07.03.2014 Views

Space Vector Modulated – Direct Torque Controlled (DTC – SVM ...

Space Vector Modulated – Direct Torque Controlled (DTC – SVM ...

Space Vector Modulated – Direct Torque Controlled (DTC – SVM ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Warsaw University of Technology<br />

Faculty of Electrical Engineering<br />

Institute of Control and Industrial Electronics<br />

Ph.D. Thesis<br />

Marcin Żelechowski, M. Sc.<br />

<strong>Space</strong> <strong>Vector</strong> <strong>Modulated</strong> <strong>–</strong> <strong>Direct</strong><br />

<strong>Torque</strong> <strong>Controlled</strong> (<strong>DTC</strong> <strong>–</strong> <strong>SVM</strong>)<br />

Inverter <strong>–</strong> Fed Induction Motor Drive<br />

Thesis supervisor<br />

Prof. Dr Sc. Marian P. Kaźmierkowski<br />

Warsaw <strong>–</strong> Poland, 2005


Acknowledgements<br />

The work presented in the thesis was carried out during author’s Ph.D. studies at the<br />

Institute of Control and Industrial Electronics in Warsaw University of Technology,<br />

Faculty of Electrical Engineering. Some parts of the work were realized in cooperation<br />

with foreign Universities:<br />

• University of Nevada, Reno, USA (US National Science Foundation grant <strong>–</strong><br />

Prof. Andrzej Trzynadlowski),<br />

• University of Aalborg, Denmark (Prof. Frede Blaabjerg),<br />

and company:<br />

• Power Electronics Manufacture <strong>–</strong> „TWERD”, Toruń, Poland.<br />

First of all, I would like to express gratitude Prof. Marian P. Kaźmierkowski for the<br />

continuous support and help during work of the thesis. His precious advice and<br />

numerous discussions enhanced my knowledge and scientific inspiration.<br />

I am grateful to Prof. Andrzej Sikorski from the Białystok Technical University and<br />

Prof. Włodzimierz Koczara from the Warsaw University of Technology for their<br />

interest in this work and holding the post of referee.<br />

Specially, I am indebted to my friend Dr Paweł Grabowski for support and<br />

assistance.<br />

Furthermore, I thank my colleagues from the Intelligent Control Group in Power<br />

Electronics for their support and friendly atmosphere. Specially, to Dr Dariusz Sobczuk,<br />

Dr Mariusz Malinowski, Dr Mariusz Cichowlas, and Dariusz Świerczyńki M.Sc.<br />

Finally, I would like thank to my whole family, particularly my parents for their love<br />

and patience.


Contents<br />

1. Introduction 1<br />

Pages<br />

2. Voltage Source Inverter Fed Induction Motor Drive 6<br />

2.1. Introduction 6<br />

2.2. Mathematical Model of Induction Motor 6<br />

2.3. Voltage Source Inverter (VSI) 12<br />

2.4. Pulse Width Modulation (PWM) 17<br />

2.4.1. Introduction 17<br />

2.4.2. Carrier Based PWM 18<br />

2.4.3. <strong>Space</strong> <strong>Vector</strong> Modulation (<strong>SVM</strong>) 22<br />

2.4.4. Relation Between Carrier Based and <strong>Space</strong> <strong>Vector</strong> Modulation 28<br />

2.4.5. Overmodulation (OM) 31<br />

2.4.6. Random Modulation Techniques 35<br />

2.5. Summary 39<br />

3. <strong>Vector</strong> Control Methods of Induction Motor 40<br />

3.1. Introduction 40<br />

3.2. Field Oriented Control (FOC) 40<br />

3.3. Feedback Linearization Control (FLC) 45<br />

3.4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control (<strong>DTC</strong>) 49<br />

3.4.1. Basics of <strong>Direct</strong> Flux and <strong>Torque</strong> Control 49<br />

3.4.2. Classical <strong>Direct</strong> <strong>Torque</strong> Control (<strong>DTC</strong>) <strong>–</strong> Circular Flux Path 53<br />

3.4.3. <strong>Direct</strong> Self Control (DSC) <strong>–</strong> Hexagon Flux Path 61<br />

3.5. Summary 64<br />

4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control with <strong>Space</strong> <strong>Vector</strong> Modulation (<strong>DTC</strong>-<strong>SVM</strong>) 66<br />

4.1. Introduction 66<br />

4.2. Structures of <strong>DTC</strong>-<strong>SVM</strong> <strong>–</strong> Review 66<br />

4.2.1. <strong>DTC</strong>-<strong>SVM</strong> Scheme with Closed <strong>–</strong> Loop Flux Control 66<br />

4.2.2. <strong>DTC</strong>-<strong>SVM</strong> Scheme with Closed <strong>–</strong> Loop <strong>Torque</strong> Control 68<br />

4.2.3. <strong>DTC</strong>-<strong>SVM</strong> Scheme with Close <strong>–</strong> Loop <strong>Torque</strong> and Flux Control<br />

Operating in Polar Coordinates 69<br />

4.2.4. <strong>DTC</strong>-<strong>SVM</strong> Scheme with Close <strong>–</strong> Loop <strong>Torque</strong> and Flux Control<br />

in Stator Flux Coordinates 70<br />

4.2.5. Conclusions from Review of the <strong>DTC</strong>-<strong>SVM</strong> Structures 71<br />

4.3. Analysis and Controller Design for <strong>DTC</strong>-<strong>SVM</strong> Method with<br />

Close <strong>–</strong> Loop <strong>Torque</strong> and Flux Control in Stator Flux Coordinates 71<br />

4.3.1. <strong>Torque</strong> and Flux Controllers Design <strong>–</strong> Symmetry Criterion Method 75<br />

4.3.2. <strong>Torque</strong> and Flux Controllers Design <strong>–</strong> Root Locus Method 78<br />

4.3.3. Summary of Flux and <strong>Torque</strong> Controllers Design 88<br />

4.4. Speed Controller Design 94<br />

4.5. Summary 98


Contents<br />

5. Estimation in Induction Motor Drives 99<br />

5.1. Introduction 99<br />

5.2. Estimation of Inverter Output Voltage 100<br />

5.3. Stator Flux <strong>Vector</strong> Estimators 104<br />

5.4. <strong>Torque</strong> Estimation 110<br />

5.5. Rotor Speed Estimation 110<br />

5.6. Summary 112<br />

6. Configuration of the Developed IM Drive Based on <strong>DTC</strong>-<strong>SVM</strong> 113<br />

6.1. Introduction 113<br />

6.2. Block Scheme of Implemented Control System 113<br />

6.3. Laboratory Setup Based on DS1103 115<br />

6.4. Drive Based on TMS320LF2406 118<br />

7. Experimental Results 122<br />

7.1. Introduction 122<br />

7.2. Pulse Width Modulation 122<br />

7.3. Flux and <strong>Torque</strong> Controllers 125<br />

7.4. <strong>DTC</strong>-<strong>SVM</strong> Control System 129<br />

8. Summary and Conclusions 138<br />

References 141<br />

List of Symbols 151<br />

Appendices 156<br />

A.1. Derivation of Fourier Series Formula for Phase Voltage<br />

A.2. SABER Simulation Model<br />

A.3. Data and Parameters of Induction Motors<br />

A.4. Equipment<br />

A.5. dSPACE DS1103 PPC Board<br />

A.6. Processor TMS320FL2406


1. Introduction<br />

The Adjustable Speed Drives (ADS) are generally used in industry. In most drives<br />

AC motors are applied. The standard in those drives are Induction Motors (IM) and<br />

recently also Permanent Magnet Synchronous Motors (PMSM) are offered. Variable<br />

speed drives are widely used in application such as pumps, fans, elevators, electrical<br />

vehicles, heating, ventilation and air-conditioning (HVAC), robotics, wind generation<br />

systems, ship propulsion, etc. [16].<br />

Previously, DC machines were preferred for variable speed drives. However, DC<br />

motors have disadvantages of higher cost, higher rotor inertia and maintenance problem<br />

with commutators and brushes. In addition they cannot operate in dirty and explosive<br />

environments. The AC motors do not have the disadvantages of DC machines.<br />

Therefore, in last three decades the DC motors are progressively replaced by AC drives.<br />

The responsible for those result are development of modern semiconductor devices,<br />

especially power Insulated Gate Bipolar Transistor (IGBT) and Digital Signal Processor<br />

(DSP) technologies.<br />

The most economical IM speed control methods are realized by using frequency<br />

converters. Many different topologies of frequency converters are proposed and<br />

investigated in a literature. However, a converter consisting of a diode rectifier, a dclink<br />

and a Pulse Width <strong>Modulated</strong> (PWM) voltage inverter is the most applied used in<br />

industry (see section 2.3).<br />

The high-performance frequency controlled PWM inverter <strong>–</strong> fed IM drive should be<br />

characterized by:<br />

• fast flux and torque response,<br />

• available maximum output torque in wide range of speed operation region,<br />

• constant switching frequency,<br />

• uni-polar voltage PWM,<br />

• low flux and torque ripple,<br />

• robustness for parameter variation,<br />

• four-quadrant operation,


1. Introduction<br />

These features depend on the applied control strategy. The main goal of the chosen<br />

control method is to provide the best possible parameters of drive. Additionally, a very<br />

important requirement regarding control method is simplicity (simple algorithm, simple<br />

tuning and operation with small controller dimension leads to low price of final<br />

product).<br />

A general classification of the variable frequency IM control methods is presented in<br />

Fig. 1.1 [67]. These methods can be divided into two groups: scalar and vector.<br />

