(θsf ) = (4.3)
b s|
|ψ b s|

The voltage model is a direct flux observer since the stator flux position is
derived directly from the stator flux estimate. The commonly referred draw-
back of the voltage model is its high sensitivity to an error in the estimated

37
stator resistance at low speeds, which is illustrated in Fig. 4.1. Expressions of
the relative magnitude error and the phase shift are given in Appendix D. At
low frequencies, the resistive voltage drop dominates so that even the slightest
error in stator resistance estimation may be disastrous. Otherwise the voltage
model is not dependent on any other parameter and at high speeds the pa-
rameter sensitivity is nearly non-existent. However, this observer can seldom
be directly implemented even for a correct estimation of the stator resistance.
The voltage model has namely two additional disadvantages due to direct in-
tegration. These two disadvantages are, first, that the stator flux estimation
will drift for biased input signals and, second, the model will not converge to
b ss (0) is incorrect. To overcome these
correct stator flux if the initial value ψ
two problems, the voltage model is often implemented as a delay element,
introducing other problems such as magnitude scaling and phase shift. A
fourth drawback of this model is the difficulty to measure the stator voltage
uss . This drawback is due to the fact that the voltage changes very rapidly
when using pulse width modulation, and aliasing is likely to occur. The alias-
ing problem may be especially problematic during low-speed operation since
the voltage pulses become very narrow. An entire pulse may be missed or
extended with the sampling period time when sampling the narrow pulses.
According to [44], the aliasing problem is normally solved by using a high
order analog low-pass filter or by using the stator voltage reference instead
of the measured stator voltage. Both of these two methods have drawbacks.
The bandwidth of the observer is limited when using a low-pass filter, and
when using the reference voltage, an accurate compensation of the blanking
time and the voltage drop of the semiconductor valves is required. As an
alternative to these two methods, a partly implemented analogue integrator

a) 1.1 b) 20

1 15
Relative magnitude error

Phase shift (deg.)

0.9 10

0.8 5

0.7 0

0.6 −5
0.001 0.01 0.1 1 23 0.001 0.01 0.1 1 23
Rotor speed (p.u.) Rotor speed (p.u.)

Figure 4.1: Relative magnitude error and phase shift of the voltage model at rated load and
bs = 1.5Rs .
R

38
was presented in [44].

4.1.2 Improved Integrator Algorithms for the Voltage Model


Hu et al. [45] have presented three improved integrator algorithms for the
voltage model. The algorithms are equivalent to direct integration but nearly
eliminate the problems with bias and incorrect initial value. The underlying
idea of the algorithms are shown in Fig. 4.2. This block circuit, referred to
as algorithm 1, is equivalent to an integrator since the transfer function from
bs iss ) to ψ
(uss − R b ss becomes (neglecting the saturation block)

bs iss
uss − R 1 uss − Rbs iss
b ss =
ψ = (4.4)
p + ω c 1 − ωc p
p + ωc
The benefit of this algorithm is that the low-frequency gain is decreased and
the feedback path of ψ b ss , allows the use of a saturation block. The saturation
block is especially useful when the voltage model is combined with a stator flux
controller. The stator flux controller maintains constant stator flux and the
amplitude of the stator flux estimate should not deviate significantly from the
reference. A second integrator algorithm, algorithm 2, is shown in Fig. 4.3.
The limiter for this algorithm is placed in polar coordinates. Algorithm 1
and 2 require that the limiters in the feedback path are set equal to the
actual stator flux amplitude. Magnitude scaling and phase shift appear in
the stator flux estimate if the limiters are incorrectly set. A third integrator
algorithm, algorithm 3, is shown in Fig. 4.4. This scheme uses a form of
adaptive magnitude compensation and is therefore not dependent on correctly
set limiters. The principle of algorithm 3 can be understood by investigating
the error e in Fig. 4.4
µ ¶ ³
1 dψ sα dψ sβ 1
e = ψbsα + ψbsβ = b s | sin(ωsf t)ωsf |ψs | cos(ωsf t)

b s|
|ψ dt dt b s|

´
−|ψb s | cos(ωsf t)ωsf |ψs | sin(ωsf t) = 0 (4.5)

Note that this equation is only equal to zero for unbiased conditions. Other-
wise, the PI-controller adjusts the magnitude compensation in order to sup-
press bias and force the observer to reach correct stator flux, even though the
initial value may be incorrect. During simulations, the author has encoun-
tered problems when tuning the PI-controller. Instabilities have occurred un-
der certain operating conditions. In the opinion of the author, the improved
integrator algorithms 1 and 2 used together with a stator flux controller are
to be preferred.

39
Saturation
ωc [ψˆ ssα ]max
min
ψ
ˆ ssα,fb p + ωc

u ssα − R̂ s i ssα 1 ψˆ ssα,lp + ψˆ ssα


p + ωc +

u ssβ − R̂ s i ssβ 1 ψ
ˆ ssβ,lp + ψˆ ssβ
p + ωc +

ψˆ ssβ,fb ωc [ψˆ ssβ ]max


min
p + ωc
Saturation

Figure 4.2: Improved integrator algorithm 1 with saturation of the feedback.

u ssα − R̂ s i ssα 1 ψˆ ssα,lp + ψˆ ssα


p + ωc + Saturation
ψ
ψ̂ cmp ψ
ψ̂ s
ψ
ˆ ssα,fb ωc
p + ωc ¥¡ Cartesian
¢  £¤¡
Cartesian

¢  £¤¡ ¥¡
ωc θ̂sf
ψ
ˆ ssβ,fb p + ωc

u ssβ − R̂ s i ssβ 1 ψˆ ssβ,lp + ψˆ ssβ


p + ωc +

Figure 4.3: Improved integrator algorithm 2 with limited amplitude.

dψ ssα dψ ssβ
ψˆ ssα +ψ
ˆ ssβ
dt dt

1 ψˆ ssα,lp + ψˆ ssα
emf ssα p + ωc +
ψ
ψ̂ cmp e ψ
ψ̂ s
ψ
ˆ ssα,fb ωc ×
PI
p + ωc ¦§ ÷
©¨ ª«§
Cartesian

Cartesian

©¨ ª«§
ωc ¦§
ψ
ˆ ssβ,fb p + ωc
emfssβ 1 ψˆ ssβ,lp + ψˆ ssβ
p + ωc +

Figure 4.4: Improved integrator algorithm 3 with adaptive amplitude compensation.

40
4.1.3 Current Model
The current model is based on the rotor voltage equation (3.20). Eliminating
the rotor current gives the rotor equation in stator coordinates, ω k = 0, as
the following
dψsf
r ψsf
+ r − jω ψsf sf
r − R R is = 0 (4.6)
dt τr
The current model for stator flux orientation in the stator orientated reference
frame can be written as
bM
L
b sR =
ψ iss (4.7)
τbr p + 1 − jωbτr
b ss
ψ b sR + L
= ψ bσ iss (4.8)
The transfer function has a complex pole whose imaginary part increases with
increasing rotor speed. The model will therefore be poorly damped at high
speeds when implemented in the stator orientated reference frame. Luckily,
the poor damping can be eliminated by transforming the current model from
the stator to a rotating reference frame. In the rotor orientated reference
frame, the current model for stator flux orientation becomes
bM
L
b rR =
ψ irs (4.9)
τbr p + 1
b rs
ψ b rR + L
= ψ bσ irs (4.10)
The current model introduces a parameter sensitivity to the inductances of
the machine. Note that the current model is commonly implemented in the
rotor flux orientated reference frame for indirect rotor flux orientation. The
rotor flux position is, in that case, derived indirectly from the integration of
the sum of the rotor speed and the slip speed.

41
4.2 Closed-Loop Stator Flux Observers
Two closed-loop flux observers will be investigated in this section. Both ob-
servers are directly parametrized. The term directly parametrized refers to
that that the observer gain matrix is mainly given by the induction machine
parameters and either constant or rather simple to calculate repeatedly in each
sample. A directly parametrized observer is therefore normally less computa-
tionally expensive, which may be important for time critical applications or
for slow microcontrollers. The first observer [20] is of Gopinath type that has
been acknowledged as a high performance observer at both low and high rotor
speeds. The second observer is a Luenberger observer with nearly constant
poles.

4.2.1 Current and Voltage Model


Two high performance closed-loop flux observers were presented in [46] and [20].
The observers attempt to combine the best properties of the current and volt-
age models and will therefore be referred to as the current and voltage model.
The observer in [20], shown in Fig. 4.5, was proposed for observing stator flux
and [46], shown in Fig. 4.6, was proposed for observing rotor flux. In fact, ei-
ther one of these two observers can be used for observing stator flux. Fig. 4.7
shows a simple block rearrangement which demonstrates that the observers
are identical for stator flux observation.
The observers shown in Figs. 4.5 and 4.6 are combinations of the current and
voltage models. In essence, the current model corrects the voltage model at
low frequencies while at high frequencies the voltage model is mainly used.
The transition frequency between the two operation modes are governed by
the gains K1 and K2 . The matrix notation for the observer can be written as
h iT h iT
s
bs ψ s s s
b = ψ
x u = u s is (4.11)
db
x ¡ s ¢
= Ab b +K ψ
x + Bu b ss
b sC − ψ (4.12)
dt
ψb ss = C x
b (4.13)

where ψb ssC is the stator flux estimate from the current model and ψ
b ss is the
stator flux estimate from the voltage model. The matrices A, B, b C and K
can be obtained directly from Fig. 4.5 as
· ¸ · ¸ h i · ¸
0 0 1 − Rbs K
A= b =
B C= 1 0 K= 1
(4.14)
0 1 0 0 K 2

It was proposed in [46] to use the following observer gains


K1 = k 1 + k 2 , K 2 = k 1 k 2 (4.15)

42
R̂ s

u ss
+
L̂σ
ir ˆM
L ψ̂ rR
ψ ψ̂sRC+
ψ ψ
ψ̂ ssC + + 1 ψ
ψ̂ ss
− jθ s
i ss e pτr + 1 e jθ +
K1
+
ˆ − + p

¬@­'® ® ¯°_±!²‰³_´_¯µ¶ °‚± ·_¯n® ³± ³'®


³_® ¶ ¯°_± ¸± ¯´‚® ¯¹ ¯® ¯°'º¯n¹ ® ¸²M¯R»
θ
1 ψs
K2
p
Voltage
model
+ 1 ψ
ψ̂sR
i ss R̂ R
− p
1
− jω
τˆ r
¬@­_® ® ¯°_±²¼³'´_¯µ¶ °‚± ·_¯n½$± ¸± ³_®
ω
³_® ¶ ¯°_± ¸± ¯´‚® ¯¹ ¯® ¯°'º¯n¹ ® ¸²M¯R»

Figure 4.5: Current and voltage model proposed for stator flux orientation.

R̂ s

u ss L̂σ
+

i sr ˆM
L ψ
ψ̂ rR ψ
ψ̂ sRC + + 1 ψ
ψ̂ ss − ψ
ψ̂ sR
i ss e − jθ e jθ −
K1
+ +
pτˆ r + 1 p +

θ ¾@¿'À À ÁÂ_ÃÄMÅ_Æ_ÁÇÈ Â‚Ã É_ÁnÀ Å_à Å_À ψs


ÅÀ È ÁÂ_à Êà ÁƂÀ ÁË ÁÀ ÁÂ_ÌÁnË À ÊÄMÁRÍ 1
K2
p
Î Å_Ç Ã ÊÏ_ÁnÄMÅ_Æ_ÁÇ

Figure 4.6: Current and voltage model proposed for rotor flux orientation.

i ss

L̂σ
ψ
ψ̂ sRC + ψ
ψ̂ssC + ψ
ψ̂ ss ψ
ψ̂ sRC + ψ
ψ̂ss
subsystem subsystem
+ − −
ψ
ψ̂ sR

i ss L̂ σ
+

(a) (b)

Figure 4.7: Block rearrangement which demonstrates that the observers in Fig 4.5 and Fig 4.6
are identical for stator flux observation. (a) Fig 4.5 redrawn. (b) Fig 4.6 redrawn.

43
The observer poles are given by
© ¡ ¢ª
det λI − A − KC = λ2 + K1 λ + K2 = 0 (4.16)
and become
λ1 = −k1 , λ2 = −k2 (4.17)
Hence, the observer poles are not dependent on rotor speed, well damped and
directly given by the observer gain matrix. Typical poles are 1-10 Hz [40].
The current and voltage model has, in fact, a third pole which is associated
b ssC . The location of this pole
with the current model that is used to obtain ψ
is dependent on whether the current model is implemented in the stator or
the rotor orientated reference frames. As shown in Section 4.1.3, a complex
pole whose imaginary part varies with rotor speed is obtained when the cur-
rent model is implemented in the stator orientated reference. The complex
pole may introduce numerical problems when transforming the observer into
discrete form. The complex pole can be eliminated by implementing the
current model in the rotor orientated reference frame. The following poles
are obtained when implementing the current model in the rotor orientated
reference frame
1
λs1 = −k1 , λs2 = −k2 , λr3 = − (4.18)
τbr
Placing the current model in the rotor orientated reference frame hence cancels
out the typically poorly damped rotor pole of an induction machine.
The current and voltage model’s sensitivity to parameter errors can be inves-
tigated by studying the relative magnitude error and the phase shift. Expres-
sions for studying the parameter sensitivity are given in Appendix D. Fig. 4.8
illustrates the parameter sensitivity at rated slip speed and k 1 = k2 = 3 Hz.
The parameter sensitivity is generally low, except for high sensitivities to the
rotor resistance and the mutual inductance.

4.2.2 Luenberger Observer with Nearly Constant Poles


This section designs poles for a Luenberger observer intended for observing
stator current and stator flux. The poles of the Luenberger observer imitate
the poles of the induction machine but are better damped. The observer gain
matrix is directly parametrized in induction machine parameters. One of the
observer poles is constant, while the other pole varies with rotor speed in order
to allow a good trade off between tracking ability and noise sensitivity. The
observer gain matrix may seem complicated and computationally expensive
to derive at a first glance. However, most of the observer gain matrix becomes
constants that only need to be calculated once. Furthermore, the poles of the

44
1.6 20

1.5 ^R =1.5R
r r 15
1.4
10 ^ ^
Relative magnitude error
1.3 Lrσ=1.5Lrσ Lsσ=1.5Lsσ

Phase shift (deg.)


5
1.2 ^ ^
Lm=1.5Lm Lsσ=1.5Lsσ
1.1 0
^
1 Rs=1.5Rs
−5
0.9 ^
Lrσ=1.5Lrσ −10
0.8 ^ =1.5L
L
^ m m ^ =1.5R
R
Rs=1.5Rs −15 r r
0.7

0.6 −20
0.001 0.01 0.1 0.4 1 2 3 0.001 0.01 0.1 0.4 1 2 3
Rotor speed (p.u.) Rotor speed (p.u.)