Variable<br />

Frequency Control<br />

Scalar based<br />

controllers<br />

<strong>Vector</strong> based<br />

controller<br />

U/f=const.<br />

Volt/Hertz<br />

is<br />

= f ( ωr<br />

)<br />

Stator Current<br />

Field Oriented<br />

Feedback<br />

Linearization<br />

<strong>Direct</strong> <strong>Torque</strong><br />

Control<br />

Passivity Based<br />

Control<br />

Rotor Flux<br />

Oriented<br />

Stator Flux<br />

Oriented<br />

<strong>Direct</strong> <strong>Torque</strong><br />

<strong>Space</strong> - <strong>Vector</strong><br />

Modulation<br />

Circle flux<br />

trajectory<br />

(Takahashi)<br />

Hexagon flux<br />

trajectory<br />

(Takahashi)<br />

<strong>Direct</strong><br />

(Blaschke)<br />

Indirect<br />

(Hasse)<br />

Open Loop<br />

NFO (Jonsson) & o&<br />

Closed Loop<br />

Flux & <strong>Torque</strong><br />

Control<br />

Fig. 1.1. General classification of induction motor control methods<br />

The scalar control methods are simple to implement. The most popular in industry is<br />

constant Voltage/Frequency (V/Hz=const.) control. This is the simplest, which does not<br />

provide a high-performance. The vector control group allows not only control of the<br />

voltage amplitude and frequency, like in the scalar control methods, but also the<br />

instantaneous position of the voltage, current and flux vectors. This improves<br />

significantly the dynamic behavior of the drive.<br />

However, induction motor has a nonlinear structure and a coupling exists in the<br />

motor, between flux and the produced electromagnetic torque. Therefore, several<br />

methods for decoupling torque and flux have been proposed. These algorithms are<br />

based on different ideas and analysis.<br />

2


1. Introduction<br />

The first vector control method of induction motor was Field Oriented Control<br />

(FOC) presented by K. Hasse (Indirect FOC) [45] and F. Blaschke (<strong>Direct</strong> FOC) [12] in<br />

early of 70s. Those methods were investigated and discussed by many researchers and<br />

have now become an industry standard. In this method the motor equations are<br />

transformed into a coordinate system that rotates in synchronism with the rotor flux<br />

vector. The FOC method guarantees flux and torque decoupling. However, the<br />

induction motor equations are still nonlinear fully decoupled only for constant flux<br />

operation.<br />

An other method known as Feedback Linearization Control (FLC) introduces a new<br />

nonlinear transformation of the IM state variables, so that in the new coordinates, the<br />

speed and rotor flux amplitude are decoupled by feedback [81, 83].<br />

A method based on the variation theory and energy shaping has been investigated<br />

recently, and is called Passivity Based Control (PBC) [88]. In this case the induction<br />

motor is described in terms of the Euler-Lagrange equations expressed in generalized<br />

coordinates.<br />

In the middle of 80s new strategies for the torque control of induction motor was<br />

presented by I. Takahashi and T. Noguchi as <strong>Direct</strong> <strong>Torque</strong> Control (<strong>DTC</strong>) [97] and by<br />

M. Depenbrock as <strong>Direct</strong> Self Control (DSC) [4, 31, 32]. Those methods thanks to the<br />

other approach to control of IM have become alternatives for the classical vector control<br />

<strong>–</strong> FOC. The authors of the new control strategies proposed to replace motor decoupling<br />

and linearization via coordinate transformation, like in FOC, by hysteresis controllers,<br />

which corresponds very well to on-off operation of the inverter semiconductor power<br />

devices. These methods are referred to as classical <strong>DTC</strong>. Since 1985 they have been<br />

continuously developed and improved by many researchers.<br />

Simple structure and very good dynamic behavior are main features of <strong>DTC</strong>.<br />

However, classical <strong>DTC</strong> has several disadvantages, from which most important is<br />

variable switching frequency.<br />

Recently, from the classical <strong>DTC</strong> methods a new control techniques called <strong>Direct</strong><br />

<strong>Torque</strong> Control <strong>–</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulated</strong> (<strong>DTC</strong>-<strong>SVM</strong>) has been developed.<br />

In this new method disadvantages of the classical <strong>DTC</strong> are eliminated. Basically, the<br />

<strong>DTC</strong>-<strong>SVM</strong> strategies are the methods, which operates with constant switching<br />

frequency. These methods are the main subject of this thesis. The <strong>DTC</strong>-<strong>SVM</strong> structures<br />

3


1. Introduction<br />

are based on the same fundamentals and analysis of the drive as classical <strong>DTC</strong>.<br />

However, from the formal considerations these methods can also be viewed as stator<br />

field oriented control (SFOC), as shown in Fig. 1.1.<br />

Presented <strong>DTC</strong>-<strong>SVM</strong> technique has also simple structure and provide dynamic<br />

behavior comparable with classical <strong>DTC</strong>. However, <strong>DTC</strong>-<strong>SVM</strong> method is characterized<br />

by much better parameters in steady state operation.<br />

Therefore, the following thesis can be formulated: “The most convenient industrial<br />

control scheme for voltage source inverter-fed induction motor drives is direct<br />

torque control with space vector modulation <strong>DTC</strong>-<strong>SVM</strong>”<br />

In order to prove the above thesis the author used an analytical and simulation based<br />

approach, as well as experimental verification on the laboratory setup with 5 kVA and<br />

18 kVA IGBT inverters with 3 kW and 15 kW induction motors, respectively.<br />

Moreover, the control algorithm <strong>DTC</strong>-<strong>SVM</strong> has been introduced used in a serial<br />

commercial product of Polish manufacture TWERD, Toruń.<br />

In the author’s opinion the following parts of the thesis are his original achievements:<br />

• elaboration and experimental verification of flux and torque controller design for<br />

<strong>DTC</strong>-<strong>SVM</strong> induction motor drives,<br />

• development of a SABER - based simulation algorithm for control and<br />

investigation voltage source inverter-fed induction motors,<br />

• construction and practical verification of the experimental setups with 5 kVA and<br />

18 kVA IGBT inverters,<br />

• bringing into production and testing of developed <strong>DTC</strong>-<strong>SVM</strong> algorithm in Polish<br />

industry.<br />

The thesis consist of eight chapters. Chapter 1 is an introduction. In Chapter 2<br />

mathematical model of IM, voltage source inverter construction and pulse width<br />

modulation techniques are presented. Chapter 3 describes basic vector control method<br />

of IM and gives analysis of advantages and disadvantages for all methods. In this<br />

chapter basic principles of direct torque control are also presented. Those basis are<br />

common for classical <strong>DTC</strong>, which is presented in Chapter 3 and for <strong>DTC</strong>-<strong>SVM</strong> method.<br />

Chapter 4 is devoted to analysis and synthesis of <strong>DTC</strong>-<strong>SVM</strong> control technique. The<br />

flux, torque and speed controllers design are presented. In Chapter 5 the estimations<br />

4


1. Introduction<br />

algorithms are described and discussed. In Chapter 6 implemented <strong>DTC</strong>-<strong>SVM</strong> control<br />

algorithm and used hardware setup are presented. In Chapter 7 experimental results are<br />

presented and studied. Chapter 8 includes a conclusion. Description of the simulation<br />

program and parameters of the equipment used are given in Appendixes.<br />

5


2. Voltage Source Inverter Fed Induction Motor Drive<br />

2.1. Introduction<br />

In this chapter the model of induction motor will be presented. This mathematical<br />

description is based on space vector notation. In next part description of the voltage<br />

source inverter is given. The inverter is controlled in Pulse Width Modulation fashion.<br />

In last part of this chapter review of the modulation technique is presented.<br />

2.2. Mathematical Model of Induction Motor<br />

When describing a three-phase IM by a system of equations [66] the following<br />

simplifying assumptions are made:<br />

• the three-phase motor is symmetrical,<br />

• only the fundamental harmonic is considered, while the higher harmonics of the<br />

spatial field distribution and of the magnetomotive force (MMF) in the air gap<br />

are disregarded,<br />

• the spatially distributed stator and rotor windings are replaced by a specially<br />

formed, so-called concentrated coil,<br />

• the effects of anisotropy, magnetic saturation, iron losses and eddy currents are<br />

neglected,<br />

• the coil resistances and reactance are taken to be constant,<br />

• in many cases, especially when considering steady state, the current and voltages<br />

are taken to be sinusoidal.<br />

Taking into consideration the above stated assumptions the following equations of<br />

the instantaneous stator phase voltage values can be written:<br />

U<br />

U<br />

dΨ<br />

A<br />

= I<br />

ARs<br />

(2.1a)<br />

dt<br />

A<br />

+<br />

dΨ<br />

B<br />

= I<br />

BRs<br />

(2.1b)<br />

dt<br />

B<br />

+


2.2. Mathematical Model of Induction Motor<br />

U<br />

dΨ<br />

C<br />

= I<br />

C<br />

Rs<br />

(2.1c)<br />

dt<br />

C<br />

+<br />

The space vector method is generally used to describe the model of the induction<br />

motor. The advantages of this method are as follows:<br />

• reduction of the number of dynamic equations,<br />

• possibility of analysis at any supply voltage waveform,<br />

• the equations can be represented in various rectangular coordinate systems.<br />

A three-phase symmetric system represented in a neutral coordinate system by phase<br />

quantities, such as: voltages, currents or flux linkages, can be replaced by one resulting<br />

space vector of, respectively, voltage, current and flux-linkage. A space vector is<br />

defined as:<br />

2<br />

k =<br />

3<br />

where: k ( t) k ( t) k ( t)<br />

A B<br />

,<br />

2<br />

[ 1⋅<br />

k () t + a ⋅ k () t + a ⋅ k () t ]<br />

A<br />

C<br />

B<br />

C<br />

(2.2)<br />

, <strong>–</strong> arbitrary phase quantities in a system of natural<br />

coordinates, satisfying the condition k ( t) + k ( t) + k ( t) = 0<br />

1, a, a 2 <strong>–</strong> complex unit vectors, with a phase shift<br />

2/3 <strong>–</strong> normalization factor.<br />

A B C<br />

,<br />

Im<br />

B<br />

a<br />

k<br />

3<br />

k<br />

2<br />

2<br />

a k ( t)<br />

C<br />

ak B<br />

k A<br />

(t)<br />

A<br />

(t)<br />

1<br />

Re<br />

2<br />

a<br />

C<br />

Fig. 2.1. Construction of space vector according to the definition (2.2)<br />

7


2. Voltage Source Inverter Fed Induction Motor Drive<br />

An example of the space vector construction is shown in Fig. 2.1.<br />

Using the space vector method the IM model equation can be written as:<br />

dΨs<br />

U<br />

s<br />

= IsR s<br />

+<br />

(2.3a)<br />

dt<br />

dΨr<br />

U<br />

r<br />

= IrR r<br />

+<br />

(2.3b)<br />

dt<br />

Ψ<br />

s<br />

jγ<br />

m<br />

= L s<br />

I + Me I<br />

(2.4a)<br />

s<br />

r<br />

Ψ<br />

r<br />

− jγ<br />

m<br />

= L r<br />

I + Me I<br />

(2.4b)<br />

r<br />

s<br />

These are the voltage equations (2.3) and flux-current equations (2.4).<br />

To obtain a complete set of electric motor equations it is necessary to, firstly,<br />

transform the equations (2.3, 2.4) into a common rotating coordinate system and<br />

secondly bring the rotor value into the stator side and thirdly. These equations are<br />

written in the coordinate system K rotating with the angular speed<br />

Ω<br />

K<br />

.<br />

U<br />

U<br />

sK<br />

rK<br />

dΨ<br />

dt<br />

sK<br />

= RsI<br />

sK<br />

+ + jΩKΨsK<br />

(2.5a)<br />

dΨ<br />

dt<br />

( ΩK<br />

− pbΩm<br />

) Ψ<br />

K<br />

rK<br />

= RrI<br />

rK<br />

+ + j<br />

r<br />

(2.5b)<br />

Ψ = L I + L I<br />

(2.6a)<br />

sK<br />

s sK<br />

M<br />

rK<br />

Ψ = L I + L I<br />

(2.6b)<br />

rK<br />

r rK<br />

M<br />

sK<br />

The equation of the dynamic rotor rotation can be expressed as:<br />

dΩ<br />

dt<br />

m<br />

[ M − M − BΩ ]<br />

= 1 e L m<br />

(2.7)<br />

J<br />

where:<br />

M<br />

e<br />

<strong>–</strong> electromagnetic torque,<br />

M<br />

L<br />

<strong>–</strong> load torque,<br />

B <strong>–</strong> viscous constant.<br />

In further consideration the friction factor will be negated ( B = 0)<br />

.<br />

The electromagnetic torque<br />

M<br />

e<br />

can be expressed by the following formulas:<br />

8


( )<br />

2.2. Mathematical Model of Induction Motor<br />

ms<br />

M Im I *<br />

e<br />

= − pb<br />

LM<br />

s<br />

I r<br />

(2.8)<br />

2<br />

( )<br />

ms<br />

M Im Ψ *<br />

e<br />

= pb<br />

s<br />

I s<br />

(2.9)<br />

2<br />

Taking into consideration the fact that in the cage motor the rotor voltage equals zero<br />

and the electromagnetic torque equation (2.9) a complete set of equations for the cage<br />

induction motor can be written as:<br />

U<br />

sK<br />

dΨ<br />

dt<br />

sK<br />

= RsIsK<br />

+ + jΩKΨs<br />

K<br />

(2.10a)<br />

dΨr<br />

K<br />

0 = RrI<br />

rK<br />

+ + j( ΩK<br />

− pbΩm<br />

) Ψr<br />

K<br />

(2.10b)<br />

dt<br />

Ψ = L I + L I<br />

(2.11a)<br />

sK<br />

s sK<br />

M<br />

rK<br />

Ψ = L I + L I<br />

(2.11b)<br />

rK<br />

r rK<br />

M<br />

sK<br />

dΩ<br />

dt<br />

m<br />

1 ⎡<br />

⎢ p<br />

J ⎣<br />

m<br />

Im Ψ *<br />

2<br />

⎤<br />

( ) − M ⎥ ⎦<br />

s<br />

=<br />

b<br />

s<br />

I s L<br />

(2.12)<br />

Equations (2.10), (2.11) and (2.12) are the basis of further consideration.<br />

The applied space vector method as a mathematical tool for the analysis of the<br />

electric machines a complete set equations can be represented in various systems of<br />

coordinates. One of them is the stationary coordinates system (fixed to the stator) α − β<br />

in this case angular speed of the reference frame is zero Ω = 0 . The complex space<br />

vector can be resolved into components α and β .<br />

U = U + jU<br />

(2.13a)<br />

sK<br />

sα sβ<br />

K<br />

I = I + jI<br />

, I<br />

rK<br />

= Irα + jIrβ<br />

(2.13b)<br />

sK<br />

sα sβ<br />

Ψ = Ψ + jΨ<br />

, Ψ<br />

rK<br />

= Ψ<br />

rβ<br />

+ jΨ<br />

rβ<br />

(2.13c)<br />

sK<br />

sα sβ<br />

In α − β coordinate system the motor model equation can be written as:<br />

U<br />

sα = RsI<br />

sα<br />

+<br />

dΨ<br />

s<br />

dt<br />

α<br />

(2.14a)<br />

9


2. Voltage Source Inverter Fed Induction Motor Drive<br />

U<br />

dΨ<br />

sβ<br />

= RsI<br />

sβ<br />

(2.14b)<br />

dt<br />

sβ +<br />

dΨ<br />

rα<br />

0 = R<br />

rIrα<br />

+ + pbΩmΨ<br />

rβ<br />

(2.14c)<br />

dt<br />

dΨ<br />

rβ<br />

0 = Rr<br />

Ir<br />

β<br />

+ − pbΩmΨ<br />

rα<br />

(2.14d)<br />

sα<br />

s sα<br />

dt<br />

Ψ = L I + L I<br />

(2.15a)<br />

sβ<br />

s sβ<br />

M<br />

M<br />

rα<br />

Ψ = L I + L I<br />

(2.15b)<br />

rβ<br />

Ψ = L I + L I<br />

(2.15c)<br />

rα<br />

rβ<br />

r rα<br />

r rβ<br />

M<br />

M<br />

sα<br />

Ψ = L I + L I<br />

(2.15d)<br />

sβ<br />

dΩ<br />

dt<br />

m<br />

1 ⎡<br />

⎢<br />

p<br />

J ⎣<br />

m<br />

2<br />

⎤<br />

( Ψ I −Ψ<br />

I ) − M<br />

⎥ ⎦<br />

s<br />

=<br />

b sα<br />

sβ<br />

sβ<br />

sα<br />

L<br />

(2.16)<br />

The relations described above by the motor equations can be represented as a block<br />

diagram. There is not just one block diagram of an induction motor. The lay-out<br />

Construction of a block diagram will depend on the chosen coordinate system and input<br />

signals. For instance, if it is assumed in the stationary α − β coordinate system that the<br />

input signal to the motor is the stator voltage, the equations (2.14-2.16) can be<br />

transformed into:<br />

dΨ<br />

dt<br />

dΨ<br />

dt<br />

dΨ<br />

dt<br />

dΨ<br />

I<br />

I<br />

dt<br />

sα<br />

sβ<br />

sα<br />

sβ<br />

rα<br />

rβ<br />

= U − R I<br />

(2.17a)<br />

sα<br />

sβ<br />

s sα<br />

= U − R I<br />

(2.17b)<br />

r rα<br />

s sβ<br />

= −R I − p Ω Ψ<br />

(2.17c)<br />

r rβ<br />

b<br />

b<br />

m<br />

m<br />

rβ<br />

= −R I + p Ω Ψ<br />

(2.17d)<br />

rα<br />

LM<br />

= 1 Ψ<br />

sα<br />

− Ψ<br />

rα<br />

(2.18a)<br />

σL<br />

σL<br />

L<br />

s<br />

s<br />

r<br />

LM<br />

= 1 Ψ<br />

sβ<br />

− Ψ<br />

rβ<br />

(2.18b)<br />

σL<br />

σL<br />

L<br />

r<br />

s<br />

r<br />

10


2.2. Mathematical Model of Induction Motor<br />

I<br />

I<br />

rα<br />

rβ<br />

LM<br />

= 1 Ψ<br />

rα<br />

− Ψ<br />

sα<br />

(2.18c)<br />

σL<br />

σL<br />

L<br />

r<br />

s<br />

r<br />

LM<br />

= 1 Ψ<br />

rβ<br />

− Ψ<br />

sβ<br />

(2.18d)<br />

σL<br />

σL<br />

L<br />

r<br />

s<br />

r<br />

dΩ<br />

dt<br />

m<br />

1 ⎡<br />

⎢<br />

p<br />

J ⎣<br />

m<br />

2<br />

⎤<br />

( Ψ I −Ψ<br />

I ) − M<br />

⎥ ⎦<br />

s<br />

=<br />

b sα<br />

sβ<br />

sβ<br />

sα<br />

L<br />

(2.19)<br />

These equations can be represented in the block diagram as shown in Fig. 2.2.<br />

M L<br />

R<br />

s<br />

U sα<br />

∫<br />

Ψ sα<br />

1<br />

σL s<br />

I sα<br />

LM<br />

σL L<br />

s<br />

r<br />

LM<br />

σL<br />

L<br />

s<br />

r<br />

ms<br />

pb<br />

2<br />

M e<br />

1<br />

J<br />

∫<br />

Ω m<br />

R r<br />

∫<br />

I rα<br />

Ψ rα<br />

1<br />

σL r<br />

p b<br />

∫<br />

R r<br />

Ψ rβ<br />

I rβ<br />

1<br />

σL r<br />

LM<br />

σL L<br />

s<br />

r<br />

LM<br />

σL<br />

L<br />

s<br />

r<br />

U sβ<br />

∫<br />

Ψ sβ<br />

1<br />

σL s<br />

I sβ<br />

R s<br />

Fig. 2.2. Block diagram of an induction motor in the stationary coordinate system<br />