Figure 4.8: Parameter sensitivity of the current and voltage model at rated slip speed when
k1 = k2 = 0.06 p.u.

observer are well damped so the complexity of the observer does not increase
when transforming it to discrete form. Nevertheless, the proposed observer
seems more complex than the Luenberger with constant gains in [43].
The following state space system has been chosen in order to design the
Luenberger observer
h iT
b = biss
x b ss
ψ u = uss (4.19)
db
x b + K(is − bis )
b x + Bu
= Ab s s (4.20)
dt
bis = C
bxb (4.21)
s

b B,
The matrices A, b Cb and the observer gain matrix K are
 Ã ! Ã !  
1 b
Ls 1 b
RR 1
 −b b b
Rs + RR + jω − jω 
b
A =   L σ Lb M
b
L σ
b
L M , Bb = b 
 Lσ 

−Rbs 0 1
h i · ¸
k1
C = 1 0 , K= (4.22)
k2

It has been found that the following observer gain matrix allows for a free

45
choice of observer poles
 Ã ! 
1 b
 −b R b R Ls
bs + R − λsc − λsf + jω 
 Lσ LbM 
 
K=
 b
λsc λsf Lσ 
 (4.23)
 b
− Rs + 
 RbR 
− jω
bM
L
The poles of the observer are placed in λsc and λsf related to stator current and
stator flux observation, respectively. The observer gains are complex, which
requires that the complex parts of the observer gain matrix K changes sign
when the rotor speed direction is reversed. The observer becomes unstable
if the complex parts of the observer gain matrix do not change sign during
a speed reversal [43]. Since the sign of the complex parts are governed by
the direction of the rotor speed, this requirement should be automatically
fulfilled.
It is suitable to derive the poles of the induction machine when selecting the
observer poles. When using stator current and stator flux as states, the poles
in the stator orientated reference frame are given by
Ls RR + LM Rs − jωLM Lσ 1 h
λ1,2 = − ∓ − 4LM Lσ Rs (RR − jωLM )
2LM Lσ 2LM Lσ
i1
2 2
+(Ls RR + LM Rs − jωLM Lσ ) (4.24)

Fig. 4.9 shows a plot of the induction machine poles, plotted in per unit. The
pole related to the stator current λ1 is poorly damped at high rotor speeds.
Furthermore, the settling time of the pole related to the stator flux λ 2 varies
a lot since its real part slides from nearly zero to -0.05 p.u. It can be observed
that the real parts of the poles vary in specific regions when the rotor speed
varies. The following derives the upper and lower limits of these regions. At
high rotor speeds, the poles converge at
Ls RR + L M Rs
λ∞
Re,1,2 = lim Re{λ1,2 } = − (4.25)
”ω→∞” 2LM Lσ
At zero rotor speed the poles become
p
Ls RR + L M Rs ± −4LM Lσ Rs RR + (Ls RR + LM Rs )2
λ01,2 = lim λ1,2 = −
ω→0 2LM Lσ
(4.26)
0
It it clear that the fastest dynamics of the stator current is given by λ 1 . Hence,

46
setting λsc equal to
p
Ls RR + L M Rs + −4LM Lσ Rs RR + (Ls RR + LM Rs )2
λsc = −ksc λ01 = −ksc
2LM Lσ
ksc > 1 (4.27)
ensures that observer estimation error decreases faster than induction machine
dynamics for all rotor speeds. The value of ksc determines how much faster
the observer is compared to the induction machine at zero rotor speed.
One may select the pole related to the stator flux observation λsf equal to
the fastest dynamics of the stator flux, i.e.
λsf ∼ λ∞
Re,2 (4.28)
However, for this pole the observer dynamics become much faster than the
induction at low rotor speeds and the observer becomes sensitive to noise.
Instead, the following pole selection is suitable in order to avoid faster observer
dynamics than necessary.
µ ¶
|ω| ∞ |ω|
λsf ∼ λRe,2 + 1 − λ02 , −ωn < ω < ωn (4.29)
ωn ωn
Eq. (4.29) means that the pole related to the stator flux observation varies
with rotor speed. More specifically, the pole varies between the maximum and
minimum real part of the stator flux pole. This way the observer dynamics
are never faster than necessary. Hence, in the range of |ω| ≤ ω n it is proposed
to select λsf as
µ ¶
|ω| p
Ls RR + L M Rs − 1 − −4LM Lσ Rs RR + (Ls RR + LM Rs )2
ωn
λsf = −ksf ,
2LM Lσ
ksf > 1, |ω| ≤ ωn (4.30)
where the constant ksf determines how much faster the observer dynamics are
compared to the induction machine dynamics at nominal frequency. There is
no need to increase the dynamic response further above nominal frequency,
as the induction machine dynamics do not increase. Hence, for rotor speeds
above nominal angular frequency it is proposed to select
Ls RR + L M Rs
λsf = −ksf , ksf > 1, |ω| > ωn (4.31)
2LM Lσ
Fig. 4.9b shows the observer poles in per unit for ksc = 1.5 and ksf = 1.1.
The differences between the observer poles and the induction machine poles
in Fig. 4.9a are:

47
a) 1 b) 1
ω=1

0.5 ω=0.5 0.5


λsc λsf
ω=0
λ1 λ2

Im{λsc,sf}
Im{λ1,2}

ω=±0.5
0 ω=0 0 ω=±1
ω=0
ω=±1

−0.5 ω=−0.5 −0.5

ω=−1
−1 −1
−0.2 −0.15 −0.1 −0.05 0 −0.2 −0.15 −0.1 −0.05 0
Re{λ1,2} Re{λsc,sf}

Figure 4.9: (a) Root loci of the induction machine. (b) Root loci of the observer poles when
ksc = 1.5 and ksf = 1.1.

• The observer improves the damping of the pole related to the stator
current. Furthermore, the observer dynamics for λsc are much faster
than the induction machine dynamics λ1 .
• The observer improves the damping of the pole related to the stator
flux, although the difference is small. The observer pole λsf varies nearly
proportional to the induction machine pole λ2 in order to obtain good
tracking ability but avoid dynamics that are too fast at low rotor speeds.

Fig. 4.10 shows the parameter sensitivity of the observer if ksc = 1.5 and
ksf = 1.1. The sensitivity is studied at nominal slip speed (nominal torque)
and the parameters of the induction machine in Appendix E are used. Equa-
tions for studying the relative magnitude scaling and phase shift are given
in Appendix D. As shown, the observer generally has a low sensitivity to
parameter errors except for a very high sensitivity to the stator resistance
near zero rotor speed.
It is not necessary to use the complicated expressions described in this section
in order to derive the observer poles and ultimately the observer gain matrix.
A simpler method is to plot the root locus of the induction machine, select
the desired observer poles and the variation of the stator flux observer pole,
and then calculate the observer gain matrix with (4.23).

48
1.6 20

1.5
15 ^L =1.5L
^ rσ rσ
1.4 R =1.5R ^
r r Rs=1.5Rs
10
Relative magnitude error

1.3 ^
Lm=1.5Lm ^
Phase shift (deg.)

5 Rr=1.5Rr
1.2
ω=1.1ω
^
1.1 0

1
−5 ω=1.1ω
^
^ ^
0.9 Rs=1.5Rs Lsσ=1.5Lsσ
−10 ^
0.8 Lm=1.5Lm
^L =1.5L
sσ sσ
^ −15
0.7 Lrσ=1.5Lrσ
0.6 −20
0.001 0.01 0.1 0.4 1 2 3 0.001 0.01 0.1 0.4 1 2 3
Rotor speed (p.u.) Rotor speed (p.u.)

Figure 4.10: Parameter sensitivity of the developed Luenberger observer variant with ksc =
1.5 and ksf = 1.1. (left) Relative magnitude error. (right) Phase shift.

49
4.3 Euler Forward Discretization
A flux observer and a PI-controller must be written on discrete form for imple-
mentation in a digital signal processor. It is suitable to choose a discretization
method that allows simple on-line updating of rotor speed and induction ma-
chine parameters. This is because the induction machine is non-linear with
respect to rotor speed and because the parameters of the induction machine
vary for different operating conditions. The Euler forward algorithm enables
simple discretization and on-line updating of induction machine parameters.
The Euler forward algorithm substitutes the derivative operator with
d z−1
p= À (4.32)
dt T
The Euler forward algorithm is an approximative discretization method but
provides an accurate discretization if the sampling frequency is at least ten
times above the underlying system dynamics [19]. Furthermore, one must
consider the stabilty region of the Euler forward method [19], [47] since the
Euler forward method requires well damped poles in order to perform well.
The stability region mapping, given by (4.32), of the unit circle in the discrete
time domain onto the Laplace domain is shown in Fig. 4.11. The unit circle
is mapped onto a circle with a center in −1/T and a radius of 1/T . Fig. 4.11
also shows the typical speed dependent poles of the induction machine, given
in the stator orientated reference frame. The imaginary part of the pole that
is associated with the rotor λsr becomes large at high rotor speeds. As such,
the pole may be located close or outside the Euler forward stability limit
and may result in an oscillatory or unstable response. Other discretization
methods must therefore be used when simulating an induction machine, or
other systems with poorly damped poles. Possible discretization method can-
didates for such systems are the trapezoidal method, Euler backward or Euler
symmetric [47].

z −1
T

Im{z} ω Im{s}

λsr

−1 stable 1 Re{z} 2 stable 1 λss Re{s}


− −
T T

unstable unstable

Figure 4.11: Mapping and stability region of the Euler forward method

50
Chapter 5

Current Control and Control of the


DC-Voltage for a Grid Connected
PWM Converter

5.1 Introduction
Vector control of a grid-connected converter is more or less equivalent to
vector current control of a three-phase electric machine. During balanced
conditions the grid can be modeled as a synchronous machine with constant
frequency and constant magnetization. A non-measurable grid flux can be
introduced in order to fully acknowledge the similarities between an electric
machine and the grid. In space vector theory, the non-measurable grid flux
becomes a space vector that defines a rotating grid flux orientated reference
frame. The grid flux vector is aligned with the d-axis in this reference frame,
and the grid voltage vector is aligned with the q-axis. Finding the position of
the grid flux orientated reference frame is equivalent to finding the position
of the grid voltage vector, since there is only a 90◦ phase angle between them.
The grid voltage can be measured, so that an accurate flux orientation can
be expected.
The control system for the vector controlled grid-connected converter here
consists of two series connected PI-controllers, i.e., a cascaded control sys-
tem. The inner control loop controls the active and the reactive grid current
components. The outer control loop determines the active current reference
by controlling the direct voltage. The reactive current reference for a grid
connected converter can be set to zero which gives the grid connected PWM
converter a unity power factor.

5.2 Synchronization to the Grid


Vector control of a grid connected converter follows the same underlying ideas
as vector control of an electric machine. The grid currents are controlled in a

51
rotating two-axis grid flux orientated reference frame. In this reference frame,
the real part of the current corresponds to reactive power while the imaginary
part of the current corresponds to active power. The reactive and active power
can therefore be controlled independently since the current components are
orthogonal.
Accurate field orientation for a grid connected converter becomes simple since
the grid flux position can be derived from the measurable grid voltages. The
grid flux position is given by
egβ egα
cos(θgf ) = , sin(θgf ) = − (5.1)
|eg | |eg |
The synchronization method described above works well and produces a
nearly sinusoidal grid current if the grid voltage is a symmetrical three-phase
system. However, in some cases the grid voltage may be non-symmetrical
and/or distorted. The non-symmetry/distorsion affects the grid current,
which becomes non-sinusoidal. There are some compensation algorithms for
non-symmetrical grid voltage. An adaptive filter was used in [48] in order to
compensate for notches in the grid voltage. Reference [49] has presented a
compensation algorithm for a grid voltage that contains a negative sequence
component.

5.3 DC-Voltage Controller


The following derivation of a direct voltage controller assumes instantaneously
impressed grid currents and perfect grid flux orientation. The references that
are used in this section are shown in Fig. 5.1. Motor references are used for
both the grid and the induction machine.
The instantaneous apparent power flowing into the grid can be written as
3 3 gf gf 3 gf
Sg = Pg + jQg = egf gf ∗
g ig = (egf gf gf
gq igq + jegq igd ) = (|eg |igq + j|eg |igd ) (5.2)
2 2 2
The active power is the real part of (5.2), thus
3
Pg = |eg |igf
gq (5.3)
2
ÐMÑ,ÒÓ|Ô3ÒÓ)Õ
u g1 PWM
Pg , Q g Ps 3Ô Ò"Machine
Ó)ՂÖn×Ø Induction
eg1 L in R in converter
i g1 i dc,g i dc,s converter machine
~ is1
i dc
~ u dc IM
~

Figure 5.1: Definition of references for the grid connected converter.

52
When neglecting the capacitor leakage, the direct voltage link power is given
by
dudc
Pdc = udc idc = udc C (5.4)
dt
Assuming that the converter losses can be neglected, the power balance in
the direct voltage link system can be written as
dudc 3
udc C = −Pg − Ps = − |eg |igf
gq − Ps (5.5)
dt 2
where Ps is the induction machine power, assumed to be independent of the
direct voltage. A transfer function between the direct voltage and active grid
current Igq can be obtained as
3|eg | gf
udc ≈ − i (5.6)
p2Cudc gq
The transfer function is non-linear. However, it seems reasonable to substitute
the direct voltage with its reference value since the objective is to maintain
a constant direct voltage. This assumption gives the following linearized
transfer function
3|eg | gf
udc ≈ − ? igq = −Gigf
gq (5.7)
p2Cudc
where the approximation is valid for small variations in the direct voltage.
Note also that the grid voltage amplitude can be assumed constant during
normal operation. Applying internal model control gives the direct voltage
link controller as
α −1 2Cu?dc
F = G = −α (5.8)
p 3|eg |
As shown in (5.8), a P-controller is obtained for regulating the direct voltage.
The P-controller is optimal for an integrator process in the sense that the
P-controller eliminates the remaining error for steps in the reference value.
However, there will be a remaining error when the induction machine is loaded
and active power flows between the direct voltage link and the machine. The
remaining error can be eliminated by adding an integrator to the direct voltage
link controller. The following is often adapted for selecting the controller
integration time in traditional PI-controller design
10 10
Ti = ≈ (5.9)
ωc α
where ωc is the cross-over frequency. The latter approximation in (5.9) is valid
since the cross-over frequency is nearly the same as the closed-loop bandwidth
for a first-order system [19].

53
The above derived direct voltage link controller is of simple design and its
stand-alone capability for maintaining constant direct voltage during heavy
electrical transients is moderate. A feed forward term is suitable to introduce
in order to increase the dynamic behavior of the controller [4]. The feed-
forward term is not required if the stator power can be considered constant
from the electrical time perspective but is useful if the opposite is true. The
stator power must be derived in order to derive the feed-forward term. The
instantaneous stator power is given by
3 3 sf sf
Ps = Re{usf i
s s
sf ∗
} = (usd isd + usf sf
sq isq ) (5.10)
2 2
The direct voltage link controller maintains a constant direct voltage by main-
taining power balance between the induction machine side and the grid side.
The constraint for constant direct voltage link voltage is that the instanta-
neous power from the grid Pg is equal to the instantaneous stator power Ps ,
hence
−Pg = Ps (5.11)
The feed forward term becomes
usf sf sf sf
sd isd + usq isq
igf
gq,f f =− (5.12)
egf
gq
Note that converter losses are neglected so the feed forward term may actually
be too high or too low depending on the power flow direction. Regardless of
this, the active reference current of the grid side converter can be written as
µ ¶
1
gf ?
igq = kp 1 + (u?dc − udc ) + igf
gq,f f (5.13)
Ti p
2Cu?dc 10
kp = −α gf Ti = (5.14)
3egq α
The negative proportional gain is because the motor references are used for the
grid. A block diagram that represents the direct voltage control is shown in
Fig. 5.2. The direct voltage link control consists of a PI-controller combined
with a feed forward term. Note that closed-loop bandwidth of the current
control is assumed to be much faster than the closed-loop bandwidth of the
direct voltage control.
gf
i gq ,ff Ps

u dc + + i gf

3 | e g | Pg − 1 u dc
PI
gq
÷
− + 2 − × pC

Figure 5.2: Block diagram of the closed-loop direct voltage control.

54
5.3.1 Direct Voltage Filtering
It may be required to filter the direct voltage due to pulsations in the instan-
taneous active power and instantaneous reactive power. For instance, instan-
taneous active power and instantaneous reactive power pulsations occur when
the grid connected PWM converter operates as a parallel active power filter
that filters current harmonics. Given that the grid voltage is symmetrical and
without voltage harmonics, the 5th and 7th current harmonics (for instance)
create the following pulsation in the instantaneous apparent power
3 ³∗ ∗
´ 3³
j(6ωg t+ϕ5 ) j(6ωg t+ϕ7 )
´
Sg,(6) = eg ig,(5) + ig,(7) = |eg ||ig,(5) |e + |eg ||ig,(7) |e
2 2
(5.15)
Hence, the instantaneous apparent power pulsates with 6ωg . The pulsations
that the instantaneous apparent power pulsates cause the direct voltage to
pulsate with 6ωg and this is especially significant if a small capacitor is used
on the dc-link. The grid connected PWM converter cannot counteract the
pulsations in the direct voltage and simultaneously operate as a parallel active
power filter. Luckily, the direct voltage pulsations can be separated from the
“dc” direct voltage with a low-pass filter [50]. The low-pass filter ideally
ensures that the direct voltage PI-controller does not interfere with the active
power filtering.
Note that the design of the direct voltage PI-controller may become more
advanced when a low-pass filter is placed in the feedback path [50].
Note that the filtered direct voltage should only be used in the direct voltage
PI-controller. The non-filtered direct voltage should be used in the pulse
width modulator in order to compensate for direct voltage pulsations and
therefore obtain correct voltage time area in each sampling period.
Note that direct voltage filtering also can be required for other scenarios, for
instance, when the grid voltage contains a negative sequence component [51].
The fundamental current times the negative sequence grid voltage then create
a 2ωg pulsation in the instantaneous apparent power.