α − β<br />

This representation of the induction motor is not good for use to design a control<br />

structure, because the output signals flux, torque and speed depend on both inputs. From<br />

the control point of view this system is complicated. That is the reason why there are a<br />

11


2. Voltage Source Inverter Fed Induction Motor Drive<br />

few methods proposed to decouple the flux and torque control. It is achieved, for<br />

example, by the orientation of the coordinate system to the rotor or stator flux vectors.<br />

Both control systems are described further in Chapter 3.<br />

The equations (2.17), (2.18), (2.19) and the block diagram presented in the Fig. 2.2<br />

can be used to build a simulation model of the induction motor. It was used in a<br />

simulation model, which is presented in Appendix A.2.<br />

2.3. Voltage Source Inverter (VSI)<br />

The three-phase two level VSI consists of six active switches. The basic topology of<br />

the inverter is shown in Fig. 2.3. The converter consists of the three legs with IGBT<br />

transistors, or (in the case of high power) GTO thyristors and free-wheeling diodes. The<br />

inverter is supplied by a voltage source composed of a diode rectifier with a C filter in<br />

the dc-link. The capacitor C is typically large enough to obtain adequately low voltage<br />

source impedance for the alternating current component in the dc-link.<br />

DC side<br />

PWM Converter<br />

U dc<br />

2<br />

U dc<br />

2<br />

0<br />

C<br />

C<br />

S A<br />

+<br />

S A<br />

-<br />

T 1<br />

T 3<br />

T 5<br />

D S B<br />

+<br />

S C<br />

+<br />

1<br />

T 2<br />

D 3<br />

D 5<br />

T 4<br />

T 6<br />

D S<br />

2<br />

B<br />

- D S<br />

4 C<br />

-<br />

D 6<br />

I A<br />

I<br />

A<br />

U B<br />

I C<br />

AB B C<br />

AC side<br />

R A<br />

R B<br />

R C<br />

L A<br />

U A<br />

L B<br />

U B<br />

L C<br />

U C<br />

E A<br />

E B<br />

E C<br />

N<br />

IM<br />

Fig. 2.3. Topology of the voltage source inverter<br />

12


2.3. Voltage Source Inverter (VSI)<br />

The voltage source inverter (Fig. 2.3) makes it possible to connect each of the three<br />

motor phase coils to a positive or negative voltage of the dc link. Fig. 2.4 explains the<br />

fabrication of the output voltage waves in square-wave, or six-step, mode of operation.<br />

The phase voltages are related to the dc-link center point 0 (see Fig. 2.3).<br />

a)<br />

U A0<br />

1<br />

U dc<br />

2<br />

1 2 3 4 5 6<br />

0<br />

1<br />

− U dc<br />

2<br />

π<br />

2π<br />

ωt<br />

b)<br />

U B0<br />

1<br />

U dc<br />

2<br />

0<br />

1<br />

− U dc<br />

2<br />

π<br />

2π<br />

ωt<br />

c)<br />

U C0<br />

1<br />

U dc<br />

2<br />

0<br />

1<br />

− U dc<br />

2<br />

π<br />

2π<br />

ωt<br />

d)<br />

U AB<br />

U dc<br />

2<br />

U dc<br />

3<br />

1<br />

U dc<br />

3<br />

0<br />

1<br />

− U dc<br />

3<br />

2<br />

− U dc<br />

3<br />

π<br />

2π<br />

ωt<br />

− U dc<br />

e)<br />

U A<br />

2<br />

U dc<br />

3<br />

1<br />

U dc<br />

3<br />

0<br />

1<br />

− U dc<br />

3<br />

2<br />

− U dc<br />

3<br />

π<br />

2π<br />

ωt<br />

Fig. 2.4. The output voltage waveforms in six-step mode<br />

The phase voltage of an inverter fed motor (Fig. 2.4e) can be expressed by Fourier<br />

series as [16, 66]:<br />

U<br />

A<br />

2<br />

= U<br />

π<br />

∞<br />

∑<br />

dc<br />

n=<br />

1<br />

1<br />

sin<br />

n<br />

( nωt) = U ( ) sin( nωt)<br />

∞<br />

∑<br />

m n<br />

n=<br />

1<br />

(2.20)<br />

where:<br />

U<br />

dc<br />

- dc supply voltage,<br />

13


2. Voltage Source Inverter Fed Induction Motor Drive<br />

U<br />

m<br />

2<br />

= U - peak value of the n-th harmonic,<br />

nπ<br />

( n) dc<br />

n = 1+6k, k = 0, ±1, ±2,…<br />

Derivation of the formula (2.20) is presented in Appendix A.1.<br />

a) U 1<br />

(100)<br />

b) U 2<br />

(110)<br />

U dc<br />

U dc<br />

A B C<br />

A B C<br />

c) U 3<br />

(010)<br />

d) U 4<br />

(011)<br />

U dc<br />

U dc<br />

A B C<br />

A B C<br />

e) U 5<br />

(001)<br />

f) U 6<br />

(101)<br />

U dc<br />

U dc<br />

A B C<br />

A B C<br />

g) U 0<br />

(000)<br />

h) U 7<br />

(111)<br />

U dc<br />

U dc<br />

A B C<br />

A B C<br />

Fig. 2.5. Switching states for the voltage source inverter<br />

From the equation (2.20) the fundamental peak value is given as:<br />

2<br />

U m () 1<br />

= U<br />

dc<br />

(2.21)<br />

π<br />

14


2.3. Voltage Source Inverter (VSI)<br />

This value will be used to define the modulation index M used in pulse width<br />

modulation (PWM) methods (see section 2.4).<br />

For the sake of the inverter structure, each inverter-leg can be represented as an ideal<br />

switch. The equivalent inverter states are shown in Fig. 2.5.<br />

There are eight possible positions of the switches in the inverter. These states<br />

correspond to voltage vectors. Six of them (Fig. 2.5 a-f) are active vectors and the last<br />

two (Fig. 2.5 g-h) are zero vectors. The output voltage represented by space vectors is<br />

defined as:<br />

2 3<br />

⎧ j(<br />

v−1)<br />

π<br />

⎪<br />

U<br />

dce<br />

v = 1...6<br />

U = 3<br />

v ⎨<br />

(2.22)<br />

⎪0<br />

v = 0,7<br />

⎩<br />

The output voltage vectors are shown in Fig. 2.6.<br />

Im<br />

U 3<br />

(010)<br />

U 2<br />

(110)<br />

U 4<br />

(011)<br />

U 0<br />

(000)<br />

U 1<br />

(100)<br />

U 7<br />

(111)<br />

Re<br />

U 5<br />

(001) U 6<br />

(101)<br />

Fig. 2.6. Output voltage represented as space vectors<br />

Any output voltage can in average be generated, of course limited by the value of the<br />

dc voltage. In order to realize many different pulse width modulation methods are<br />

proposed [13, 27, 30, 38, 46, 47, 51, 52, 105] in history. However, the general idea is<br />

15


2. Voltage Source Inverter Fed Induction Motor Drive<br />

based on a sequential switching of active and zero vectors. The modulation methods are<br />

widely described in the next section.<br />

Only one switch in an inverter-leg (Fig. 2.3) can be turned on at a time, to avoid a<br />

short circuit in the dc-link. A delay time in the transistor switching signals must be<br />

inserted. During this delay time, the dead-time T D transistors cease to conduct. Two<br />

control signals S A +, S A - for transistors T 1 , T 2 with dead-time T D are presented in Fig.<br />

2.7. The duration of dead-time depends of the used transistor. Most of them it takes 1-<br />

3µs.<br />

S A<br />

+<br />

S A<br />

-<br />

t<br />

T D<br />

T D<br />

t<br />

T s<br />

Fig. 2.7. Dead-time effect in a PWM inverter<br />

Although, this delay time guarantees safe operation of the inverter, it causes a serious<br />

distortion in the output voltage. It results in a momentary loss of control, where the<br />

output voltage deviates from the reference voltage. Since this is repeated for every<br />

switching operation, it has significant influence on the control of the inverter. This is<br />

known as the dead-time effect. This is important in applications like a sensorless direct<br />

torque control of induction motor. These applications require feedback signals like:<br />

stator flux, torque and mechanical speed. Typically the inverter output voltage is needed<br />

to calculate it. Unfortunately, the output voltage is very difficult to measure and it<br />

requires additional hardware. Because of that for calculation of feedback signals the<br />

reference voltage is used. However, the relation between the output voltage and the<br />

reference voltage is nonlinear due to the dead-time effect [8]. It is especially important<br />

16


2.4. Pulse Width Modulation (PWM)<br />

for the low speed range when voltage is very low. The dead-time may also cause<br />

instability in the induction motor [52].<br />

Therefore, for correct operation of control algorithm proper compensation of deadtime<br />

is required. Many approaches are proposed to compensate of this effect [2, 3, 8, 29,<br />

54, 64, 76].<br />

The dead-time compensation is directly connected with estimation of inverter output<br />

voltage. Therefore, compensation algorithm, which is used in final control structure of<br />

the inverter is presented in Chapter 5.<br />

2.4. Pulse Width Modulation (PWM)<br />

2.4.1. Introduction<br />

In the voltage source inverter conversion of dc power to three-phase ac power is<br />

performed in the switched mode (Fig. 2.3). This mode consists in power semiconductors<br />

switches are controlled in an on-off fashion. The actual power flow in each motor phase<br />

is controlled by the duty cycle of the respective switches. To obtain a suitable duty<br />

cycle for each switches technique pulse width modulation is used. Many different<br />

modulation methods were proposed and development of it is still in progress [13, 27,<br />

30, 38, 46, 47, 51, 52, 105].<br />

The modulation method is an important part of the control structure. It should<br />

provide features like:<br />

• wide range of linear operation,<br />

• low content of higher harmonics in voltage and current,<br />

• low frequency harmonics,<br />

• operation in overmodulation,<br />

• reduction of common mode voltage,<br />

• minimal number of switching to decrease switching losses in the power<br />

components.<br />

The development of modulation methods may improve converter parameters. In the<br />

carrier based PWM methods the Zero Sequence Signals (ZSS) [46] are added to extend<br />

17


2. Voltage Source Inverter Fed Induction Motor Drive<br />

the linear operation range (see section 2.4.2). The carrier based modulation methods<br />

with ZSS correspond to space vector modulation. It will be widely presented in section<br />

2.4.4.<br />

All PWM methods have specific features. However, there is not just one PWM<br />

method which satisfies all requirements in the whole operating region. Therefore, in the<br />

literature are proposed modulators, which contain from several modulation methods.<br />

For example, adaptive space vector modulation [79], which provides the following<br />

features:<br />

• full control range including overmodulation and six-step mode, achieved by the<br />

use of three different modulation algorithms,<br />

• reduction of switching losses thanks to an instantaneous tracking peak value of<br />

the phase current.<br />

The content of the higher harmonics voltage (current) and electromagnetic<br />

interference generated in the inverter fed drive depends on the modulation technique.<br />

Therefore, PWM methods are investigated from this point of view. To reduce these<br />

disadvantages several methods have been proposed. One of these methods is random<br />

modulation (RPWM). The classical carrier based method or space vector modulation<br />

method are named deterministic (DEPWM), because these methods work with constant<br />

switching frequency. In opposite to the deterministic methods, the random modulation<br />

methods work with variable frequency, or with randomly changed switching sequence<br />

(see section 2.4.6).<br />

2.4.2. Carrier Based PWM<br />

The most widely used method of pulse width modulation are carrier based. This<br />

method is also known as the sinusoidal (SPWM), triangulation, subharmonic, or<br />

suboscillation method [16, 52]. Sinusoidal modulation is based on triangular carrier<br />

signal as shown in Fig. 2.8. In this method three reference signals U Ac , U Bc , U Cc are<br />

compared with triangular carrier signal U t , which is common to all three phases. In this<br />

way the logical signals S A , S B , S C are generated, which define the switching instants of<br />

the power transistors as is shown in Fig. 2.9.<br />

18


2.4. Pulse Width Modulation (PWM)<br />

U dc<br />

U Ac<br />

S A<br />

U Bc<br />

S B<br />

U Cc<br />

U t<br />

S C<br />

A B C<br />

Carrier<br />

N<br />

Fig. 2.8. Block scheme of carrier based sinusoidal PWM<br />

U dc<br />

2<br />

U t<br />

U Ac<br />

U Bc<br />

0<br />

−U dc<br />

2<br />

U Cc<br />

1<br />

S A<br />

0<br />

1<br />

S B<br />

0<br />

1<br />

0<br />

U A<br />

S C<br />

0<br />

2 3U dc<br />

1 3U dc<br />

0<br />

−1 3U dc<br />

−2 3U dc<br />

U dc<br />

U AB<br />

0<br />

−U dc<br />

0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02<br />

Fig. 2.9. Basic waveforms of carrier based sinusoidal PWM<br />

19


2. Voltage Source Inverter Fed Induction Motor Drive<br />

The modulation index m is defined as:<br />

U<br />

m<br />

m = (2.23)<br />

U<br />

m(t)<br />

where:<br />

U<br />

m<br />

- peak value of the modulating wave,<br />

U<br />

m(t)<br />

- peak value of the carrier wave.<br />

The modulation index m can be varied between 0 and 1 to give a linear relation<br />

between the reference and output wave. At m=1, the maximum value of fundamental<br />

U<br />

dc<br />

peak voltage is , which is 78.55% of the peak voltage of the square wave (2.21).<br />

2<br />

The maximum value in the linear range can be increased to 90.7% of that of the<br />

square wave by inserting the appropriate value of a triple harmonics to the modulating<br />

wave. It is shown in Fig. 2.10, which presents the whole range characteristic of the<br />

modulation methods [67]. This characteristic include also the overmodulation (OM)<br />

region, which is widely described in section 2.4.5.<br />

π U<br />

2 U<br />

A<br />

dc<br />

⋅100<br />

100<br />

[%]<br />

90.7<br />

78.5<br />

SVPWM<br />

or SPWM with ZSS<br />

OM<br />

Six step<br />

operation<br />

SPWM<br />

m<br />

1 1.155 3.24 M<br />

0.785 0.907 1<br />

Fig. 2.10. Output voltage of VSI versus modulation index for different PWM techniques<br />

20


2.4. Pulse Width Modulation (PWM)<br />

If the neutral point N on the AC side of the inverter is not connected with the DC<br />

side midpoint 0 (Fig. 2.3), phase currents depend only on the voltage difference<br />

between phases. Therefore, it is possible to insert an additional Zero Sequence Signal<br />

(ZSS) of the 3-th harmonic frequency, which does not produce phase voltage distortion<br />

and without affecting load currents. A block scheme of the modulator based on the<br />

additional ZSS is shown in Fig. 2.11 [46].<br />

U dc<br />

U Ac<br />

U Bc<br />

U<br />

*<br />

Ac<br />

U<br />

*<br />

Bc<br />

U Cc<br />

U Cc<br />

*<br />

S A<br />

S B<br />

S C<br />

A B C<br />

Calculation<br />

of ZSS<br />

Carrier<br />

U t<br />

N<br />

Fig. 2.11. Generalized PWM with additional Zero Sequence Signal (ZSS)<br />

The type of the modulation method depends on the ZSS waveform. The most popular<br />

PWM methods are shown in Fig. 2.12 where unity the triangular carrier waveform gain<br />

U<br />

dc<br />

U<br />

is assumed and the signals are normalized to . Therefore, ± dc<br />

saturation limits<br />

2<br />

2<br />

correspond to ±1. In Fig. 2.12 only phase “A” modulation waveform is shown as the<br />

modulation signals of phase “B” and “C” are identical waveforms with 120º phase shift.<br />

The modulated methods illustrated in Fig. 2.12 can be separated into two groups:<br />

continuous and discontinuous. In the continuous PWM (CPWM) methods, the<br />

modulation waveform are always within the triangular peak boundaries and in every<br />

carrier cycle triangle and modulation waveform intersections. Therefore, on and off<br />

switchings occur. In the discontinuous PWM (DPWM) methods a modulation<br />

waveform of a phase has a segment which is clamped to the positive or negative DC<br />

21


2. Voltage Source Inverter Fed Induction Motor Drive<br />

bus. In this segments some power converter switches do not switch. Discontinuous<br />

modulation methods give lower (average 33%) switching losses. The modulation<br />

method with triangular shape of ZSS with 1/4 peak value corresponds to space vector<br />

modulation (SVPWM) with symmetrical placement of the zero vectors in a sampling<br />

period. It will be widely describe in section 2.4.4. In Fig. 2.12 is also shown sinusoidal<br />

PWM (SPWM) and third harmonic PWM (THIPWM) with sinusoidal ZSS with 1/4<br />

peak value corresponding to a minimum of output current harmonics [63].<br />

a) b) c)<br />

SPWM THIPWM SVPWM<br />

1<br />

0.8<br />

1<br />

0.8<br />

U A<br />

1<br />

0.8<br />

U A<br />

0.6<br />

0.4<br />

0.6<br />

0.4<br />

U A0<br />

0.6<br />

0.4<br />

U A0<br />

0.2<br />

0.2<br />

0.2<br />

0<br />

0<br />

0<br />

-0.2<br />

U N0<br />

U A<br />

=U A0<br />

U N0<br />

U N0<br />

-0.2<br />

-0.2<br />

-0.4<br />

-0.4<br />

-0.4<br />

-0.6<br />

-0.6<br />

-0.6<br />

-0.8<br />

-0.8<br />

-0.8<br />

-1<br />

-1<br />

-1<br />

0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02<br />

Time<br />

0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02<br />

Time<br />

0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02<br />

Time<br />

d) e) f)<br />

DPWM1 DPWM2 DPWM3<br />

1<br />

1<br />

1<br />

0.8<br />

0.8<br />

U A0<br />

0.8<br />

U A0<br />

0.6<br />

0.6<br />

0.6<br />

0.4<br />

0.4<br />

U A0<br />

U A<br />

U A<br />

U A<br />

0.4<br />

0.2<br />

0.2<br />

0.2<br />

0<br />

0<br />

0<br />

-0.2<br />

-0.2<br />

-0.2<br />

-0.4<br />

U N0<br />

-0.4<br />

-0.4<br />

-0.6<br />

-0.8<br />

-0.6<br />

-0.8<br />

U N0<br />

-0.6<br />

-0.8<br />

U N0<br />

-1<br />

-1<br />

-1<br />

0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02<br />

Time<br />

0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02<br />

Time<br />

0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02<br />

Time<br />

Fig. 2.12. Waveforms for PWM with added Zero Sequence Signal a) SPWM, b)THIPWM, c) SVPWM,<br />

d) DPWM1, e) DPWM2, f) DPWM3<br />

2.4.3. <strong>Space</strong> <strong>Vector</strong> Modulation (<strong>SVM</strong>)<br />

The space vector modulation techniques differ from the carrier based in that way,<br />

there are no separate modulators used for each of the three phases. Instead of them, the<br />

reference voltages are given by space voltage vector and the output voltages of the<br />

inverter are considered as space vectors (2.22). There are eight possible output voltage<br />

vectors, six active vectors U 1 - U 6 , and two zero vectors U 0 , U 7 (Fig. 2.13). The<br />

reference voltage vector is realized by the sequential switching of active and zero<br />

vectors.<br />

In the Fig. 2.13 there are shown reference voltage vector U c and eight voltage<br />

vectors, which corresponds to the possible states of inverter. The six active vectors<br />

22


2.4. Pulse Width Modulation (PWM)<br />

divide a plane for the six sectors I - VI. In the each sector the reference voltage vector<br />

U c is obtained by switching on, for a proper time, two adjacent vectors. Presented in<br />