5.4 Open-loop Reactive Power Control


The reactive power exchange with the grid is controlled by the reactive current
component. The simplest method to control the reactive power is via an open-
loop. Taking the imaginary part of (5.2) gives the reactive reference current
as
gf ? 2
igd = gf Q?g (5.16)
3egq

55
5.5 Design of a Current Controller for a Grid Con-
nected PWM Converter
The design of a current controller for a grid connected converter is similar
to the design of a current controller for an induction machine. It is assumed
that the grid is stiff and that an L-filter is connected between the converter
and the grid.
The utility grid (the left part of Fig. 5.1) can be written in space vector form
as
digf
g Rin gf gf
ugf gf
g − jegq
=− i − jωgf ig + (5.17)
dt Lin g Lin
It is convenient to exclude some terms from (5.17) before continuing the de-
sign of the current controller. Two terms can be excluded. Decoupled internal
model control implies that the current cross coupling is handled by a separate
loop. Hence, the term jωgf igf g can be excluded from the transfer function.
Second, the term jegfgq can also be excluded since it can be compensated by
integral action or compensated with a feed-forward term. The transfer func-
tion between voltage and current, when excluding the two mentioned terms,
becomes
1
igf
g ≈ ugf
g = Gug
gf
(5.18)
pLin + Rin
Placing the closed-loop pole in the desired bandwidth α gives the controller
as
FG α α/p
= = (5.19)
1 + FG p + α 1 + α/p
α −1 b
F = bin + αRin
G = αL (5.20)
p p
The current controller for the grid connected converter becomes an ordinary
PI-controller. The reference voltage vector is given, with current cross cou-
pling handled by a separate loop, by
µ ¶
1 ¡ gf ? ¢
uggf ? = kp 1 + ig − igf
g ωgf igf
+ jb gf
g + jegq (5.21)
Ti p
b
kp = α Lbin Ti = Rin (5.22)
bin
L
The grid flux frequency ω
bgf can with high accuracy be assumed to be 2π50 rad/s
(in Europe). The feed-forward term jegf gq is not required if the grid voltage is
constant but is very useful for compensating sudden grid voltage sags.

56
Chapter 6

Carrier Based Modulation

The choice of modulator highly affects the performance of a PWM con-


verter. Modulation is therefore an important issue and an intense research
topic. Different modulators produce, for instance, different voltage and cur-
rent distortion, noise, RFI, torque ripple and can be of various levels of sim-
plicity/complexity. Furthermore, different modulators also affect converter
switching losses and stresses and even the losses of an electric machine. This
section only discusses carrier based pulse width modulation. It is assumed
that the carrier frequency is fixed and the pulse ratio is high. Reference [6]
contains descriptions of other types of modulators and a more detailed de-
scription of carrier based modulation.

6.1 Introduction
Given reference voltages, a carrier based modulator determines the switching
instants of the semiconductor switches of a PWM converter in order to obtain
a correct voltage time area within one switching cycle. The voltage-time
area is normally controlled without feedback. For a two-level voltage source
converter, one can show that the possible converter voltage vectors in the
two-axis stator orientated reference frame become six active and two zero
vectors, as shown in Fig. 6.1. The six active voltage vectors make up a voltage
hexagon. The converter can only deliver voltage vectors that are within this
hexagon.
Carrier based modulation can be implemented either as space vector PWM
or triangular PWM. Space vector PWM provides direct control of the con-
verter switching states while triangular PWM can be considered to control
the switching states indirectly. Both methods have essentially fixed carrier
and switching frequencies, however. This chapter discusses only triangular
PWM, also called suboscillation method. The switching instants during tri-
angular PWM are given from a comparison between the three-phase voltage
references and a triangular carrier wave in the three-phase system. Fig. 6.2
shows the generation of the switching control logics when using triangular

57
PWM. The voltage references vary slowly compared to the triangular carrier
wave if the pulse ratio between the switching frequency and the fundamental
voltage references is sufficiently high. The high pulse ratio results in that the
voltage harmonics that the PWM converter produces only appear near mul-
tiples of the carrier frequency. Fig. 6.3 shows the amplitude spectrum of the
PWM converter phase voltage when using three different pulse width modu-
lation methods. These three methods are, first, when the voltage references
are pure sinusoidals as in Fig. 6.3a, and, second and third, when two different
zero sequence signals are added to the voltage references as in Fig. 6.3b and
Fig. 6.3c. Zero sequence signals will be treated more thoroughly in a following
section. As shown in Fig. 6.3, the voltage harmonics of the PWM converter
phase voltage are located around multiplies of the switching frequency. Note
in Fig. 6.3c, though, that a more distributed amplitude spectrum is obtained
when using the zero sequence DPWM1. This is because the DPWM1 is a
so-called discontinuous zero sequence signal and the instantaneous switching
frequency varies between the mean switching frequency and the triangular
carrier wave frequency. Discontinuous zero sequence signals will be treated
more thoroughly in a following section.
u(010) jβ u(110)

m=1 m=0.91
u(011) u(100)
α
u(000) u(111) 2
u dc
3

m=0.79

u(001) u(101)

Figure 6.1: Voltage hexagon of a PWM voltage source converter.

58
a) 1
u* , carrier

*
us1
*
0 us2
u*
123

s3
−1
17 17.05 17.1 17.15 17.2 17.25 17.3 17.35 17.4
Tsw/2
b) 1.5

1
sw1

0.5 0 1 1 1 1 1 0
0
17 17.05 17.1 17.15 17.2 17.25 17.3 17.35 17.4

c) 1.5

1
sw2

0.5
0 0 1 1 1 0 0
0
17 17.05 17.1 17.15 17.2 17.25 17.3 17.35 17.4

d) 1.5

1
sw3

0.5 0
0 0 0 1 0 0
0
17 17.05 17.1 17.15 17.2 17.25 17.3 17.35 17.4
Time (ms)

Figure 6.2: Example of the switching control logic generation for three-phase triangular
PWM.

59
a)
1
Amplitude

0.5

0
f1 1 2 3 4 5 6 7 8 9 10

b)
1
Amplitude

0.5

0
f1 1 2 3 4 5 6 7 8 9 10

c)
1
Amplitude

0.5

0
f1 1 2 3 4 5 6 7 8 9 10
Frequency (kHz)

Figure 6.3: Simulated voltage spectrum of the PWM converter phase voltage at m = 0.74 and
fc = 3 kHz. The spectrum is normalized with the amplitude of the fundamental component.
(a) Spectrum for pure sinusoidal voltage references. (b) Spectrum when the zero sequence
signal SPWM-sym is added to the original voltage references. (c) Spectrum when the zero
sequence signal DPWM1 is added to the original voltage references.

60
a) 1

usα−esα (kV) 0.5


0
−0.5
−1
17 17.05 17.1 17.15 17.2 17.25 17.3 17.35 17.4
t=0 t=T
b) 120

100
(A)

80
i

60
17 17.05 17.1 17.15 17.2 17.25 17.3 17.35 17.4
t=0 Time (ms) t=Tsw

Figure 6.4: Regular asymmetric sampling. (a) Inductor voltage and carrier wave. (b) Current
and sampling instants.

6.2 Regular Sampling


The current begins and ends at its mean value within one half switching
period for carrier based modulation. The time instants when the current
is at its mean values coincide with the time instants when the triangular
carrier wave reaches its maximum and minimum value. Regular sampling of
the three-phase currents at these time instants allows a rather low sampling
frequency in order to derive the current mean values. Furthermore, regular
sampling also reduces the need for filtering the sampled currents. Fig. 6.4
shows the triangular carrier wave, one phase inductor voltage and one phase
current with marked sampling events. Regular sampling can be carried out
as symmetrical regular sampling and asymmetrical regular sampling, which
correspond to sampling frequencies of fc and 2fc , respectively, where fc is the
carrier wave frequency. Asymmetrical regular sampling enables the possibility
of doubling the closed-loop bandwidth of the current control and slightly less
current distortion compared to symmetrical regular sampling [6].

6.3 Modulation Index

The modulation index is the normalized fundamental voltage of a PWM con-


verter, defined as [6]
û(1) û(1)
m= = , 0≤m≤1 (6.1)
û(1),sixstep 2udc /π
where û(1) is the crest fundamental component of the converter output volt-
age. It is common to define three circles in the voltage hexagon, as shown

61
in Fig. 6.1. These three circles correspond to three important modulation
indexes. The inner circle corresponds to the maximum linear modulation in-
dex for pure sinusoidal reference voltages and corresponds to m = 0.79. The
linear modulation index can be increased up to the maximum mlin max = 0.91
by adding triplen harmonics to the reference voltages. This is where linear
modulation without low-order voltage harmonics ends for a PWM converter.
The maximum linear modulation index corresponds to a circle that just fits
within the voltage hexagon. Above mlin max = 0.91 the fundamental of the volt-
age output still increases but the modulation is no longer linear and low order
harmonics appear in the phase voltages and phase currents. The low order
harmonics include all odd harmonics with the exception of triplen harmonics,
hence 5th, 7th, 11th... Note that the low order harmonics are not necessarily
a problem. The influence from torque ripple that the harmonics produce is
often low at high loads and especially for high inertia loads. The maximum
fundamental voltage that a voltage source converter can deliver occurs when
the integration of the applied voltage vectors, i.e., the flux, slides along the
six sides of the voltage hexagon in Fig. 6.1. In the three phase system the
voltage references now have the square wave shapes. PWM longer occurs in
this operation mode and this operation mode is instead referred to as sixstep
operation. By definition, sixstep operation corresponds to a modulation index
equal to 1.
The region in the voltage hexagon that is outside the circle corresponding to
mlin
max = 0.91 but within the voltage hexagon is referred to as the overmodula-
tion region. The overmodulation region is useful for maximizing the voltage
capability of the PWM converter during continuous operation, for instance
in traction applications, and is also useful for improving the current step re-
sponse. Voltage vectors in the region outside the voltage hexagon cannot
be obtained however and correspond to voltage saturation. Actions should
be taken in order to avoid integrator windup of the current controllers when
voltage saturation occurs. Furthermore, it is possible to slightly affect the
current dynamics by freely choosing a proper voltage vector on the hexagon
boundary during voltage saturation. Above all, the choice of voltage vector
on the hexagon boundary affects the cross coupling between the active and
reactive current components during voltage saturation [52].

6.4 Triangular PWM with Zero Sequence Injection


Triangular PWM is in its simplest form carried out by comparing sinusoidal
voltage references with a triangular carrier wave, so called sinusoidal modula-
tion. However, the maximum linear modulation is 0.79 for sinusoidal modula-
tion and the current distortion is not minimized when the voltage references
are sinusoidals [7]. The modulation index can be increased and the current

62
éá"ì_ï å"á'æ,ê‚ã î ë'ææ,ãRì'ì'íWá"ç,ì î ä á"ä â_ì#æ,ð çá"è äã î ä æ,í
ñ ò
ï î á'ã äê‚ç,ì'ë'õ'ç,ù"ã ä è í$çá"ãRì únúnç,å'ä î æ ÷
i s1 is 2 i s3

÷ öç,ã â_ê‚â'á"ä îì'ä á'ä é,ì í)á"íWðæ,ø'î ã ù'ä ë_æì'è æé,ì æ å'ç,è å"æ@éá"ê‰ë"æì'íWç,î ä á"ì
u dc ò
Ù + u s1'
+ + +
u
u
Ù
s1
+ +
Ù
u s 2'

+ + + − àá â'ã ä å'æ,ã
6
s2
+ + u s3' + + + −

u s3

ã ä ë'ó ë'è ï ænñ é,è ä á"ì'ê‚ô ë_å"æ,á"ì'è îíWççõ"î ä æ á"ì


+ −

ÚÜÛÝ ÚbÞ ß 2
u dc
1
+ + -1
-1/2
uz

Figure 6.5: Block diagram of triangular PWM using SPWM-sym zero sequence and with
compensation for converter non-idealities.

distortion may be reduced by adding a zero sequence to the sinusoidal waves.


The zero sequence contains only triplen harmonics. The added triplen har-
monics do not appear in the phase voltages nor in the phase currents in a three
phase system with an unconnected neutral point. Fig. 6.5 shows a block dia-
gram of triangular PWM using the zero sequence SPWM-sym [7]. The zero
sequence signal is added to the original sinusoidal reference voltages and the
voltage references are compared with the triangular carrier wave in three sep-
arate comparators. Although not really part of the modulation, Fig. 6.5 also
shows compensation for converter non-idealities. The objective of the com-
pensation is to obtain the correct voltage time area in each sampling period.
The compensation consists of compensation for the blanking time and on-
state voltage drop of the semiconductor switches and compensation for direct
voltage variations [53].

6.5 Some Zero Sequence Signals


It is common to distinguish between triangular PWM methods that use “con-
ventional” and those that use discontinuous zero sequence signals. Conven-
tional PWM modulates all three phase-legs with the same switching frequency
while discontinuous PWM leaves one of the phase-legs unmodulated during
each switching period, see Fig 6.6. The switching frequency of conventional
PWM is therefore equal to the carrier frequency while the mean switching
frequency for all three phase-legs of discontinuous PWM becomes 2/3 of the
carrier frequency. The dominating voltage harmonics of discontinuous PWM
still appear at multiples of the carrier frequency however, as shown in Fig. 6.3.
Fig. 6.7 shows the voltage references and zero sequence signal of two com-
mon conventional PWM methods. These two methods are SPWM-sym and

63
t 000 t100 t110 t111 t111 t110 t100 t 000 t100 t110 t111 t111 t110 t100
2 2 2 2 2 2 2 2 2 2 2 2 2 2

sw1 0 1 1 1 1 1 1 0 sw1 1 1 1 1 1 1

sw 2 0 0 1 1 1 1 0 0 sw 2 0 1 1 1 1 0

sw 3 0 0 0 1 1 0 0 0 sw 3 0 0 1 1 0 0

T T T T
2 2 2 2
(a) (b)

Figure 6.6: Switching pattern of the semiconductor switches for: (a) Conventional three-
phase PWM. (b) Discontinuous three-phase PWM.

0.5 0.5

0.25 0.25
Voltage (pu)

Voltage (pu)

0 0

−0.25 −0.25

−0.5 −0.5
0 π/3 2π/3 π 4π/3 5π/3 2π 0 π/3 2π/3 π 4π/3 5π/3 2π
ωt (rad) ωt (rad)

Figure 6.7: Voltage references and corresponding continuous zero sequence signal at m = 0.7
for two continuous PWM methods. The y-axes are normalized with udc and the voltage
reference of phase 1 and the zero sequence signal are plotted with thick lines. The continuous
PWM methods are: (left) SPWM-sym. (right) SPWM-opt.

SPWM-opt. SPWM-sym adds an offset to the reference voltages which places


the maximum and minimum voltage references symmetrically around the time
axis. The symmetrical distribution of the voltage references results in a sym-
metrical time distribution of the zero vectors. SPWM-opt has a not quite
symmetrical time distribution of the zero vectors which slightly reduces the
current distortion compared to SPWM-sym [7].
Discontinuous PWM leaves one of the phase legs unmodulated each switching
period, i.e., one zero vector in the voltage hexagon being used during each
switching period. Alternatively, discontinuous PWM implies that one of the
voltage references is clamped to either the positive or the negative direct
voltage link busbar voltage in the three-phase domain. Fig. 6.8 shows the
voltage references and zero sequence of six discontinuous PWM methods.
The voltage references of DPWM0, DPWM1 and DPWM2 are clamped to

64
negative or positive dc-busbar voltage during 60◦ , DWPW3 is clamped during
30◦ while DPWMMAX and DPWMMIN are clamped during 120◦ .
Table 6.1 describes how to generate the zero sequence signals in Fig. 6.7
and Fig. 6.8 [7], [54]. All PWM methods in Table 6.1 have maximum linear
modulation index 0.91, except for SPWM-opt that has maximum linear mod-
ulation index 0.88 [7]. The table rows for DPWM0, DPWM1, and DPWM2
should be read as follows. The original sinusoidal reference voltages are first
phase shifted. The maximum absolute value of the phase shifted reference

Table 6.1: Generation of some common zero sequences for triangular PWM.
Sequence name Zero sequence generation

SPWM-sym uz = −[min{u?1 , u?2 , u?3 } + max{u?1 , u?2 , u?3 }]/2

umax = max{u?1 , u?2 , u?3 }, umin = min{u?1 , u?2 , u?3 }


umax + umin
SPWM-opt usym = − , usum = umax + umin
2
usum (umax + usum )(umin + usum )
uopt =
2 umax (umax + usum ) + umin (umin + usum )
uz = usym + uopt

γ = 0, us?0 = e−j(γ−π/6) us?


? udc
DPWM0 assume |u?0 ?0 ?0
1 | ≥ |u2 |, |u3 |, then uz = sign{u1 } − u?1
2

γ = π/6, us?0 = e−j(γ−π/6) us?


? udc
DPWM1 assume |u?0 ?0 ?0
1 | ≥ |u2 |, |u3 |, then uz = sign{u1 } − u?1
2

γ = 2π/6, us?0 = e−j(γ−π/6) us?