Fig. 2.13 the reference vector U c can be implemented by the switching vectors of U 1 , U 2<br />

and zero vectors U 0 , U 7 .<br />

U 3<br />

(010)<br />

II<br />

U 2<br />

(110)<br />

III<br />

I<br />

U 4<br />

(011)<br />

(t 2<br />

/T s<br />

)U 2<br />

α<br />

U c<br />

U 0<br />

(000) U 1<br />

(100)<br />

U 7<br />

(111) (t 1<br />

/T s<br />

)U 1<br />

IV<br />

VI<br />

U 5<br />

(001) U<br />

V<br />

6<br />

(101)<br />

Fig. 2.13. Principle of the space vector modulation<br />

The reference voltage vector U c is sampled with the fixed clock frequency<br />

and next a sampled value ( )<br />

T s<br />

f = 1 T ,<br />

U is used for calculation of times t 1 , t 2 , t 0 and t 7 . The<br />

c<br />

signal flow in space vector modulator is shown in Fig. 2.14.<br />

s<br />

s<br />

U dc<br />

f s<br />

U c<br />

(T s<br />

)<br />

Sector<br />

selection<br />

S A<br />

S B<br />

S C<br />

U c<br />

A B C<br />

t 1<br />

t 2<br />

t 0<br />

t 7<br />

Calculation<br />

N<br />

Fig. 2.14. Block scheme of the space vector modulator<br />

23


2. Voltage Source Inverter Fed Induction Motor Drive<br />

The times t 1 and t 2 are obtained from simple trigonometrical relationships and can be<br />

expressed in the following equations:<br />

2 3<br />

t<br />

1<br />

= MTs<br />

sin( π 3 −α<br />

)<br />

(2.24a)<br />

π<br />

2 3<br />

t<br />

2<br />

= MTs<br />

sin( α )<br />

(2.24b)<br />

π<br />

Where M is a modulation index, which for the space vector modulation is defined as:<br />

M<br />

U<br />

=<br />

U<br />

c<br />

1( six−step)<br />

U<br />

c<br />

=<br />

2<br />

U<br />

π<br />

dc<br />

(2.25)<br />

where:<br />

U<br />

c<br />

- vector magnitude, or phase peak value,<br />

U1 ( six − step)<br />

- fundamental peak value ( U dc<br />

π )<br />

2 of the square-phase voltage<br />

wave.<br />

The modulation index M varies from 0 to 1 at the square-wave output. The length of<br />

the U c vector, which is possible to realize in the whole range of α is equal to<br />

3<br />

U<br />

3<br />

This is a radius of the circle inscribed of the hexagon in Fig. 2.13. At this condition the<br />

modulation index is equal:<br />

dc<br />

.<br />

M<br />

=<br />

3<br />

U<br />

3<br />

2<br />

U<br />

π<br />

dc<br />

dc<br />

= 0.907<br />

(2.26)<br />

This means that 90.7% of the fundamental at the square wave can be obtained. It<br />

extends the linear range of modulation in relation to 78.55% in the sinusoidal<br />

modulation techniques (Fig. 2.10).<br />

After calculation of t 1 and t 2 from equations (2.24) the residual sampling time is<br />

reserved for zero vectors U 0 and U 7 .<br />

t =<br />

s<br />

− ( + = t + t<br />

(2.27)<br />

0 ,7<br />

T t1<br />

t2<br />

)<br />

0<br />

7<br />

24


2.4. Pulse Width Modulation (PWM)<br />

The equations for t 1 and t 2 are identically for all space vector modulation methods.<br />

The only difference between the other type of <strong>SVM</strong> is the placement of zero vectors at<br />

the sampling time.<br />

The basic <strong>SVM</strong> method is the modulation method with symmetrical spacing zero<br />

vectors (SVPWM). In this method times t 0 and t 7 are equal:<br />

( T − t ) 2<br />

t = =<br />

s<br />

−<br />

(2.28)<br />

0<br />

t7<br />

1<br />

t2<br />

For the first sector switching sequence can be written as follows:<br />

U 0 → U 1 → U 2 → U 7 → U 2 → U 1 → U 0 (2.29)<br />

This vector switching sequence in the SVPWM method is shown in Fig. 2.15a. In<br />

this case zero vectors are placed in the beginning and in the end of period U 0 , and in the<br />

center of the period U 7 . In one sampling period all three phases are switched. To realize<br />

the reference vector can also be used an other switching sequence, for example:<br />

U 0 → U 1 → U 2 → U 1 → U 0 (2.30)<br />

or<br />

U 1 → U 2 → U 7 → U 2 → U 1 (2.31)<br />

These sequences are shown in Fig. 2.15b and 2.15c respectively. In these cases only<br />

two phases switch in one sampling time, and only one zero vector is used U 0 (Fig.<br />

2.15b) or U 7 (Fig. 2.15c). This type of modulation is called discontinuous pulse width<br />

modulation (DPWM).<br />

a) b) c)<br />

S A<br />

0<br />

1 1 1<br />

1<br />

1<br />

1<br />

0<br />

S A<br />

0 1 1<br />

1<br />

0<br />

S A<br />

1<br />

1 1 1<br />

1<br />

1<br />

1<br />

1<br />

S B<br />

0 0<br />

1 1<br />

1<br />

1<br />

0<br />

0<br />

S B<br />

0<br />

0 1<br />

0<br />

0<br />

S B<br />

0 1<br />

1 1<br />

1<br />

1<br />

1<br />

0<br />

S C<br />

0 0 0<br />

1<br />

1<br />

0<br />

0<br />

0<br />

S C<br />

0 0<br />

0<br />

0<br />

0<br />

0 0 1<br />

1<br />

1<br />

1<br />

0<br />

0<br />

t 0 /4 t 1 /2 t 2 /2 t 0 /4<br />

t 0 /4 t 1 /2 t 2 /2 t 0 /4<br />

t 0 /2 t 1 /2<br />

t 2 t 1 /2 t 0 /2<br />

t 1 /2 t 2 /2<br />

S C<br />

U 1 U 2 U 2 U 1<br />

t 0 t 2 /2<br />

t 1 /2<br />

T s<br />

T s<br />

T s<br />

U 0 U 1 U 2 U 7 U 7 U 2 U 1 U 0<br />

U 0 U 1 U 2 U 1 U 0<br />

U 7<br />

Fig. 2.15. <strong>Space</strong> vectors in the sampling period a) SVPWM, b), c) DPWM<br />

The idea of discontinuous modulation is based on the assumption that one phase is<br />

clamped by 60° to lower or upper of the dc bus voltage. It gives only one zero state per<br />

sampling period (Fig. 2.15b, c). The discontinuous modulation provides 33% reduction<br />

25


2. Voltage Source Inverter Fed Induction Motor Drive<br />

of the effective switching frequency and switching losses. The discontinuous space<br />

vector modulation techniques, like all the space vector methods, correspond to the<br />

carrier based modulation method. It will be widely described in the next section.<br />

a)<br />

DPWM1<br />

U 3 (010)<br />

t 0 = 0<br />

t 7 = 0<br />

U 2 (110)<br />

1<br />

0.8<br />

U A0<br />

U A<br />

t 0 = 0<br />

t 7 = 0<br />

0.6<br />

0.4<br />

U 4 (011)<br />

t 7 = 0<br />

t 0 = 0<br />

U 0 (000) U 1 (100)<br />

U 7 (111)<br />

0.2<br />

0<br />

-0.2<br />

t 7 = 0<br />

t 0 = 0<br />

-0.4<br />

U N0<br />

t 0 = 0<br />

t 7 = 0<br />

-0.6<br />

-0.8<br />

t 0 = 0<br />

t 7 = 0<br />

U 6 (101)<br />

-1<br />

U 5 (001)<br />

0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02<br />

Time<br />

b)<br />

DPWM2<br />

U 3 (010)<br />

t 7 = 0<br />

U 2 (110)<br />

1<br />

0.8<br />

U A0<br />

0.6<br />

t 0 = 0 t 0 = 0<br />

0.4<br />

U A<br />

U 4 (011)<br />

U 0 (000) U 1 (100)<br />

0.2<br />

0<br />

U 7 (111)<br />

-0.2<br />

t 7 = 0<br />

U 5 (001)<br />

t 0 = 0<br />

t 7 = 0<br />

U 6 (101)<br />

-0.4<br />

-0.6<br />

U<br />

-0.8<br />

N0<br />

-1<br />

0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02<br />

Time<br />

c)<br />

DPWM3<br />

t 7 = 0<br />

U 3 (010)<br />

t 7 = 0<br />

t 0 = 0<br />

t 0 = 0<br />

U 2 (110)<br />

1<br />

0.8<br />

0.6<br />

U A0<br />

U A<br />

U 4 (011)<br />

t 0 = 0<br />

t 0 = 0<br />

t 7 = 0<br />

U 0 (000) U 1 (100)<br />

U 7 (111)<br />

t 7 = 0<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

-0.4<br />

t 7 = 0<br />

t 7 = 0<br />

U 5 (001)<br />

t 0 = 0<br />

t 0 = 0<br />

U 6 (101)<br />

-0.6<br />

U<br />

-0.8<br />

N0<br />

-1<br />

0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02<br />

Time<br />

d)<br />

DPWM4<br />

U 3 (010)<br />

t 0 = 0<br />

U 2 (110)<br />

1<br />

0.8<br />

0.6<br />

U A<br />

U A0<br />

U 4 (011)<br />

t 7 = 0<br />

t 7 = 0<br />

U 0 (000) U 1 (100)<br />

U 7 (111)<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

t 0 = 0<br />

t 0 = 0<br />

-0.4<br />

-0.6<br />

U N0<br />

U 5 (001)<br />

t 7 = 0<br />

U 6 (101)<br />

-0.8<br />

-1<br />

0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 0.016 0.018 0.02<br />

Time<br />

Fig. 2.16. The discontinuous space vector modulation<br />

26


2.4. Pulse Width Modulation (PWM)<br />

In the Fig. 2.16 there are shown several different kinds of space vector discontinues<br />

modulation. It can be seen that the type of method depends on the moved do not switch<br />

sectors. These sectors are adequately moved on 0°, 30°, 60°, 90° and denoted as<br />

DPWM1, DPWM2, DPWM3 and DPWM4. Fig. 2.16 also shows voltage waveforms for<br />

each methods. For the carrier based methods with ZSS these waveforms are identical<br />

(Fig. 2.12).<br />

From the type of modulation it depends also harmonic content, what is presented in<br />

Fig. 2.17 for the SVPWM and DPWM1 methods.<br />

Fig. 2.17. The output line to line voltage harmonics content a) SVPWM, b) DPWM 1<br />

In Fig. 2.17 harmonics of output line to line voltage are shown. The voltage<br />

frequency domain representation is composed of the series discrete harmonics<br />

components. These are clustered about multiplies of the switching frequency. In this<br />

case the switching frequency was 5 kHz. Spectrum for every modulation methods is<br />

different. In Fig. 2.17 the differences between SVPWM and DPWM1 modulation<br />

method can be seen. However, characteristic feature for all methods, which work with<br />

constant switching frequency is clustered higher harmonics round multiplies of the<br />

switching frequency. These type of modulation methods are named deterministic PWM<br />

(DEPWM). The modulation method influence also for current distortion, torque ripple<br />

and acoustic noise emitted from the motor. Modulation techniques are still being<br />

improved for reduction of these disadvantages. One of the proposed methods is a<br />

random PWM (RPWM) (see section 2.4.6).<br />

27


2. Voltage Source Inverter Fed Induction Motor Drive<br />

2.4.4. Relation Between Carrier Based and <strong>Space</strong> <strong>Vector</strong> Modulation<br />

All the carrier based methods have equivalent to the space vector modulation<br />

methods. The type of carrier based method depends on the added ZSS, as shown in<br />

section 2.4.2, and type of the space vector modulation depending on the time of zero<br />

vectors t 0 and t 7 .<br />

A comparison of carrier based method with <strong>SVM</strong> is shown in Fig 2.18. There is<br />

shown a carrier based modulation with triangular shape of ZSS with 1/4 peak value.<br />

This method corresponds to the space vector modulation (SVPWM) with symmetrical<br />

placement of zero vectors in sampling period. In Fig. 2.18b is presented discontinuous<br />

method DPWM1 for carrier based and for <strong>SVM</strong> techniques.<br />

In the carrier based methods three reference signals U * Ac , U * Bc , U * Cc are compared<br />

with triangular carrier signal U t , and in this way logical signals S A , S B , S C are generated.<br />

In the space vector modulation duration time of active (t 1 , t 2 ) and zero (t 0 , t 7 ) vectors are<br />

calculated, and from these times switching signals S A , S B , S C are obtained. The gate<br />

pulses generated by both methods are identical.<br />

The carrier based PWM methods are simple to implement in hardware. Through the<br />

compare reference signals with triangular carrier signal it receives gate pulses.<br />

However, a PWM inverter is generally controlled by a microprocessor/controller<br />

nowadays. Thanks to the representation of command voltages as space vector, a<br />

microprocessor using suitable equations can calculate duration time and realize<br />

switching sequence easily.<br />

It is possible to implement all carrier based modulation methods using the space<br />

vector technique. The active vector times t 1 and t 2 equations are identically for all space<br />

vector modulation methods. But every method demand suitable equation for the zero<br />

vectors t 0 and t 7 .<br />

The eight voltage vectors U 0 - U 7 correspond to the possible states of the inverter<br />

(Fig. 2.13). Each of these states can be composed by a different equivalent electrical<br />

circuit. In Fig 2.19 the circuit for the vector U 1 is presented.<br />

28


2.4. Pulse Width Modulation (PWM)<br />

a)<br />

b)<br />

S A<br />

S B<br />

Carrir based PWM<br />

S A<br />

S B<br />

Carrir based PWM<br />

S C<br />

S C<br />

U Ac<br />

*<br />

U Bc<br />

*<br />

U Ac<br />

*<br />

U Cc<br />

*<br />

U Bc<br />

*<br />

U Cc<br />

*<br />

S A<br />

0<br />

1 1 1<br />

1<br />

1<br />

1<br />

0<br />

S A<br />

0<br />

0 1 1<br />

1<br />

1<br />

0<br />

0<br />

S B<br />

S C<br />

0 0 1 1<br />

0 0 0 1<br />

t 0<br />

/4 t 1<br />

/2 t 2<br />

/2 t 0<br />

/4<br />

1 1 0 0<br />

1 0 0 0<br />

t 0<br />

/4 t 1<br />

/2 t 2<br />

/2 t 0<br />

/4<br />

<strong>Space</strong> vector PWM<br />

S B<br />

S C<br />

0 0 0 1<br />

0 0 0 0<br />

t 0<br />

/2 t 1<br />

/2<br />

1 0 0 0<br />

0 0 0 0<br />

t 2<br />

t 1<br />

/2 t 0<br />

/2<br />

<strong>Space</strong> vector PWM<br />

T s<br />

T s<br />

U 0<br />

U 1<br />

U 2<br />

U 7<br />

U 7<br />

U 2<br />

U 1<br />

U 0<br />

U 0<br />

U 1<br />

U 2<br />

U 1<br />

U 0<br />

Fig. 2.18. Comparison of carrier based PWM with space vector PWM a) SVPWM, b) DPWM1<br />

A<br />

U A0<br />

U dc<br />

2<br />

U A<br />

0<br />

U N0<br />

N<br />

U dc<br />

2<br />

U B<br />

U C<br />

U B0<br />

=U C0<br />

B<br />

C<br />

Fig. 2.19. Equivalent circuit of VSI for the U 1 vector<br />

29


2. Voltage Source Inverter Fed Induction Motor Drive<br />

Taking into consideration the electrical circuit in Fig. 2.19 the voltage distribution<br />

can be obtained. The voltages can be written as:<br />

2<br />

1<br />

1<br />

U A<br />

= U dc<br />

; U B<br />

= − Udc<br />

; U C<br />

= − U<br />

dc<br />

(2.32)<br />

3<br />

3<br />

3<br />

1<br />

1<br />

1<br />

U A0<br />

= U dc<br />

; U B0<br />

= − Udc<br />

; U C0<br />

= − U<br />

dc<br />

(2.33)<br />

2<br />

2<br />

2<br />

U<br />

N0<br />

1<br />

= U<br />

A0<br />

−U<br />

AN<br />

= − Udc<br />

(2.34)<br />

6<br />

This analysis may be repeated for all vectors provided to obtain voltages presented in<br />

Table 2.1.<br />

Table 2.1. The voltages for the eight converter output vectors<br />

U 0<br />

U 1<br />

U 2<br />

U 3<br />

U 4<br />

U 5<br />

U 6<br />

U 7<br />

U<br />

A0<br />

1<br />

−<br />

2<br />

1<br />

U dc<br />

2<br />

1<br />

U dc<br />

2<br />

1<br />

−<br />

2<br />

1<br />

−<br />

2<br />

1<br />

−<br />

2<br />

1<br />

U dc<br />

2<br />

1<br />

U dc<br />

2<br />

U dc<br />

U dc<br />

U dc<br />

U dc<br />

U<br />

B0<br />

UC0<br />

U<br />

A<br />

U<br />

B<br />

UC<br />

U<br />

N0<br />

1 1<br />

1<br />

− U<br />

dc<br />

− U dc 0 0 0 − U<br />

dc<br />

2 2<br />

2<br />

1 1 2 1 1 1<br />

− U dc<br />

− U dc U<br />

dc<br />

− U dc − U<br />

dc<br />

− U<br />

dc<br />

2 2 3 3 3 6<br />

1 1 1 1 2 1<br />

U dc − U dc<br />

U dc<br />

U dc<br />

− U dc U<br />

dc<br />

2 2 3 3 3 6<br />

1 1 1 2 1 1<br />

U dc<br />

− U dc − U dc<br />

U<br />

dc − U<br />

dc<br />

− U<br />

dc<br />

2 2 3 3 3 6<br />

1 1 2 1 1 1<br />

U dc<br />

U dc<br />

− U<br />

dc U<br />

dc<br />

U dc<br />

U dc<br />

2 2 3 3 3 6<br />

1 1 1 1 2 1<br />

− U<br />

dc<br />

U dc<br />

− U dc<br />

− U dc<br />

U<br />

dc<br />

U<br />

dc<br />

2 2 3 3 3 6<br />

1 1 1 2 1 1<br />

− U dc<br />

U dc<br />

U dc<br />

− U dc<br />

U<br />

dc<br />

U<br />

dc<br />

2 2 3 3 3 6<br />

1 1<br />

1<br />

U dc<br />

U dc 0 0 0 U<br />

dc<br />

2 2<br />

2<br />

The average value for sampling time of U NO voltage can be written as follows:<br />

U<br />

N0<br />

1 U<br />

dc ⎛ 1 1<br />

= ⎜−<br />

t0<br />

− t1<br />

+ t2<br />

+ t7<br />

T 2 ⎝ 3 3<br />

s<br />

⎞<br />

⎟<br />

⎠<br />

for the sectors I, III, V (2.35)<br />

and<br />

U<br />

N0<br />

1 U<br />

dc ⎛ 1 1<br />

= ⎜−<br />

t0<br />

− t2<br />

+ t1<br />

+ t7<br />

T 2 ⎝ 3 3<br />

s<br />

⎞<br />

⎟ for the sectors II, IV, VI (2.36)<br />

⎠<br />

30


2.4. Pulse Width Modulation (PWM)<br />

From the above equations and taking into consideration equations (2.24) and (2.27)<br />

the zero vectors times for different kinds of modulation can be calculated.<br />