? udc
DPWM2 assume |u?0 ?0 ?0
1 | ≥ |u2 |, |u3 |, then uz = sign{u1 } − u?1
2

udc
DPWM3 assume |u?2 | ≤ |u?1 | ≤ |u?3 |, then uz = sign{u?1 } − u?1
2

udc
DPWMMAX uz = − max{u?1 , u?2 , u?3 }
2

udc
DPWMMIN uz = − − min{u?1 , u?2 , u?3 }
2

65
a) b)
0.5 0.5

0.25 0.25

Voltage (pu)
Voltage (pu)

0 0

−0.25 −0.25

−0.5 −0.5
0 π/3 2π/3 π 4π/3 5π/3 2π 0 π/3 2π/3 π 4π/3 5π/3 2π
ωt (rad) ωt (rad)
c) d)
0.5 0.5

0.25 0.25
Voltage (pu)
Voltage (pu)

0 0

−0.25 −0.25

−0.5 −0.5
0 π/3 2π/3 π 4π/3 5π/3 2π 0 π/3 2π/3 π 4π/3 5π/3 2π
ωt (rad) ωt (rad)
e) f)
0.5 0.5

0.25 0.25
Voltage (pu)
Voltage (pu)

0 0

−0.25 −0.25

−0.5 −0.5
0 π/3 2π/3 π 4π/3 5π/3 2π 0 π/3 2π/3 π 4π/3 5π/3 2π
ωt (rad) ωt (rad)

Figure 6.8: Voltage references and corresponding zero sequence signals at m = 0.7 for six
discontinuous PWM methods. The y-axes are normalized with udc and the voltage reference
of phase 1 and the zero sequence signal are plotted with thick lines. The discontinuous PWM
methods are:
(a) DPWM0. (b) DPWM1. (c) DPWM2. (d) DPWM3. (e) DPWMMAX. (f) DPWMMIN.

66
voltages is then identified. The non phase shifted reference voltage which
corresponds to the phase shifted reference voltage with maximum absolute
value is then used to generate the zero sequence signal.
The loss factor d2 is one measure of PWM performance. More specific, the
loss factor is a measure of the phase current distortion and the harmonic
copper losses of an electric machine are proportional to the loss factor [6].
The loss factor is given by [6]
s
Z
1 Ts ³ ´2
Ih,rms = i(t) − i(1) (t) dt (6.2)
Ts 0
2
2
Ih,rms
d = 2 (6.3)
Ih,rms,sixstep
where i(t) is a distorted phase current, i(1) (t) is the fundamental component
of the distorted phase current and Ts is the fundamental period time. The loss
factor is calculated for a pure inductive system and the back-emf is increased
accordingly with the modulation index. Normalization with the harmonic
distortion of sixstep operation makes the loss factor independent of the load
inductance. A computer program can be constructed which calculates the
loss factor in the stator orientated reference frame.
Fig. 6.9 shows the loss factor of the PWM methods in Table 6.1 as a func-
tion of modulation index. Fig. 6.9a compares conventional and discontinuous
PWM methods for the same carrier frequency of 3 kHz. The conventional
PWM methods have less loss factor throughout the entire linear modulation
range. The difference of SPWM-sym and SPWM-opt is very small. Fig. 6.9b
shows a comparison between conventional and discontinuous PWM methods
for the same effective switching frequency. The loss factors of discontinuous
PWM methods are less than the loss factors of conventional PWM methods
for approximately m > 0.7. This makes discontinuous PWM methods an
interesting alternative to conventional PWM for PWM converters that of-
ten operate at high modulation indexes, for instance a grid connected PWM
converter.
The torque factor q can also be used as a measure of PWM performance. The
torque factor q is a measure of the active current distortion, i.e., the torque
pulsations. The torque factor is here defined as
s
Z
1 Ts ³ ´2
Ith,rms = Iq − Iq,(dc) dt (6.4)
Ts 0

Ith,rms
q= (6.5)
Ith,rms,sixstep
67
a) 0.06 b) 0.025
6 1
6
0.05 2
0.02

0.04 7
7 3
Loss factor

Loss factor
3 0.015 4
4
0.03 5
5
0.01
1
0.02
2
0.005
0.01

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Modulation index Modulation index

Figure 6.9: Loss factor as a function of modulation index. (a) Continous and discontinuous
PWM methods both use a carrier frequency of 3 kHz. (b) Continous PWM methods use
a carrier frequency of 3 kHz while discountinous PWM methods use a carrier frequency of
1.5 · 3 = 4.5 kHz. 1: SPWM-sym. 2: SPWM-opt. 3: DPWM3. 4: DPWM1. 5: DPWM0
and DPWM2. 6: DPWMMAX. 7: DPWMMIN.

where Iq is the distorted active current component, Iq,(dc) is the mean of


the active current component. The definition of the torque factor is slightly
different than in [7], although the underlying idea is the same.
Fig. 6.10 shows the torque factor of the PWM methods in Table 6.1 as a func-
tion of the modulation index. The torque distortion of conventional PWM
methods is lower than the torque distortion of discontinuous PWM methods
in nearly the entire linear modulation range for both equal carrier frequency
and equal effective switching frequency. The low torque distortion of conven-
tional PWM methods is due to the symmetrical time distribution of the two
converter zero vectors [7].
The switching losses of discontinuous PWM methods have a dependency on
the phase angle between voltage and current. This is because the switching
losses decrease if the phase currents are at maximum when the voltage refer-
ences are clamped to the positive or negative dc-bus bar voltage. Fig 6.11a
shows what happens in one PWM converter phase-leg when the voltage ref-
erences, with an injected zero sequence DPWM1, are in phase with the phase
currents. As shown in the phase potential and the switching pattern of one
phase leg in Fig 6.11b-f, no switchings occur when the phase current is at
maximum. Fig. 6.12 shows the switching losses of the conventional and dis-
continuous PWM methods for both equal carrier frequency and equal effec-
tive switching frequency. The switching losses have been derived with the loss
model in [55]. It is assumed that diodes and transistors have equal switching
losses and that the switch-on losses are equal to the switch-off losses. The

68
a) 0.8 b)
6 0.5 6
0.7 5
4 5
0.6 3 3 7
4 0.4
Torque ripple

Torque ripple
0.5 1
7 0.3
0.4
2
0.3 0.2
0.2
2 0.1
0.1 1
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Modulation index Modulation index

Figure 6.10: Torque ripple as a function of modulation index. (a) Continous and discontinu-
ous PWM methods both use a carrier frequency of 3 kHz. (b) Continous PWM methods use
a carrier frequency of 3 kHz while discountinous PWM methods use a carrier frequency of
1.5 · 3 = 4.5 kHz. 1: 1: SPWM-sym. 2: SPWM-opt. 3: DPWM3. 4: DPWM1. 5: DPWM0
and DPWM2. 6: DPWMMAX. 7: DPWMMIN.

switching losses that are shown in Fig. 6.12 therefore have a limited accuracy
and should merely be seen as a qualitative description of the switching losses
of different PWM methods. The switching losses are plotted relative to the
switching losses of conventional PWM methods, which have no dependency
on the phase angle. The switching losses of discontinuous PWM for equal
carrier frequency in Fig. 6.12a are 17 % to 50 % lower than the switching
losses of continuous PWM. The mean switching loss reduction is approxi-
mately 33 % which is due to that the effective switching frequency is 33 %
lower for the discontinuous PWM methods. Fig. 6.12b shows that conven-
tional and discontinuous PWM methods have the same mean switching losses
for equal effective switching frequency. The switching losses of discontinuous
PWM methods vary between 25 % lower and 25 % higher than conventional
PWM methods.
Although not shown here, DPWMMAX and DPWMMIN place unequal
losses and unequal thermal stresses on the upper and lower semiconductors
within one phase leg [54]. This makes DPWMMAX and DPWMMIN less
desirable zero sequences, although one may consider alternating between DP-
WMMAX and DPWMMIN in order to obtain evenly distributed losses.
The graphs in Fig. 6.13 show a comparison of the current distortion of
SPWM-sym and DPWM0 during one switching period. Fig. 6.13a shows
that the q-axis current ripple of DPWM0 is approximately twice as high the
q-axis current distortion of SPWM-sym for equal carrier frequency. For equal
mean switching frequency the current triangles of DPWM0 become smaller
which reduces current distortion, as shown in Fig. 6.13b.

69
a)
1
u1,i1,carrier

−1
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02
Time (s)
b)
1
v1

−1
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02
Time (s)
c) 1.5

1
T1+

0.5

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02
Time (s)
d) 1.5

1
D1−

0.5

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02
Time (s)
e) 1.5

1
T1−

0.5

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02
Time (s)
f) 1.5

1
D1+

0.5

0
0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02
Time (s)

Figure 6.11: Switching pattern when using DPWM1. The phase current is at maximum when
the voltage reference is clamped and switching losses are reduced. (a) Voltage reference (solid)
and current (dashed) of one phase, triangular carrier wave (dotted). (b) Phase potential. (c)
Switching pattern of the upper transistor in one phase leg. (d) Switching pattern of the lower
diode in one phase leg. (e) Switching pattern of the lower transistor in one phase leg. (f)
Switching pattern of the upper diode in one phase leg.
70
a) 1.4 b) 1.4
X: DPWMMAX, DPWM0 DPWM1 DPWM2
1.3 1.3
DPWMMIN
1.2 1.2
Relative switching losses

Relative switching losses


1.1 1.1 X
SPWM−sym, SPWM−opt
1 1
0.9 DPWM0 DPWM1 DPWM2 0.9
DPWM3
0.8 0.8
X
0.7 0.7
X: DPWMMAX,
0.6 0.6
DPWMMIN
DPWM3
0.5 0.5
0 30 60 90 120 150 180 0 30 60 90 120 150 180
Phase angle (deg) Phase angle (deg)

Figure 6.12: Switching losses for conventional and discontinuous PWM. (a) Continous and
discontinuous PWM methods both use a carrier frequency of 3 kHz. (b) Continous PWM
methods use a carrier frequency of 3 kHz while discountinous PWM methods use a carrier
frequency of 1.5 · 3 = 4.5 kHz.

a) b)
β, q β, q

α, d α, d
sq

sq
issβ, isf

issβ, isf

0 0

SPWM−sym
SPWM−sym
DPWM0
DPWM0

0 0
issα, isf issα, isf
sd sd

Figure 6.13: Stator current ripple for SPWM-sym and DPWM0. (a) Both PWM-methods
use a carrier frequency of 3 kHz. (b) SPWM-sym uses a carrier frequency of 3 kHz while
DPWM0 uses a carrier frequency of 4.5 kHz.

71
6.6 Avoiding Short Time Duration Switching States
A semiconductor switch requires a certain time to fully reach its on- and off-
states. Short time duration switching states are therefore undesirable since
they do not affect the output voltage but only create switching losses. The
minimum on-duration switching state is around 1-100 µs [6], where the lower
figure rather than the upper figure is relevant when using IGBT:s. Short time
duration switching states are especially likely for high modulation indexes.
Furthermore, short time duration switching states can also occur when using
triangular PWM with discontinuous zero sequence injection. This is because
the clamped voltage references and the triangular carrier wave have the same
magnitude and accidental short time duration switchings are likely to occur.
In practice, avoiding short time duration switching states results in a slight
reduction in the maximum modulation index [6].

6.7 Summary of Triangular PWM with Zero Sequence


Injection
Discontinuous PWM methods reduce the current distortion at high modula-
tion indexes for the same mean switching frequency and can greatly reduce
the switching losses. The total losses of a PWM converter can be reduced by
5-10 % [55].
Discontinuous PWM methods seem to be suitable for a grid connected PWM
converter since it normally operates at high modulation indexes only. Dis-
continuous PWM methods have lower loss factors compared to conventional
PWM methods for high modulation indexes, assuming an equal mean switch-
ing frequency. Furthermore, especially the discontinuous PWM method DPWM1
seems suitable to use since the switching losses of DPWM1 are at minimum at
0◦ and 180◦ phase angle and a grid connected PWM converter often operate
at unity power factor. The loss factor and the switching losses reduction of
DPWM1 compared to conventional PWM methods depend on how the ef-
fective switching frequency is chosen. For the machine side PWM converter,
one may switch between different PWM methods in order to obtain minimum
switching losses as the phase angle changes [56]. Note that in order to main-
tain a low loss factor and a low torque ripple, also the carrier frequency must
be changed when switching between continuous and discontinuous PWM.
An interesting choice is to let the mean switching frequency of DPWM1 be
twice as high as the permissable switching frequency of conventional PWM
methods. The increased mean switching frequency of DPWM1 allows for a
smaller sizing of the grid filter [57]. Meanwhile, DPWM1 gives the same
switching losses as conventional PWM methods at unity power factor.

72
The modulation of the grid side PWM converter and the machine side PWM
converter can also be co-ordinated in order to cancel out two common voltage
pulses [58] for certain grounding schemes. The essence of this modulation
technique is that the grid side PWM converter generates the zero sequence of
the machine side PWM converter.

73
74
Chapter 7

Simulations

7.1 Software Setup


A block diagram of the proposed control system for the double-sided PWM
converter is shown in Fig. 7.1. Both the induction machine and the grid con-
nected PWM converter are vector controlled and use triangular PWM with
injected zero sequence SPWM-sym. For simplicity, the control system has
been drawn with space vectors although the real and imaginary parts are
treated separately in the actual implementation. The current and voltage
model is here used as stator flux observer and the current model is imple-
mented in the rotor orientated reference frame. Euler forward discretization
is used to transform PI-controllers and the well-damped flux observer into
discrete form.
The PI-controllers in the machine control and in the control of the grid con-
nected PWM converter use back-calculation [19] in order to avoid integrator
windup. The principle of back-calculation for the grid connected converter is
shown in Fig. 7.2. The original control errors, say e?u , may be too large, which
results in that isf ?
sq must be limited in order to protect the semiconductor
switches. If this happens, integrating e?u causes unnecessary current over-
shoot. The overshoot can be avoided by back-calculating the “new” control
error eu that is given by the limited value of isf ?0
sq .

7.1.1 Changes in the Control System


The control system is nearly identical to the proposed control system in the
previous chapter. Two minor changes have been made, however. First, volt-
age decoupling of the induced emf has been included in the reference stator
voltage vector. The rotor speed has very fast dynamics in the simulations,
and the decoupling term is required to obtain the correct torque. Second,
the integration time for the direct voltage controller has been decreased to
5/α. This gives a faster tracking of the direct voltage reference. Unfortu-
nately, the decreased integration time has also been found to give a larger

75
direct voltage overshoot/undershoot when the power flow is reversed. The
direct voltage overshoot/undershoot during power flow reversal is considered
acceptable, however.
j3ωgf T / 2
e
jθˆ gf jθˆ 'gf
e e
û ülýRüRþ,ÿ
eg1
Cartesian

e g2 j
− jθˆ gf
j | eg |
e g3
| eg | e
jωg L̂ g
i g1
123 i sg i gf
g 
u g1
i g2 − jθˆ gf + + u gf
g

u sg αβ


i g3 αβ e
PI jθˆ 'gf 
u g2
+ + e 123
÷ i gf
gd
 
u g3
2
×

Qg 3
ji gf
û ü ü ü Rû $üRWû þ[ü

gq
u dc        
+
PI j   


u dc
i qgf,ff ÷ | eg |

×
i sf
sq u sf
sq


 


ψ̂ sf θ
sd
Current
e jθsf e jθsf
ˆ ˆ'
voltage
ψ̂ sf jωψˆ sf
e − jθsf
ˆ model sd sd

is1 jω e j3ωT / 2
123 i ss i sf
s 
u s1
e − jθsf
is2 ˆ
αβ + u sf
s

u ss 
αβ 

e jθsf
ˆ' us2
is3 PI
+ 123
L̂ σ (i sf
sq )
2 
u s3
ψˆ sf
sd − L̂ σi sd
sf

+ i sf
sd


PI

ü û,[ÿWû û ü WüRþ[Wþ ü ü Rû û,ÿWü û ÿ û ü


ψ sf  +
sd ! " #  
  $
  
 
÷ 2 ji sf
sq

    "      %  


jTe 
× 3p

Figure 7.1: Block diagram of the vector control system for the double-sided PWM converter.

i gf
g (k)
je gf
gq ( k )
i ffgq (k ) i gf
gd, ( k )
jωgf L̂ g i ggf (k ) +
jωgf L̂ g
+
& ')(*' +


u dc(k) e u (k) 
+ i gf
 ' gf
gq ( k ) i gd, ( k )
− ei (k) 
+ u gf
g , (k )
kp kp
+ − + + + + + +
limit +
u dc(k)
− + − u gf '

T e u (k) T ei (k) + g (k )
1/ k p 1/ k p
Ti (z − 1) − Ti (z − 1)
T T −
 e u (k )  ei (k )
Ti Ti

Figure 7.2: PI-controllers with anti-windup for the grid connected converter.

76
7.1.2 Simulation Environment
The simulations have been carried out with Matlab/Simulink 5.3. The simu-
lation computer uses a 450 MHz processor and 256 Mb RAM. Fifty seconds in
real-life correspond to 0.1 s in the simulation program. The simulation speed
is reduced during long time simulations because the virtual RAM of the hard
disk is accessed. The control system is implemented in Simulink S-functions.
Observer states and controller integrator sums have been implemented as
persistive variables in the Simulink S-functions. This implementation of the
control system gives a simulation environment that is rather similar to C-
programming of a physical floating point digital signal processor.

7.2 Safety Limits


A number of safety limits have been implemented in the control system. The
purpose of the safety limits is to obtain a better simulation agreement with
a real physical system. The safety limits are, in some cases, also required for
proper operation of the back-calculation anti-windup and for proper operation
of the stator flux observer.
The safety limits described in this section are speed limit, voltage limit and
current limit. The current limit is divided into three parts; namely torque
limit, current amplitude limit and power limit. The section describes why the
safety limits are required and how they are set. Mainly transient limits are
described. However, continuous limits are believed to be similar, with the dif-
ference that lower limits must be chosen. In addition to these software safety
limits, one may, for instance, also consider a direct voltage software limit. It
may not be required to trip the machine side converter if the direct voltage
deviation becomes too large. Instead, the machine side PWM converter can
assist the grid side PWM converter with direct voltage regulation if the direct
voltage deviation is above a certain value. A software direct voltage limit is
not included in these simulations, however. This is because the direct voltage
remains fairly constant due to the feed-forward term.

7.2.1 Speed Limit


A speed limit can be required for at least two reasons. First, the mechanical
design of the induction machine may not permit rotor speeds above a certain
value. Second, the torque capability of the system is reduced when operat-
ing above base speed in the flux weakening region. Depending on the load
characteristics, an unstable operating point may be reached. The maximum
speed has been set at ±1.4 pu in the simulations.

77
7.2.2 Stator Current Limits
Three stator current limits have been identified. These three current limits
are torque limit, amplitude limit and power limit.

Torque Limit
The torque limit is required in order to limit the torque below the pull out
torque, see Section 3.3.2. There may also be a torque limit/torque response
limit due to mechanical constraints. The torque limit has been performed by
limiting the torque reference within
à !2
b
ψsdsf
[Te? ]max max
min = ±Te (7.1)
ψ base

where ψ base is the base flux linkage. The torque limit hence accounts for
variations in the stator flux. The maximum torque, when actual stator flux is
equal to base flux, has been set at 150 % of base torque, which is well below
the pull-out torque for this machine. The active current limit becomes
2|Te? |
[isf max
sq ]Te = (7.2)
3pψbsf
sd

Current Amplitude Limit


The current amplitude limit has been divided into a continuous limit and a
transient limit. The purpose of the continuous amplitude limit is to avoid
overheating the the induction machine. The continuous amplitude limit de-
pends on how much cooling the machine receives, which in turn depends on
the speed.
During shorter transients, the current amplitude can be allowed to exceed
the continuous rating. A transient current amplitude limit can therefore be
introduced. The transient current limit is likely to be given by the ratings
of the semiconductor switches, because a semiconductor switch has a shorter
thermal time constant than an induction machine. For instance, some IGBT
manufacturers claim 150 % overload capability for 60 seconds [59], while an
induction machine may be able to handle 150 % overload for 10-20 minutes.
Both the reactive and the active current componenents should be limited.
The transient reactive current limit should be high for stator flux orientation.
This is due to the cross coupling between the active stator current and stator
flux. The active stator current greatly reduces the stator flux during torque
transients if the reactive current becomes limited. The continuous and tran-
sient reactive current limit are therefore set at 1 pu. The transient active

78
current limit depends on the present reactive current according to
q
sf max max )2 − (isf )2
[isq ]I = (Is,trans (7.3)
sd

max
The maximum transient stator current amplitude Is,trans is set at 1.5 pu. For
a grid connected PWM converter that operates at unity power factor, the
transient active current limit becomes

[igf max
gq ]I
max
= Ig,trans (7.4)
max
The maximum transient grid current amplitude Ig,trans is set to 1.5 pu.

Power Limit
A power limit has been implemented for the stator current. The purpose
of the power limit is to ensure that a power balance can be maintained on
the direct voltage link. There can be both transient and continuous power
limits. Only the transient power limit is treated here but the continuous limit
is believed to be similar, only a lower power limit must be chosen.
The maximum transient power that the grid connected PWM converter can
transfer is, according to the previous described grid current amplitude limit
3
Pgmax = |eg |Ig,trans
max
(7.5)
2
The power limit on the grid side sets the following limit on the transient active
stator current
sf max
Pgmax − usf sf
sd isd
[isq ]P = (7.6)
|usf
sq |
The power limit Pgmax on the grid side is set at 1.45 pu, i.e., 0.05 pu below
the maximum theoretical value.

Summary of Current Limits


The control system calculates [isf max sf max
sq ]Te , [isq ]I , [isf max
sq ]P each sample. The
three current limits are all positive and the resulting stator active current
reference becomes
n o n o
sf ? sf max sf max sf max
isq = min [isq ]Te , [isq ]I , [isq ]P · sign Te? (7.7)

The current of the grid connected PWM converter is only exposed to the
current amplitude limit.

79
7.2.3 Voltage Limit
The voltage limit is considered to be required for two reasons. First, a contin-
uous voltage limit is required for continuous operation in the flux weakening
region. The continuous voltage limit is set at 1 pu. The stator flux reference
is set at 1.2 pu, hence, the base speed for flux weakening becomes 0.83 pu.
The flux reference above the base speed for flux weakening is set at 1/ω. Sec-
ond, a transient software voltage limit is also useful. The transient software
voltage limit is especially useful when PWM voltages on the machine side
and grid side are not measured. The software transient voltage limit ideally
ensures that the stator flux observer and the back-calculation of the current
controller receives the correct voltage even during voltage saturation. The
software transient voltage is given by the voltage hexagon of the grid side and
machine side voltage source converters, and depends on the present direct
voltage and the reference voltage vector position, see [52] for further details.

80
7.3 Simulation Results
This section presents the simulation results of a double-sided PWM converter
that uses the control system that was described in the previous chapters.
The parameters of the 300 kW induction machine given in Appendix E are
used and the remaining simulation parameters are shown in Table 7.1. It
is assumed that the control system has perfect knowledge of the induction
machine and grid filter parameters and no measurement noise or dc-offsets
exist. The load is a constant torque load of 0.5 p.u. An ideal semiconductor
switch model is used, hence, a switch is modeled as a short-circuit in its on-
state and as an open circuit in its off-state. SPWM-sym is used as a zero
sequence signal.

Table 7.1: Simulation parameters.


Machine side converter and control parameters:
Switching frequency 4.8 kHz ¿ 96 p.u.
Sampling frequency 4.8 kHz ¿ 96 p.u.
Closed-loop current bandwidth 150.0 Hz ¿ 3 p.u.
Closed-loop stator flux bandwidth 5.0 Hz ¿ 0.1 p.u.
Current and voltage model, σ1 = σ2 3.0 Hz ¿ 0.06 p.u.
Stator flux reference below base speed 2.11 Vs ¿ 1.18 p.u.
Base speed for flux weakening 1245.0 rpm ¿ 0.83 p.u.
Grid side converter and control parameters:
Switching frequency 4.8 kHz ¿ 96 p.u.
Sampling frequency 4.8 kHz ¿ 96 p.u.
Closed-loop current bandwidth 250.0 Hz ¿ 5 p.u.
Closed-loop direct voltage bandwidth 8.5 Hz ¿ 0.17 p.u.
direct voltage link:
direct voltage link voltage 1200.0 V ¿ 1.74 p.u.
direct voltage link capacitor 39.6 mF ¿ 0.058 p.u.
Grid filter:
Inductance 0.44 mH ¿ 0.1 p.u.
Resistance 14.0 mΩ ¿ 0.01 p.u.

81
Figs. 7.3 and 7.4 show induction machine variables and grid side variables,
respectively, during a speed reversal from 0.7 p.u. to -1.4 p.u. The speed
reversal is obtained with torque control. The speed reversal takes only 0.35 s.
The fast speed reversal is due to the four-quadrant capability of the double-
sided PWM converter, and partly due to the load torque. The torque response
time (10 % to 90 %) is slightly less than 3 ms. The stator current amplitude
in Fig. 7.3f is maintained below 1.5 p.u. The reactive stator current varies
in order to counteract the cross coupling between active stator current and
stator flux, and in order to weaken the flux above the rotor speed 0.83 p.u. The
stator current is limited due to the power limit at t=0.29 s. The result of the
power limit can be seen in Fig. 7.4b and Fig. 7.4f, i.e. the power limit ensures
that the grid power and the grid current amplitude is maximum 1.45 p.u. The
stator power limit ensures a power balance on the direct voltage link and the
grid side PWM converter can, hence, maintain constant direct voltage. The
stator power limit also reduces the torque, as shown in Fig. 7.3b. The stator
power limit together with the feed-forward term of the induction machine
power gives a nearly constant direct voltage during the speed reversal, as
shown in Fig. 7.4a. The reactive power in Fig. 7.4c is zero during the speed
reversal, hence a unity power factor is maintained.
Fig. 7.5 shows the torque-speed trajectory during the speed reversal. The
induction machine and also the double-sided PWM converter operate in all
four quadrants during the speed reversal. The four-quadrant operation gives
fast speed dynamics. Three safety limits are encountered during the speed
reversal. These safety limits are maximum transient torque, maximum tran-
sient power and maximum speed. Note that it may be difficult to obtain this
“optimal” torque-speed trajectory with a simple speed controller.

82
a) 1
0.5
Speed (pu)

0
−0.5
−1
−1.5
−2
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)
b) 1
0.5
Torque (pu)

0
−0.5
−1
−1.5
−2
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)
c) 1.5
1
(pu)

0.5
0
sd
ψssα,ψsf

−0.5
−1
−1.5
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)
d) 1
0.75
0.5
(pu)

0.25
0
sd
isf

−0.25
−0.5
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)
e) 1
0.5
0
(pu)

−0.5
sq

−1
isf

−1.5
−2
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)
f) 1.5
1
issα (pu)

0.5
0
−0.5
−1
−1.5
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)

Figure 7.3: Induction machine variables during speed reversal. (a) Rotor speed. (b) Ele-
cromechanical torque. (c) Stator flux α-component and stator flux amplitude. (d) Reactive
stator current. (e) Active stator current. (f) Stator current β-component.

83
a) 1.76
1.75
udc (pu)

1.74
1.73
1.72
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)
b) 1.5
1
p (pu)

0.5
0
g

−0.5
−1
−1.5
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)
c) 0.6
0.4
q (pu)

0.2
0
g

−0.2
−0.4
−0.6
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)
d) 0.6
0.4
igf (pu)

0.2
0
gd

−0.2
−0.4
−0.6
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)
e) 1.5
1
igf (pu)

0.5
0
gq

−0.5
−1
−1.5
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)
f) 1.5
1
is (pu)

0.5
0

−0.5
−1
−1.5
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)

Figure 7.4: Grid variables during speed reversal. The stator power is fed-forward to the
direct voltage control and the stator power is limited. (a) Direct voltage. (b) Active grid
power. (c) Reactive grid power. (d) Reactive grid current. (e) Active grid current. (f) Grid
current β-component.

84
1.5

1
Start

0.5 Maximal
transient
Speed (pu)

torque 2 1
0
3 4
−0.5
Maximal
transient
power
−1
Max. speed End
X
−1.5
−1.5 −1 −0.5 0 0.5 1 1.5
Torque (pu)

Figure 7.5: Speed-torque trajectory during speed reversal. The induction machine operates
in all four quadrants.

85
Fig. 7.6 shows the grid side variables for the same speed reversal as in Figs. 7.3
and 7.4. However, no feed-forward term for the direct voltage control is used in
Fig. 7.6 and the stator power is not limited. As shown in Fig. 7.6a, the direct
voltage experiences high variations. The variations are mainly due to the fact
that no feed-forward term is used in the direct voltage control. Furthermore,
the grid current amplitude is limited to 1.5 p.u. at t=0.3 s, hence the grid
side PWM converter can only supply an active power of 1.5 p.u. to the direct
voltage link. The induction machine draws more than 1.5 p.u. active power
from the direct voltage link due to the absence of the stator power limit, and
the direct voltage therefore decreases at t=0.3 s.

86
a) 2
1.9
udc (pu)

1.8
1.7
1.6
1.5
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)
b) 1.5
1
0.5
pg (pu)

0
−0.5
−1
−1.5
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)
c) 0.6
0.4
qg (pu)

0.2
0
−0.2
−0.4
−0.6
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)
d) 0.6
0.4
(pu)

0.2
0
gd
igf

−0.2
−0.4
−0.6
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)
e) 1.5
1
(pu)

0.5
0
gq
igf

−0.5
−1
−1.5
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)
f) 1.5
1
isgα (pu)

0.5
0
−0.5
−1
−1.5
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)

Figure 7.6: Grid variables during speed reversal. The stator power is not fed-forward to the
direct voltage control and the stator power is not limited. (a) Direct voltage. (b) Active grid
power. (c) Reactive grid power. (d) Reactive grid current. (e) Active grid current. (f) Grid
current β-component.

87
Fig. 7.7 shows that the grid connected PWM converter can operate at an
arbitrary power factor. Furthermore, the grid voltage sag of 0.2 p.u. at
t=0.95 s hardly affects the direct voltage or the active and reactive grid power.
The induction machine variables are not shown, but the induction machine
operates in steady state with a load torque of 0.5 p.u.

88
1.744
1.742
udc (pu)

1.74
1.738
1.736
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 0.11 0.12 0.13 0.14
Time (s)
b) 0
−0.1
pg (pu)

−0.2
−0.3
−0.4
−0.5
−0.6
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 0.11 0.12 0.13 0.14
Time (s)
c) 0.6
0.4
qg (pu)

0.2
0
−0.2
−0.4
−0.6
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 0.11 0.12 0.13 0.14
Time (s)
d) 0.6
0.4
(pu)

0.2
0
gd
igf

−0.2
−0.4
−0.6
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 0.11 0.12 0.13 0.14
Time (s)
e) 0
−0.1
(pu)

−0.2
−0.3
gq
igf

−0.4
−0.5
−0.6
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 0.11 0.12 0.13 0.14
Time (s)
1.5
1
isgα,esgα (pu)

0.5
0
−0.5
−1
−1.5
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.1 0.11 0.12 0.13 0.14
Time (s)

Figure 7.7: Grid variables during power factor control and grid voltage dip. (a) Direct
voltage. (b) Active grid power. (c) Reactive grid power. (d) Reactive grid current. (e)
Active grid current. (f) Grid current and grid voltage β-components.

89
Fig. 7.8 shows grid variables when the grid side PWM converter operates
as both a direct voltage regulator and as a parallel active power filter. The
grid side PWM converter now uses a vector current controller with dead-beat
gains [7] and the sampling frequency is increased to 9.6 kHz, i.e. twice the
switching frequency. The direct voltage is filtered in the control system with a
first order low-pass filter, cut-off frequency of 30 Hz. Fig. 7.8d shows the grid
converter current, Fig. 7.8e shows the harmonic load current and Fig. 7.8f
shows the resulting grid current. The instantaneous reactive power method,
proposed in [60], is used to detect the load current harmonics. Note that
only harmonics that contribute to the ac instantaneous reactive power are
detected and cancelled, as shown in Fig. 7.8b. The grid current will therefore
not be perfectly sinusoidal. On the other hand, the harmonics that contribute
to the ac instantaneous reactive power can be detected and cancelled instan-
taneously [60]. The pulsating instantaneous reactive power creates a small
300 Hz pulsation in the direct voltage. The induction machine variables are
not shown, but the induction machine is speed controlled and the constant
torque load increases from 0.5 p.u. to 1.0 p.u. at t=0.1 s. The active grid
current in Fig. 7.8c therefore changes at t=0.1 s in order to maintain a power
balance on the direct voltage link. The grid side PWM converter voltage
saturates for each flank in the harmonic load current. The voltage saturation
introduces cross coupling between the reactive grid current and the active grid
current. The voltage saturation also results in the grid side PWM converter
not being able to track the current harmonics perfectly and current spikes
occur in the otherwise nearly sinusoidal grid current.

90
a) 1.748
1.744
udc (pu)

1.74
1.736
1.732
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.22
Time (s)
b) 1.8
1.2
(pu)

0.6
0
gd
igf

−0.6
−1.2
−1.8
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.22
Time (s)
c) 1.8
1.2
(pu)

0.6
0
gq
igf

−0.6
−1.2
−1.8
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.22
Time (s)
d) 1.8
1.2
ig1 (pu)

0.6
0
−0.6
−1.2
−1.8
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.22
Time (s)
e) 1.8
1.2
(pu)

0.6
0
iload

−0.6
1

−1.2
−1.8
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.22
Time (s)
f) 2.4
1.6
(pu)

0.8
0
igrid

−0.8
1

−1.6
−2.4
0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 0.22
Time (s)

Figure 7.8: Grid variables during active power filtering operation. (a) Direct voltage. (b)
Reactive converter current. (c) Active converter current. (d) Converter phase current. (e)
Harmonic load phase current. (f) Grid phase current.

91
92
Chapter 8

Experiments

This chapter presents the experimental results of the developed stator flux
orientated vector control of an induction machine. The control of the grid
connected PWM converter is not studied here and the direct voltage is charged
from a dc-generator. A brief description of the laboratory hardware is given.

8.1 Hardware Description


This section describes the hardware that was used during the experiments.
Fig 8.1 shows a coarse block diagram of the experimental arrangement. Fig 8.2
to Fig 8.4 show photographs of the experimental arrangement.

8.1.1 Double-Sided PWM Converter


The double-sided PWM converter uses Semikron SKIPPACK:s 342GD120-
314CTV that are rated 1200 V/300 A. The bridges are equipped with their
own integrated drivers. The bridges are also equipped with their own temper-
ature and current sensors that are used for protection. Additional hardware
for protecting the bridges against overcurrent and overvoltage on the direct
voltage link was also used in the laboratory arrangement. Fiber optics were
used for galvanic insulation between drivers and the bridges. Only the ma-
chine side PWM converter was used in the experiments in this chapter.

Direct Voltage Link


The direct voltage were set at 570 V in the experiments in this chapter. The
direct voltage link can be charged in two ways, either via the grid side PWM
converter or via a separate dc-source. In the following experiments, the direct
voltage link is charged from a dc-generator and the grid side PWM converter
is not used. The dc-generator has quite slow direct voltage control dynamics
and large direct voltage variations were encountered during the experiments.

93
Synchronous DC
machine machine
~ =
D B 4 ;E94 ;7F<=>-?03@6A71B C 71B
G 6@C8H971;I -/.103254 687:94 ;7:<=>-?01@6A73B C 71B

Induction DC
is us
machine Resolver Tacho- machine
meter
=
df
e
ir d
a
ur bc Thyristor
^ _`a
converter
i dc u dc
J KMLON <=-QP G R @8;5H8S .1C @B I

3
IEA-MIMO
8 3
<TVU @B R 73.19HB 7 R 736C 9 G W .1XYZ4 71[OI

<TVU @8B\T]N
Post-processing
@U\;.1C .
programming

Figure 8.1: Block diagram of experimental arrangement.

Figure 8.2: Photograph of experimental arrangement. Control computer (middle), PC for


C-programming (middle), measurement PC and sample/hold/filter box (left), double-sided
PWM converter (right) and induction machine/dc machine (background).

94
Figure 8.3: The experimental arrangement from another perspective.

Figure 8.4: The induction machine (right) and the loading dc-machine.

95
8.1.2 Induction Machine

The induction machine was a 22 kW ASEA MAG 180L 55-4 with wounded
rotor. During the experiments, the rotor of the 22 kW machine was short-
circuited and no external rotor resistance was used. It is possible to measure
the rotor currents in order to evaluate different flux observers. However, the
rotor currents have not been used at all in the control system during the
experiments. The induction machine parameters have been calculated from
standard no-load and locked rotor tests. The power cable resistance has been
included in the stator resistance.

A dc-machine has been used to load the induction machine. The dc-machine
was fed by a thyristor converter.

Unshielded power cables have been used to feed the induction machine and
the dc-machine. The unshielded power cables are believed to have a negative
effect on the noise level of the experimental system. The length of the power
cables that feed the induction machine were roughly 4 m. Inductor filters
of 0.5 mH have been placed between the PWM converter and the induction
machine in order to reduce stresses on the stator winding insulation.

8.1.3 Control Computer

The control system of the induction machine was implemented in a IEA-


MIMO control computer [11]. The IEA-MIMO, developed by Anders Carlsson
at Lund University of Technology, is based on a Texas TMS320c30 floating
point signal processor. The IEA-MIMO is clocked by a 40 MHz crystal os-
cillator and has eight analog outputs, a 16-bit DSPLINK expansion bus for
input/output and permits simultaneous sampling of up to 16 analogue chan-
nels. The 16 analogue inputs were filtered with first order low-pass RC-filters.
The cut-off frequency of the filter is 45 kHz, so the filters had minor influ-
ence. The A/D-conversion of the 16 channels is carried out by one 12-bit
flash A/D-converter and takes about 16 µs. The TMS320c30 has 2 kWords
(2048x32 bits) of on-chip memory and the IEA-MIMO has two blocks of ad-
ditional external memory, namely 128 kWords static RAM and 256 kWords
ROM. The IEA-MIMO can easily be programmed with the C language or the
assembly language with the aid of an ordinary PC. The control computer was
mounted in a rack together with input/output cards, overvoltage/overcurrent
protection cards, resolver card, which were all developed at the department.
Even the modulator card has been mounted in the rack. All control cables
were shielded and grounded at one end.

96
8.1.4 Pulse Width Modulation
The pulse width modulator was a digital IEA-PWM6 [11]. The modulator
uses triangular PWM and can modulate up to six phases synchronously with
a common 10-bit resolution carrier triangular wave. Hence, the modulator
can modulate both a grid side PWM converter and a machine side PWM
converter synchronously. The modulator was linked to the control computer
via the expansion bus. The modulator also generates interrupts for the control
computer. The interrupts are synchronized with the triangular carrier wave
and are used to trigger the simultaneous regular sampling of 16 channels.

8.1.5 Sensors
The experimental system has been equipped with a large amount of sen-
sors. This provides a flexible system and the possibility to perform several
interesting future experiments. The control computer used only a few of the
available sensors during the experiments in this chapter. More specific, the
control computer used stator current sensors, rotor position sensor and direct
voltage sensor.

Current Sensors
Currents have been measured with 300 A LEM-modules. The LEM-modules
for measuring phase currents were wounded with three turns while the LEM-
module on the direct voltage link was wounded with one turn.

Voltage Sensors
Voltages have been measured with Analog Devices AD210. Observe that the
control system did not use measured stator voltages. Reference voltages were
used in the software stator flux observers.

Position/Speed Sensor
The induction machine has been equipped with both a resolver and a dc-
tachometer. Only the resolver signal enters the control computer. The re-
solver circuitry is based on an Analog Devices AD2S80A resolver to digital
converter. The resolver circuitry was configured to 12 bits and the reference
frequency was 5 kHz. This resolver configuration allows a theoretical speed
range of 70-18750 rpm, given a sampling frequency of 4.9 kHz. The plan is to
modify the resolver circuitry to 16 bits within a near future. The rotor posi-
tion entered the control computer via the expansion bus. The dc tachometer
was not used by the control system, only by the measurement system.

97
Torque Sensor
The shaft torque has been measured with strain gauges and a HBM torque
transducer and amplifier.

Temperature Sensors
Four PT-100 sensors have been placed in the stator winding of the induction
machine. The sensors are intended to provide a rough estimation of the
thermal conditions.

8.1.6 Measurement System


The measurement system was used to collect experimental data for later post-
processing. The measurement system used LabView and National Instru-
ments SCXI sample/hold modules and programmable low-pass filters. The
maximum sampling frequency with 16 measured variables is 35 kHz. The
sampling frequency during the experiments was 1.5 kHz and the cut-off fre-
quency of the filter was 400 Hz.

8.2 Software Description


The software implementation is essentially as described in the previous chap-
ters in this thesis. However, some minor control changes have been imple-
mented. Measurement offsets in the current sensors were stored before actual
operation started and compensated by the control program during actual op-
eration. No torque was applied until the machine was fully magnetized. The
control program compensated for the blanking time and on-state voltage drop
of the semiconductor valves and the direct voltage ripple was also compen-
sated.
The Luenberger with nearly constant poles was used for stator flux feedback.
The parameters of the observer and the control system are shown in Table 8.1.
The stator flux reference below base speed was constant and set at the nom-
inal stator flux linkage. The software limits for maximum transient torque,
current amplitude and stator power were set at 218 Nm, 100 A and 30 kW,
respectively.
The current decoupling term in the stator flux controller was empirically
weighted with 0.85, hence the influence from the decoupling term was re-
duced. The weighting factor resulted in less stator flux variations during
torque transients in the experiments.
The implementation of the control computer and the modulator resulted in
a one sample calculational delay for this system. The calculation delay was

98
not considered in the current control. However, it has been found useful
to assume uss (k) = us?
s (k − 1) since reference voltages were used instead of
measured stator voltage in the stator flux observers.
The resolver provides the rotor position only. A speed observer [61] has been
used to derive the rotor speed from the measured rotor position. The speed
observer were also capable of estimating the load torque.
The execution time of the control program including A/D-conversion was
approximately 120 µs, i.e. 59 % of the sampling period. No special efforts
have been carried out in order to make the control code more time efficient.
Table 8.1 summarizes the parameters of the experimental arrangement and
the control system parameters.

8.3 Laboratory Experiences with Stator Flux Observers


Four stator flux observers were tested experimentally. These four observers,
described earlier in Chapter 4, were the traditional voltage model, the im-

Table 8.1: Parameters of experimental arrangement.


Nominal values of the induction machine:
Stator voltage (Y) 380 V
Stator current (Y) 44 A
Frequency (electrical) 50 Hz
Rotor speed 1440 rpm
Torque 145 Nm
Power 22 kW
Power factor 0.89
Induction machine parameters:
Rotor resistance 0.137 Ω ¿ 0.0274 p.u.
Stator resistance 0.125 Ω ¿ 0.0250 p.u.
Rotor leakage inductance 1.85 mH ¿ 0.1167 p.u.
Stator leakage inductance 1.14 mH ¿ 0.0718 p.u.
Mutual inductance 39.8 mH ¿ 2.51 p.u.
Rotor inertia 0.334 kgm2
PWM Converter:
Semikron SKIPPACK 342GD120-314CTV 1200V/300A
Switching frequency 4.9 kHz ¿ 98.0 p.u.
direct voltage link:
Nom./max. direct voltage 570/800 V ¿ 1.5/2.1 p.u.
Capacitor bank 20.0 mF ¿ 1.5 p.u.
Control system parameters:
Sampling frequency 4.9 kHz ¿ 98.0 p.u.
Closed-loop current bandwidth 150.0 Hz ¿ 3.0 p.u.
Closed-loop stator flux bandwidth 5.0 Hz ¿ 0.1 p.u.
Luenberger observer, |λsc |/|λsf | 17.5/0.5-7.5 Hz ¿ 0.35/0.01-0.15 p.u.

99
proved integrator algorithm 2 for the voltage model, the current and voltage
model and the Luenberger observer with nearly constant poles. The volt-
age model did not work even though it was tried in order to compensate
for measurement offset. Uncompensated measurement offset quickly caused
the stator flux estimate to drift. The remaining three stator flux observers
worked satisfactorily though. No detailed comparison of the three functional
observers was carried out. The Luenberger observer with nearly constant
poles was used in the following section that contains experimental results.

8.4 Experimental Results


Fig 8.5 shows the experimental results when the torque reference is controlled
by a hysteres function of rotor speed. The torque reference is reversed each
time the rotor speed passes through ±1000 rpm. The machine is not loaded
with any external torque. Note that the stator flux amplitude and the torque
are estimated quantities. The torque response is fast and accurate but minor
torque overshoots seem to occur. The stator flux and the torque are nearly
decoupled but disturbances can be spotted in the stator flux during the torque
transients. Some minor oscillations in the torque are encountered when the
speed passes zero rpm. It is unclear if this depends on the stator flux feedback
or a too low resolver resolution.

100
a) 1500
1000
n (rpm)

500
0
−500
−1000
−1500
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)
b) 80
40
T (Nm)

0
e

−40
^

−80
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)
c) 1.2
1.1
ψsd (Vs)

1
^ sf

0.9
0.8
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)
d) 40
35
30
(A)

25
sd
isf

20
15
10
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)
e) 40
20
isf (A)

0
sq

−20
−40
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)
f) 40
20
is (A)

0

−20
−40
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)

Figure 8.5: Induction machine variables during steps in the torque. (a) Rotor speed. (b)
Estimated torque. (c) Estimated stator flux amplitude. (d) Reactive stator current. (e)
Active stator current. (f) Stator current β-component.

101
Fig 8.6 shows the results of an experiment in which steps in the stator flux
reference are carried out. The reference steps have been generated by an
external signal generator that is connected to one of the input cards. The
machine was not loaded with any external torque. Observe that the stator
flux steps have been carried out below nominal stator flux linkage, i.e., the
saturation level has changed slightly compared to the nominal case. The
stator flux control works well and the torque remains fairly constant. The
response time (10%-90%) of the stator flux amplitude is approximately 60 ms.

102
a) 600
580
n (rpm)

560
540
520
500
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)
b) 20
15
Te (Nm)

10
5
^

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)
c) 1
0.9
ψsd (Vs)

0.8
^ sf

0.7
0.6
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)
d) 40
30
(A)

20
sd
isf

10
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)
e) 20
15
(A)

10
sq
isf

5
0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)
f) 40
20
is (A)

0

−20
−40
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Time (s)

Figure 8.6: Induction machine variables during steps in the stator flux amplitude. (a) Rotor
speed. (b) Estimated torque. (c) Estimated stator flux amplitude. (d) Reactive stator
current. (e) Active stator current. (f) Stator current β-component.

103
Fig. 8.8 shows the results of an experiment in which the rotor has been ac-
celerated from standstill with 150 % of nominal torque, i.e., 218 Nm. The
machine was not loaded with any external torque. The torque response is
very good. The system enters the power limit of 30 kW at t =0.45 s and the
torque is reduced. Flux weakening is applied when the speed passes 1340 rpm
and the stator flux is reduce. The final speed is 1730 rpm, i.e. 120 % of nom-
inal speed. Higher speed than this was not tested due to concerns for the
wounded rotor windings.
Fig. 8.7 shows a closer view of the estimated torque and active stator current
when 150 % of nominal torque is applied. The torque response time (10%-
90%) is approximately 3 ms. The fast torque response is because no back-
emf is present and nearly the whole converter output voltage can be used to
increase the active current.

220
200
Est. torque (Nm), active stator current (A)

180
160
140
120
100
80
60
40
Est. torque
20 Active current
0
0.14 0.145 0.15 0.155 0.16 0.165 0.17
Time (s)

Figure 8.7: Closer view of estimated torque and active stator current during acceleration
from standstill with 150 % of nominal torque.

104
a) 2000
1500
n (rpm)

1000
500
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)
b) 250
200
Te (Nm)

150
100
50
0
−50
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)
c) 1.4
1.2
ψsd (Vs)

1
sf

0.8
0.6
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)
d) 60
50
40
isd (A)

30
sf

20
10
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)
e) 100
80
60
(A)

40
sq
isf

20
0
−20
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)
f) 100
50
issβ (A)

0
−50
−100
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 0.55 0.6 0.65
Time (s)

Figure 8.8: Induction machine variables during acceleration from standstill with 150 % of
nominal torque. (a) Rotor speed. (b) Estimated torque. (c) Estimated stator flux amplitude.
(d) Reactive stator current. (e) Active stator current. (f) Stator current β-component.

105
Low speed operation with nominal torque was investigated experimentally.
The torque reference in the experiment was set by an outer speed control
loop. The speed controller was designed as recommended in [19]. The closed-
loop bandwidth of the speed loop was set 100 times below the closed-loop
bandwidth of the current loop. The induction machine was loaded with nom-
inal torque and the speed was decreased until controlled speed operation was
no longer possible.
The minimum speed when using the 12 bit resolver and nominal load torque
was about 60 rpm, i.e. close to the theoretical value of 70 rpm. A simple
speed estimator [17] was implemented in order to decrease the minimum speed
further. The speed estimator is based on the derivative of the stator flux
position minus the slip speed [17], hence
" Ã !# " #
d ψbs
1 d b
ψ s
sβ sβ
ω
bsf = tan−1 = Ã ! 2
dt ψbsα
s
ψbsβ
s dt ψbsα s
1+
ψbsα
s

dψbsβ
s
dψbsα
s
bs
ψsα − ψbsβ
s
bs is )ψbsα bs issα )ψbs
dt dt (ussβ − R sβ
s
− (ussα − R sβ
= = (8.1)
b s|
|ψ 2 b s|
|ψ 2

bs isf
L sq
ω
bsl = [steady state] = (8.2)
τbr (ψbsd − L
sf bσ isf )
sd
ω
b = ωbsf − ω bsl (8.3)
The speed estimator operated rather well but the speed estimate was noisy
when using the speed estimator in this experimental system. The speed es-
timate was therefore filtered with a first-order low-pass filter. The cut-off
frequency of the low-pass filter was set 10 times above the closed-loop band-
width of the speed control loop.
Fig. 8.9 shows the results of the sensorless experiment. The induction ma-
chine is loaded with nominal torque, 145 Nm. The resolver and the dc-
tachometer are not used by the control computer. The speed is instead esti-
mated with (8.3). The minimum speed with nominal torque is approximately
10 rpm. The dc-tachometer signal is very noisy and it could not be deter-
mined whether the speed pulsations in Fig. 8.9a were due to noise, due to the
control system or due to the loading dc-machine and thyristor converter. A
torque reference step of 110 % of nominal torque was applied at t = 0.5 s. As
shown, the torque control is still good at this low speed.

106
a) 50
40
n (rpm)
30
20
10
0
−10
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time (s)
b) 170
165
Te (Nm)

160
155
150
^

145
140
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time (s)
c) 1
0.98
^ sf (Vs)

0.96
0.94
sd
ψ

0.92
0.9
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time (s)
d) 35
33
isd (A)

31
29
sf

27
25
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time (s)
e) 60
56
isq (A)

52
48
sf

44
40
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time (s)
f) 80
40
is (A)

0

−40
−80
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Time (s)

Figure 8.9: Induction machine variables during sensorless low speed operation. (a) Rotor
speed. (b) Estimated torque. (c) Estimated stator flux amplitude. (d) Reactive stator
current. (e) Active stator current. (f) Stator current β-component.

107
Fig. 8.10 shows an experimental magnetization of the machine. The max-
imum reactive stator current is limited to 50 A in this experiment. The
magnetization takes approximately 0.5 s.
It was observed that disturbances entered the sensors and the control com-
puter when the PWM converter was switched on. The disturbances occasion-
ally caused the stator flux estimate to drift very quickly. A software limit
was therefore placed on the stator flux estimate’s α- and β-components. The
limit was set to 2 Vs.
a) 1.2
1
ψsd (Vs)

0.8
0.6
^ sf

0.4
0.2
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Time (s)
b) 60
50
40
isd (A)

30
sf

20
10
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
Time (s)

Figure 8.10: Magnetization of the induction machine. (a) Stator flux amplitude. (b) Reactive
stator current.

108
Chapter 9

Conclusions

A vector control scheme for the double-sided PWM converter has been de-
veloped. The induction machine uses stator flux orientated vector control
with a synchronous PI-controller. Stator flux orientated vector control has
low parameter sensitivity in the medium/high speed region but unfortunately
also has a cross coupling between torque and stator flux. A decoupling term
has been developed for the cross coupling, and computer simulations as well
as laboratory experiments, show that the decoupling term is successful. A
Luenberger observer variant has been developed for stator flux feedback. The
poles of the observer imitate the poles of the induction machine but are better
damped.
The grid connected PWM converter is controlled by a grid flux orientated
vector control scheme that uses a synchronous PI-controller. The grid con-
nected PWM converter controls the grid current and the direct voltage. A
feed-forward term that is proportional to the induction machine loading is
used in the direct voltage control. Computer simulations show that the direct
voltage remains fairly constant even for heavy torque transients when using
the feed-forward term. The fairly constant direct voltage is also due to a
power limit on the stator power of the induction machine.
Laboratory experiments with a PWM converter and 22 kW induction machine
have verified a speed range of 170:1 and a torque response time of 3 ms at
stand still. Furthermore, a torque capability below base speed of at least
150 % of nominal torque has also been verified. Torque and stator flux have
been controlled nearly independently.
The parameter sensitivity or noise sensitivity of the developed control schemes
has not been studied in great detail and this remains to be done. One can
especially expect the parameters of the induction machine to vary depending
on temperature and saturation.
A similar control scheme as described in this thesis, but more refined and
adapted to a wind turbine application, will soon be tested on a 600 kW
variable-speed stall regulated wind turbine.

109
110
Chapter 10

Future Work

The author recommends the following topics for future work on the double-
sided PWM converter.
Study the possibility to reduce the number of components of a double-sided
PWM converter. A reduced number of components is likely to reduce the cost
and size of a double-sided PWM converter and can also increase reliability.
For instance, one method to reduce the number of components is to reduce
the number of sensors. It is also recommended to compare different speed-
sensorless algorithms.
Improve the control, and especially the co-ordinated control of the grid side
PWM converter and the machine side PWM converter, so that operation with
a dc-capacitor of only a few microfarads becomes troublefree.
The control system of the double-sided PWM converter should be able to
handle any kind of situation with built-in control algorithms, for instance fault
detection, fault handling, overload control, self-commisioning, line transients
and an unbalanced grid.
Improve the total drive efficiency when using a double-sided PWM converter.
Find a good balance between converter losses, motor losses and filters. De-
velop a control method that optimizes the efficiency. Study the influence from
different PWM methods and switching frequencies on the drive efficiency.
Decrease the dependency on the stator resistance for direct stator flux orien-
tation at low speeds. The dependency on the stator resistance can for instance
be obtained by estimating the stator resistance on-line at low speeds, or by
designing a stator flux observer with less dependency on the stator resistance.
Investigate the differences between stator flux orientation and rotor flux ori-
entation in more detail.

111
112
References

[1] Ion Boldea and S.A. Nasar, Electric Drives, CRC Press, Boca Raton,
1999.
[2] Dan Jones, “Global motion market trends,” Lecture notes for a tutorial
held at PCIM’98, Nürnberg, Germany, 1998.
[3] Naoya Eguchi and Teruo Imura, “Self-commutated inverter for fuel cell
power plant,” in Proc. IAS 1986, Denver, USA, Sep. 28-Oct. 3 1986,
vol. 1, pp. 527–532.
[4] Helmut Kohlmeier, Olaf Niermeyer, and Dierk F. Schröder, “Highly
dynamic four-quadrant ac motor drive with improved power factor and
on-line optimized pulse pattern with promc,” IEEE Transactions on
Industry Applications, vol. 23, no. 6, pp. 1001–1009, Nov./Dec. 1987.
[5] Helmut Kohlmeier and Dierk F. Schröder, “Control of a double voltage
inverter system coupling a three phase mains with an ac-drive,” in Proc.
IAS 1987, Atlanta, USA, Oct. 18-23 1987, vol. 1, pp. 593–599.
[6] Joachim Holtz, Pulse Width Modulation for Electronic Power Conver-
sion, chapter 4, pp. 138–208, In [63], 1997.
[7] Tore Svensson, On modulation and control of electronic power convertors,
Technical report no. 186, School of Electrical and Computer Engineering,
Chalmers University of Technology, Göteborg, Sweden, 1988.
[8] Frede Blaabjerg and John K. Pedersen, “An integrated high power fac-
tor three-phase ac-dc-ac converter for ac-machines implemented in one
microcontroller,” in Proc. PESC 1999, Seattle, USA, 1993, vol. 1, pp.
285–292.
[9] Mats Alaküla and John-Eric Persson, “Vector controlled ac/ac converters
with a minimum of energy storage,” in Proc. IEE Fifth International
Conference on Power Electronics and Variable-Speed Drives, London,
UK, Oct. 26-28 1994, vol. 1, pp. 236–239.
[10] N. J. Wheeler, H. Zhang, and D. A. Grant, “Minimization of reactive
component values in back-to-back converters,” in Proc. IEE Fifth In-

113
ternational Conference on Power Electronics and Variable-Speed Drives,
London, UK, Oct. 26-28 1994, vol. 1, pp. 240–245.
[11] Anders Carlsson, “The back-to-back converter - control and design,”
1998.
[12] Steffan Hansen, Mariusz Malinowski, Frede Blaabjerg, and Marian P.
Kaźmierkowski, “Sensorless control strategies for pwm rectifier,” in Proc.
APEC 2000, New Orleans, USA, 2000, vol. 2, pp. 832–838.
[13] Felix Blaschke, “The principle of field orientation as applied to the
new TRANSVEKTOR closed-loop control system for rotating-field ma-
chines,” Siemens Review, vol. 34, pp. 217–220, May 1972.
[14] K. Hasse, “Drehzahlgelverfahren für schnelle umrichterantriebe
mit stromrichtergespeisten asynchron - kurzschlussläufermotoren,”
Regelungstechnik, pp. 60–66, 1972.
[15] Xingyi Xu, Rik De Doncker, and Donald W. Novotny, “Stator flux ori-
entation control of induction machines in the field weakening region,”
in Proc. IAS Annual Meeting 1988, Pittsburgh, USA, 1988, vol. 1, pp.
437–443.
[16] Xingyi Xu, Rik De Doncker, and Donald W. Novotny, “A stator flux
orientated induction machine drive,” in Proc. PESC 1988, Kyoto, Japan,
1988, vol. 2, pp. 870–876.
[17] Xingyi Xu and Donald W. Novotny, “Implementation of direct stator
flux orientation on a versatile dsp based system,” IEEE Transactions on
Industry Applications, vol. 27, no. 4, pp. 694–700, July/Aug. 1991.
[18] Xingyi Xu and Donald W. Novotny, “Selection of the flux reference for
induction machine drives in the field weakening region,” IEEE Transac-
tions on Industry Applications, vol. 28, no. 6, pp. 1353–1358, Nov./Dec.
1992.
[19] Lennart Harnefors, On Analysis, Control and Estimation of Variable-
Speed Drives, Ph.D. thesis, Royal Institute of Technology, Stockholm,
Sweden, 1997.
[20] Patrick L. Jansen and Robert D. Lorentz, “A physically insightful ap-
proach to the design and accuracy assessment of flux observers for field
orientated induction machine drives,” IEEE Transactions on Industry
Applications, vol. 30, no. 2, pp. 101–110, Jan./Feb. 1994.
[21] “Braking and regenerative energy with ac drives,” Rockwell Automation
Appl. Note, 1999.

114
[22] J. Jiang and J. Holtz, “An efficient braking method for controlled ac
drives with a diode rectifier front end,” in Proc. IAS 2000, Rome, Italy,
Oct. 7-12 2000, vol. 3, pp. 1137–1143.
[23] Annabelle van Zyl, René Spée, Alex Faveluke, and Shibashis Bhowmik,
“Voltage sag ride-through for adjustable-speed drives with active recti-
fiers,” IEEE Transactions on Industry Applications, vol. 34, no. 6, pp.
179–186, Nov./Dec. 1998.
[24] Ed. Wayne L. Stebbins, Select & Apply ASD Systems, IEEE Press, New
Jersey, USA, 2000.
[25] D. E. Rice, “Adjustable speed drive and power rectifier harmonics - their
effect on power system components,” in Proc. PCIC’95, San Francisco,
USA, Sept. 10-12 1984, vol. 1, pp. 269–287.
[26] Johann W. Kolar, “Three-phase unity power factor ac/dc converter sys-
tems,” Lecture notes for a tutorial held at IEEE Nordic Workshop on
Power and Industrial Electronics 2000, Aalborg, Denmark, 1998.
[27] A. R. Prasad, Phoivos D. Zioga, and Stefanos Manias, “An active power
factor correction technique for three-phase diode rectifiers,” IEEE Trans-
actions on Power Electronics, vol. 6, no. 1, pp. 83–92, Jan. 1991.
[28] Steffan Hansen, Peter Nielsen, and Paul Thogersen, “Harmonic distorsion
and reduction techniques of pwm adjustable speed drives - a cost-benefit
analysis,” in Proc. IEEE Nordic Workshop on Power and Industrial
Electronics 2000, Aalborg, Denmark, 2000, vol. 1, pp. 271–277.
[29] F. Abrahemsen and A. David, “Adjustable speed drive with active fil-
tering capability for harmonic current compensation,” in Proc. IEEE
PESC’95, Atlanta, USA, June 18-22 1995, vol. 2, pp. 1446–1453.
[30] Walter A. Hill and Suresh C. Kapoor, “Effect of two-level pwm sources
on plant power system harmonics,” in Proc. IEEE Industry Applications
Conference 1998 (IAS’98), St. Louis, USA, 1998, vol. 2, pp. 1300–1306.
[31] Zhong Erkuan, Chen Shaotong, and Thomas Lipo, “Improvements in
emi performance of inverter-fed motor drives,” in Proc. APEC 1994,
Orlando, USA, 1994, vol. 2, pp. 608–614.
[32] Frede Blaabjerg, John K. Pedersen, Ulrik Jaeger, and Paul Thoegersen,
“Single current sensor technique in the dc link of three-phase pwm-vs
inverters: A review and a novel solution,” IEEE Transactions on Industry
Applications, vol. 33, no. 5, pp. 1241–1253, Sep./Oct. 1997.
[33] Bengt Schmidtbauer, Analog och Digital Reglerteknik, Studentlitteratur,
Lund, Sweden, 1995.

115
[34] Ryuzo Ueda, Toshikatsy Sonoda, Kunio Koga, and Michiya Ichikawa,
“Stability analysis in induction motor driven V/f controlled general-
purpose inverter,” IEEE Transactions on Industry Applications, vol.
28, no. 2, pp. 472–481, March/April 1992.
[35] Donald W. Novotny and Thomas A. Lipo, Vector Control and Dynamics
of AC Drives, Clarendon Press, Oxford, 1998.
[36] Pál K. Kovács, Transient Phenomena in Electrical Machine, Elsevier,
Amsterdam, 1984.
[37] Gordon R. Slemon, “Modelling of induction machines for electric drives,”
IEEE Transactions on Industry Applications, vol. 25, no. 6, pp. 1126–
1131, Nov./Dec. 1989.
[38] Werner Leonhard, Control of Electric Drives, Springer-Verlag, Berlin,
1996.
[39] Karl-Erik Hallenius, Elektriska Maskiner, CWK Gleerup Bokförlag,
Lund, Sweden, 1972.
[40] R.D. Lorentz, T.A Lipo, and D.W. Novotny, Motion Control with In-
duction Motors, chapter 5, pp. 209–276, In [63], 1997.
[41] Robert D. Lorenz and Donald W. Novotny, “Optimal utilization of in-
duction machines in field orientated drives,” in Proc. Electric Energy
Conference 1987, Adelaide, Australia, 1987, vol. 1, pp. 259–264.
[42] Ian T. Wallace, Donald W. Novotny, Robert D. Lorentz, and Deep-
akraj M. Divan, “Verification of enhanced dynamic torque per ampere
in saturated induction machines,” IEEE Transactions on Industry Ap-
plications, vol. 30, no. 5, pp. 1193–1201, Sep./Oc. 1994.
[43] Bo Peterson, Induction Machine Speed Estimation - Observations on
Observers, Ph.D. thesis, Lund Institute of Technology, Lund, Sweden,
1996.
[44] Mats Alaküla and Anders Carlsson, “An induction machine servo with
one current controller and an improved flux observer,” in Proc. IECON
1993, Maui, USA, 1993, vol. 3, pp. 1991–1996.
[45] J. Hu and B. Wu, “New integration algorithms for estimating motor flux
over a wide speed range,” in Proc. PESC 1997, St. Louis, USA, 1997,
vol. 2, pp. 1075–1081.
[46] Patrick L. Jansen, Robert D. Lorentz, and Donald W. Novotny,
“Observer-based direct field orientation: Analysis and comparison of al-
ternative methods,” IEEE Transactions on Industry Applications, vol.
30, no. 2, pp. 945–953, July/Aug. 1994.

116
[47] Jouko Niiranen, “Efficient and accurate symmetric euler algorithm for
electromechanical simulations,” in Proc. IEEE Nordic Workshop on
Power and Industrial Electronics 1998, Helsinki, Finland, 1998, vol. 1,
pp. 229–234.
[48] Vladimir Blasko, “Adaptive filtering for selective elimination of higher
harmonics from line current of a voltage source converter,” in Proc. IAS
1998, St. Louis, USA, 1998, vol. 2, pp. 1222–1280.
[49] H. S. Kim, H. S. Mok, G. H. Choe, D. S. Hyun, and S. Y. Choe, “Design
of current controller for 3∼phase pwm converter with unbalanced input
voltage,” in Proc. IEEE PESC 1998, Fukuoka, Japan, 1998, vol. 1, pp.
503–509.
[50] Martin Bojrup, “Advanced control of active filters in a battery charger
application,” 1999.
[51] S. Bhattacharya, T.M. Frank, D.M. Divan, and B. Banerjee, “Parallel
active filter system implementation and design issues for utility interface
of adjustable speed drive systems,” in Proc. IAS Annual Meeting 1996,
San Diego, USA, 1996, vol. 2, pp. 1032–1039.
[52] Rolf Ottersten and Jan Svensson, “Vector current controlled grid con-
nected pwm converter - deadbeat control and saturation strategies,” in
Proc. 1998 IEEE Nordic Workshop on Power and Industrial Electronics,
Espoo, Finland, 1998, vol. 1, pp. 65–70.
[53] Finbarr Moynihan, “Fundamentals of dsp-based control for ac machines,”
Lecture notes for a tutorial held at PCIM’98, Nürnberg, Germany, 1998.
[54] Ahmet M. Hava, Russel J. Kerkman, and Thomas A. Lipo, “Simple
analytical and graphical tools for carrier based pwm methods,” in Proc.
PESC 1997, St. Louis, USA, 1997, vol. 2, pp. 1462–1471.
[55] Frede Blaabjerg, Ulrik Jaeger, Stig Munk-Nielsen, and John K. Pedersen,
“Power losses in pwm-vsi inverter using npt or pt igbt devices,” IEEE
Transactions on Power Electronics, vol. 10, no. 3, pp. 358–366, May 1995.
[56] Mariusz Malinowski and Marian P. Kaźmierkowski, “Dsp implementa-
tion of adaptive modulation for three-phase pwm rectifier/inverter,” in
Proc. IEEE Nordic Workshop on Power and Industrial Electronics 2000,
Aalborg, Denmark, 2000, vol. 1, pp. 288–292.
[57] F.R. Walsh, J.F. Moynihan, P.J. Roche, M.G. Egan, and J.M.D. Murphy,
“Analysis and influence of modulation scheme on the sizing of the input
filter in a pwm rectifier system,” in Proc. EPE 1999, Trondheim, Norway,
1997, vol. 2, pp. 929–933.

117
[58] Hyeoun-Dong Lee and Seung-Ki Sul, “A common-mode voltage reduction
in converter-inverter system by shifting active space vector in sampling
period,” in Proc. Power Conversion, May 1998, vol. PC, pp. 273–280.
[59] “Skiippack technical information,” Semikron Product Overview, 1998.
[60] H. Akagi, Y Kanazawa, and A Nabae, “Instantaneous reactive power
compensators comprising switching devices without energy storage com-
ponents,” IEEE Transactions on Industry Applications, vol. 20, no. 3,
pp. 625–630, May/June 1984.
[61] Robert D. Lorenz and Keith Van Patten, “High resolution velocity es-
timation for all digital, ac servo drives,” in Proc. IAS Annual Meeting
1988, Pittsburgh, USA, 1988, vol. 1, pp. 363–368.
[62] Jorma Luomi, “Transient phenomena in electrical machines,” Lecture
notes for a course in electrical drives held at Chalmers University of
Technology, Göteborg, Sweden, 1997.
[63] Ed. Bimal K. Bose, Power Electronics and Variable Frequency Drives,
IEEE Press, New York, 1997.

118
Appendix A

Glossary of Symbols, Subscripts,


Superscripts and Abbreviations

Symbols
Boldface characters denote space vectors or matrices. F and G are transfer
functions. All others are scalars. Space vectors in flux orientated reference
frames are given by f = fd + jfq , where d is the flux orientated axis (reactive
axis) and q is the torque axis (active axis).

A System matrix
B Input matrix
C Output matrix
C Capacitance F
eg , eg Grid voltage and grid voltage space vector, respectively V
F Controller
G Transfer function
i, i Current and current space vector, respectively A
I Identity matrix √
j Imaginary operator, −1
J Inertia kgm2
k Constant
kp Proportional gain
L Inductance H
p Derivative operator
p Pole pair
P Instantaneous active power W
Q Instantaneous reactive power VAr
R Resistance Ω
s Slip
S Instantaneous apparent power VA
t Time s

119
T Sampling period s
Te Instantaneous electro-mechanical torque Nm
Ti Controller integration time
u System input
u, u Voltage and voltage space vector, respectively V
Z Impedance Ω
ω Electrical angular velocity of the rotor rad/s
ωg , ωgf Angular frequency of the grid rad/s
ωk Angular frequency of an arbitrary variable rad/s
ωrf Angular frequency of the rotor flux rad/s
ωs Angular stator frequency in steady state rad/s
ωsf Angular frequency of the stator flux rad/s
ωsl Angular slip frequency rad/s
α Closed-loop bandwidth rad/s
λ Pole rad
ψ, ψ Flux linkage and flux linkage space vector, respectively Vs
σ Total leakage factor
τ Time constant s
θ Flux vector angle rad
|| Absolute value
b Estimated
e Error

Subscripts

br Braking
c Cut-off frequency
C Current model
d Real part (alligned with flux linkage) in a flux orientated
reference frame
dc Direct current
dec Decoupled
e Electro-mechanical
fb Feedback
g Grid
gf Grid flux
in Input (filter)
lp Low pass filter
m Mutual, T-model
M Mutual, inverse Γ-model
n, nom Nominal
p Pull-out
120
q Imaginary part in a flux orientated reference frame
r Rotor, T-model
R Rotor, inverse Γ-model
rf Rotor flux
s Stator
sc Stator current
sf Stator flux
sf Stator flux
trans Transient
α Real part in the stator orientated reference frame
β Imaginary part in the stator orientated reference frame
σ Leakage
0 No-load

Superscripts

base Base value


gf Grid flux orientated reference frame
k Arbitrary reference frame
max Maximum
nom Nominal
r Rotor orientated reference frame
rf Rotor flux orientated reference frame
s Stator orientated reference frame
sf Stator flux orientated reference frame

Complex conjugate
?
Reference

Abbreviations

AFD Adjustable frequency drive


BJT Bipolar junction transistor
D/A Digital to analog
dc, DC Direct current
det Determinant
DSP Digital signal processor
emf Electromotive force
GTO Gate turn on thyristor
IGBT Insulated gate bipolar transistor
MOSFET
121
PI Proportional integration
PWM Pulse width modulation
RFI Radio frequency interference
V/f Volts-per-Hertz

122
Appendix B

Space Vectors

Some principles of space vectors are briefly described in this appendix. It is


assumed that the simplifications in Section 3.1 are valid. It is also assumed
that the three-phase variables, f1 , f2 , f3 , have a sum equal to zero, i.e., a
zero-sequence signal is not present.

B.1 Space Vector Notation


The three-phase variables, f1 , f2 , f3 , define a space vector according to

s 2 ³ j0 j2π/3 j4π/3
´
f = f1 e + f 2 e + f3 e (B.1)
3 µ ¶
2 1
Re{f s } = fαs = f1 − (f2 + f3 ) = [symmetry] = f1 (B.2)
3 2
1 1
Im{f s } = fβs = √ (f2 − f3 ) = [symmetry] = √ (f1 + 2f2 ) (B.3)
3 3
Note that the gain 2/3 in (B.1) implies an amplitude transformation, i.e.,
the amplitude of the space vector is the same as the amplitude of the three-
phase variables. Fig. B.1 illustrates the space vector and its construction. The
inverse transformation from the stator-fixed complex plane to the three-phase

Im β 3f / 2

f3 f
e j2 π / 3
f2

α (fixed)
f1 Re

e j4 π / 3

Figure B.1: A space vector in the stator-fixed complex plane.

123
system becomes

f1 = fαs (B.4)

1 s 3 s
f2 = − fα + fβ (B.5)
2 √2
1 s 3 s
f3 = − fα − f (B.6)
2 2 β
The stator-fixed complex plane is referred to as the stator orientated reference
frame in this thesis.

B.2 Coordinate Transformations


The coordinate system transformation of variables in the stator orientated
reference frame to a rotating reference frame is given by

f k = f s e−jθk (B.7)

The real and imaginary components in the rotating reference frame become

Re{f k } = fdk = fαs cos(θk ) + fβs sin(θk ) (B.8)


Im{f k } = fqk = −fαs sin(θk ) + fβs cos(θk ) (B.9)

Fig B.2 shows the coordinate transformation and the projections of the space
vector on the axes. The coordinate transformation can also be reversed, hence

f s = f k ejθk (B.10)
The real and imaginary components in the stator orientated reference frame
become

Re{f s } = fαs = fdk cos(θk ) − fqk sin(θk ) (B.11)


Im{f s } = fβs = fdk sin(θk ) + fqk cos(θk ) (B.12)

q β

fβs f
ωk
d
f qk
f dk θk
α (fixed)
f αs

Figure B.2: Coordinate transformation between stator-fixed complex plane and rotating
complex plane.

124
The derivative operator p is replaced with p + jωk when transforming ro-
tor/stator voltage equations from the stator orientated reference frame to
a rotating reference frame [19]. The derivative operator p is replaced with
p − jωk when transforming rotor/stator voltage equations from a rotating
reference frame to the stator orientated reference frame.

B.3 Torque in Terms of Space Vectors


The instantaneous power in a symmetrical three-phase system is
3¡ s s ¢ 3 3
P = u1 i1 + u2 i2 + u3 i3 = uα iα + usβ isβ = Re {us i∗s } = Re {u∗s is } (B.13)
2 2 2
where the last three right-hand statements assume amplitude-invariant trans-
formation. The last two right-hand statements in (B.13) are valid in any ref-
erence frame and the superscripts are therefore dropped. The stator voltage
equation in the rotor orientated reference frame becomes
dψrs
urs = Rs irs
+ + jω ψrs (B.14)
dt
Multiplying (B.14) with 3ir∗ s /2 and taking the real part gives
½µ r
¶ ¾
3 3 d ψ s
Re {urs ir∗
s } = Re Rs irs + + jω ψrs ir∗
s (B.15)
2 2 dt
The terms in (B.15) represent the following power balance
Pelec = Ploss + Pf ield + Pmech (B.16)
The mechanical power is hence given by
3
Pmech = Re {jω ψrs ir∗
s } (B.17)
2
According to the mechanical subsystem, the mechanical power is also given
by
ω
Pmech = Te (B.18)
p
The instantaneous torque becomes [62]
3 3 3
Te = pRe {jψs i∗s } = − pIm {ψs i∗s } = pIm {ψ∗s is } (B.19)
2 2 2
Superscripts are dropped since (B.19) is valid in all reference frames. Eq. (B.19)
in the stator flux orientated reference frame can be simplified to
3 sf sf
Te = pψsd isq (B.20)
2
125
The instantaneous torque can be written in several ways. For instance, a
derivation similar to the above can be carried out with the rotor voltage
equation of the induction machine in the stator orientated reference frame.
The instantaneous torque then becomes [62]
3 3
Te = − pIm {ψ∗r ir } = − pIm {ψ∗R iR } (B.21)
2 2
Eq. (B.21) in the rotor flux orientated reference frame can be simplified to
3 rf rf 3 rf rf
Te = − pψrd irq = − pψRd iRq (B.22)
2 2
which can rewritten as
3 Lm rf rf 3 rf rf
Te = p ψrd isq = pψRd isq (B.23)
2 Lr 2

126
Appendix C

Base Values

For a machine with the nominal voltage Un , the nominal current In and the
nominal electric angular frequency ωn , the base values in the three-phase
system, stator orientated reference frame and in a flux orientated reference
frame can be defined as shown in Table C.1 [62].

Table C.1: Base Values.


Base value Three-phase system, Flux orientated
stator orientated reference frame reference frame

Base voltage Un 2Unph

Base current In 2In

Base angular frequency ωn ωn



Unph 2Unph
Base flux linkage
ωn ωn

Base power
(active, reative 3Unph In 3Unph In
and imaginary)

3Unph In 3Unph In
Base torque
ωn /p ωn /p

Unph Unph
Base impedance
In In

1 1
Base time
ωn ωn

127
In addition to Table C.1, dividing the equation of motion
J dω
= Te − T l (C.1)
p dt
where ω is the electrical rotor speed in rad/s, with the base torque and ex-
tending the left hand side of (C.1) with ωn2 /ωn2 gives
µ ¶
ω
d
J ωn2 ωn Te − T l
= (C.2)
p 3Unph In d(ωn t) 3Unph In
ωn /p ωn /p
i.e., the per-unit equation
dωpu
TJ = Te,pu − Tl,pu (C.3)
dtpu
where the normalized moment of inertia is [62]
µ ¶2
ωn
2 J
J ωn p
TJ = = ωn (C.4)
p 3Unph In 3Unph In
ωn /p
Finally, the relation between the angular velocity ω and the rotor position θ

ω= (C.5)
dt
is divided by the base angular velocity, giving [62]
ω dθ
= (C.6)
ωn d(ωn t)
i.e., the per-unit equation

ωpu = (C.7)
dtpu

128
Appendix D

Parameter Sensitivity of Stator Flux


Observers

The parameter sensitivity of a stator flux observer can be investigated by


studying the relative magnitude error |ψ e s | and the phase shift θesf of the
estimated stator flux compared to actual stator flux. For perfect stator flux
orientation, the relative magnitude error is equal to one and the phase shift
is equal to zero.

D.1 Open-Loop Observers


D.1.1 Voltage Model
The relative magnitude error and phase shift of the voltage model in steady
state become
µ ¶Ã b
!
b
ψs 1 1 + jτr ωsl Rs − Rs
= 1−j (D.1)
ψs Ls 1 + jστr ωsl ωs
b s|

e s| =
|ψ (D.2)
|ψ s |
½ ¾
bs
ψ
θesf = arg (D.3)
ψs

129
D.2 Closed-Loop Observers
D.2.1 Current and Voltage Model
The relative magnitude error and the phase shift of the current and voltage
model are [46]
K2
K = K1 + (D.4)
p
b sC
ψ bs 1 + jτr ωsl ¶ µ 1 + jb
L
µ
σ τbr ωsl

= (D.5)
ψsC Ls 1 + jb
τr ωsl 1 + jστr ωsl
µ ¶Ã b
!
b
ψsV 1 1 + jτr ωsl Rs − Rs
= 1−j (D.6)
ψsV Ls 1 + jστr ωsl ωs
ψb sC b sV
ψ
bs K + jω s
ψ ψsC ψsV
= (D.7)
ψs K + jωs
s
b ss |

e
|ψs | = (D.8)
|ψss |
θesf = arg{ψ b ss } − arg{ψss } (D.9)

D.2.2 Luenberger Observer


The state space system of the Luenberger with nearly constant poles, pre-
sented in Section 4.2.2, is
h iT
x bs b
b = is ψ ss
u = uss (D.10)
db
x b + K(is − bis )
b x + Bu
= Ab s s (D.11)
dt
bis = C
bxb (D.12)
s

The actual state space system of the induction machine is


h iT
s s
x = is ψ s u = uss (D.13)
dx
= Ax + Bu (D.14)
dt
iss = Cx (D.15)
The observed states and the actual states become
³ ´
b = (pI − A
x b + K C)
b −1 b + KC(pI − A) B u
B −1
(D.16)
x = (pI − A)−1 Bu (D.17)

130
e ss | and the phase shift θesf are given by the
The relative magnitude error |ψ
following
h i
C sf = 0 1 (D.18)
b ss
ψ C sf x b
= (D.19)
ψss C sf x
s
b ss |

e
|ψs | = (D.20)
|ψss |
θesf = θbsf − θsf = arg{ψ
b ss } − arg{ψss } (D.21)
The observer errors in steady state are obtained by substituting the derivative
operator p with jωs . The expressions for the relative magnitude error and
the phase are rather complicated and are conveniently evaluated by using
mathematical software with arithmetic capacity.

131
132
Appendix E

Induction Machine Data

The induction machine that is used in the theoretical and simulation parts of
this thesis is a Siemens 1PP8 317 - 4ZZ90-Z with the following data:
Nominal Values

Voltage (Y) 690 V

Current (Y) 285 A

Frequency (electrical) 50 Hz

Power 300 kW

Power factor 0.89

Rotor speed (mechanical) 1518 rpm

Torque 1970 Nm

Machine parameters at 50 Hz

Rotor resistance Rr 19.6 mΩ ¿ 0.014 p.u.

Stator resistance Rs 16.8 mΩ ¿ 0.012 p.u.

Rotor leakage inductance Lrσ 0.76 mH ¿ 0.17 p.u.

Stator leakage inductance Lsσ 0.40 mH ¿ 0.091 p.u.

Mutual inductance Lm 14.8 mH ¿ 3.33 p.u.

Core loss resistance Rc 255.1 Ω ¿ 182.5 p.u.

133
134
Appendix F

Publications

135

You might also like