Relations between carrier based and <strong>SVM</strong> methods are presented in Table 2.2. This<br />

table presents also the zero vector (t 0 , t 7 ) times equations for the most significant<br />

modulation methods.<br />

Table 2.2. Relation between carrier based and <strong>SVM</strong> methods<br />

Modulation<br />

method<br />

Waveform of the<br />

ZSS (Fig. 2.13)<br />

Calculation of t 0<br />

and t 7<br />

SPWM<br />

no signal<br />

( U N0<br />

= 0)<br />

t<br />

t<br />

0<br />

0<br />

T ⎛ ⎞<br />

=<br />

s<br />

4<br />

⎜1<br />

− M cos α ⎟<br />

2 ⎝ π ⎠<br />

Ts<br />

⎛ 2<br />

= ⎜1−<br />

M<br />

α<br />

2 ⎝ π<br />

⎞<br />

( cosα<br />

) + 3 sin ⎟<br />

⎠<br />

for sectors I, III, V<br />

for sectors II, IV, VI<br />

t<br />

7<br />

= Ts<br />

− t0<br />

− t1<br />

− t2<br />

THIPWM<br />

SVPWM<br />

Sinusoidal with<br />

1/4 amplitude<br />

Triangular with<br />

1/4 amplitude<br />

t<br />

t<br />

t<br />

0<br />

0<br />

Ts<br />

⎛ 4 ⎛ 1 ⎞⎞<br />

= ⎜1−<br />

M ⎜cosα<br />

− cos3α<br />

⎟⎟<br />

for sectors I, III, V<br />

2 ⎝ π ⎝ 4 ⎠⎠<br />

Ts<br />

⎛ 2 ⎛<br />

1 ⎞⎞<br />

= ⎜1−<br />

M ⎜cosα<br />

+ 3 sinα<br />

− cos3α<br />

⎟⎟<br />

2 ⎝ π ⎝<br />

2 ⎠⎠<br />

for sectors II, IV, VI<br />

7<br />

= Ts<br />

− t0<br />

− t1<br />

− t2<br />

t<br />

( T − t − ) 2<br />

0<br />

= t7<br />

=<br />

s 1<br />

t2<br />

DPWM1<br />

Discontinuous<br />

t 0<br />

t<br />

= 0<br />

7<br />

= Ts<br />

− t1<br />

− t2<br />

t 7<br />

t<br />

= 0<br />

0<br />

= Ts<br />

− t1<br />

− t2<br />

π<br />

3<br />

π<br />

6<br />

when n ≤ α < ( 2n<br />

+ 1)<br />

π<br />

6<br />

π<br />

3<br />

when ( 2n + 1) ≤ α < ( n + 1)<br />

n = 0, 1,<br />

2,<br />

3,<br />

4,<br />

5<br />

Waveforms of the ZSS presented in Table 2.2 are shown in Fig. 2.12.<br />

2.4.5. Overmodulation (OM)<br />

At the end of the linear range (Fig. 2.10) the inverter output voltage is 90.7% of the<br />

maximum output peak voltage in six-step mode (see equation 2.21). The nonlinear<br />

31


2. Voltage Source Inverter Fed Induction Motor Drive<br />

range between this point and six-step mode is called overmodulation. This part of the<br />

modulation techniques is not so important in vector controlled drives methods for the<br />

sake of big distortion current and torque. For example, the overmodulation can be<br />

applied in drives working in open loop control mode to increase the value of inverter<br />

output voltage.<br />

The overmodulation has been widely discussed in the literature [16, 33, 55, 75, 89].<br />

Some of methods are proposed as extensions of the carrier based modulation and others<br />

as extensions of space vector modulation. In the carrier based methods overmodulation<br />

algorithm is realized by increasing reference voltage beyond the amplitude of the<br />

triangular carrier signal. In this case some switching cycles are omitted and each phase<br />

is clamped to one of the dc busses.<br />

The overmodulation region for space vector modulation is shown in Fig. 2.20. The<br />

maximum length of vector U c possible to realization in whole range of α angle is equal<br />

3<br />

3<br />

U dc<br />

. It is a radius of the circle inscribed of the hexagon. This value corresponds to<br />

the modulation index equal to 0.907 (see equation 2.26). To realize higher values a<br />

voltage overmodulation algorithm has to be applied. At the end of the overmodulation<br />

region is a six-step mode (at M = 1).<br />

U 3<br />

(010)<br />

U 2<br />

(110)<br />

Six-step mode<br />

M = 1<br />

U 4<br />

(011)<br />

U 0<br />

(000) U 1<br />

(100)<br />

U 7<br />

(111)<br />

(t 2<br />

/T s<br />

)U 2<br />

α<br />

(t 1<br />

/T s<br />

)U 1<br />

U c<br />

Overmodulation range<br />

0.907 < M < 1<br />

Linear range<br />

M ≤ 0.907<br />

U 5<br />

(001) U 6<br />

(101)<br />

Fig. 2.20. Definition of the overmodulation range<br />

32


2.4. Pulse Width Modulation (PWM)<br />

If the value of the reference voltage beyond maximal value in the linear range vector<br />

U c can not be realized for whole range of α angle. However, average voltage value can<br />

be obtained for modification of the reference voltage vector. Because of the modified<br />

reference voltage vector overmodulation algorithms are not widely used in vector<br />

control methods of drive. To modify the reference voltage vector different algorithm<br />

may be applied. Overmodulation range can be considered as one region [33], or it can<br />

be divided into two regions [16, 55, 75, 89].<br />

In the algorithm where overmodulation region is considered as two regions two<br />

modes depending on the reference voltage value were defined. In mode I the algorithm<br />

modifies only the voltage vector amplitude, in mode II both the amplitude and angle of<br />

the voltage vector.<br />

Overmodulation mode I is shown in Fig. 2.21.<br />

U 3<br />

(010)<br />

U 2<br />

(110)<br />

U c<br />

U 4<br />

(011)<br />

U 0<br />

(000)<br />

α<br />

U c<br />

*<br />

θ<br />

U 1<br />

(100)<br />

U 7<br />

(111)<br />

U 5<br />

(001) U 6<br />

(101)<br />

Fig. 2.21. Overmodulation mode I<br />

In this mode voltage vector U c crosses the hexagon boundary at two points in each<br />

sector. There is a loss of fundamental voltage in the region where reference vector<br />

exceeds the hexagon boundary. To compensate for this loss, the reference vector<br />

amplitude is increased in the region where the reference vector is in hexagon boundary.<br />

A modified reference voltage trajectory proceeds partly on the hexagon and partly on<br />

the circle. This trajectory is shown in Fig. 2.21.<br />

33


2. Voltage Source Inverter Fed Induction Motor Drive<br />

In the hexagon trajectory part only active vectors are used. The duration of these<br />

vectors t 1 and t 2 are obtained from trigonometrical relationships and can be expressed in<br />

the following equations:<br />

3 cosα<br />

− sinα<br />

t 1<br />

= T s<br />

(2.37a)<br />

3 cosα<br />

+ sinα<br />

t<br />

= −<br />

(2.37b)<br />

2<br />

Ts<br />

t1<br />

t 0<br />

= t 7<br />

= 0<br />

(2.37c)<br />

The output voltage waveform is given approximately by linear segments for the<br />

hexagon trajectory and sinusoidal segments for the circular trajectory. Boundary of the<br />

segments is determined by a crossover angle θ which depends on the reference voltage<br />

value. As known from Fig. 2.21 the upper limit in mode I is when θ = 0°. Then the<br />

voltage trajectory is fully on the hexagon. The fundamental peak value generated in this<br />

way voltage is 95% of the peak voltage of the square wave [75]. It gives modulation<br />

index M = 0.952.<br />

For the modulation index higher then 0.952 the overmodulation mode II is applied.<br />

The overmodulation mode II is shown in Fig. 2.22. In this mode not only the reference<br />

vector amplitude is modified but also an angle. The reference angle from α to α *<br />

changed.<br />

is<br />

U 3<br />

(010)<br />

U 2<br />

(110)<br />

α h<br />

U c<br />

*<br />

U c<br />

U 4<br />

(011)<br />

U 0<br />

(000)<br />

α<br />

∗<br />

α<br />

α h<br />

U 1<br />

(100)<br />

U 7<br />

(111)<br />

U 5<br />

(001) U 6<br />

(101)<br />

Fig. 2.22. Overmodulation mode II where both amplitude and angle is changed<br />

34


2.4. Pulse Width Modulation (PWM)<br />

The trajectory of U * c is maintained on the hexagon which defines amplitude of the<br />

reference voltage vector. The angle is calculated from the following equations:<br />

∗<br />

α<br />

⎧<br />

⎪0<br />

⎪<br />

⎪ α −α<br />

h π<br />

= ⎨<br />

⎪π<br />

6 −α<br />

h<br />

6<br />

⎪<br />

⎪<br />

π 3<br />

⎩<br />

for<br />

0 ≤ α ≤ α<br />

α < α < π 3 −α<br />

h<br />

π 3 −α<br />

≤ α ≤ π 3<br />

h<br />

h<br />

h<br />

(2.38)<br />

where: α h <strong>–</strong> hold-angle.<br />

This angle uniquely controls the fundamental voltage. It is a nonlinear function of the<br />

modulation index [16, 55].<br />

In Fig. 2.22 is shown the reference vector trajectory generated for the first sector.<br />

This trajectory is obtained in three steps. First part, if angle α is less than the respective<br />

value of α h , the algorithm holds the vector U c * at the vertex (U 1 ). Next part is for α from<br />

α h to π 3 −α<br />

h<br />

. In this angle range the reference vector moves along the hexagon. In the<br />

last range, from π 3 −α<br />

h<br />

to α<br />

h<br />

, the vector U * c is held until the next vertex (U 2 ).<br />

The overmodulation mode II works up to the six-step mode for α h equal zero. The<br />

six-step mode characterized by selection of the switching vector for one-sixth of the<br />

fundamental period. In this case the maximum possible inverter output voltage is<br />

generated.<br />

2.4.6. Random Modulation Techniques<br />

The pulse width modulation technique is important for drive performance in respect<br />

to voltage and current harmonics, torque ripple, acoustic noise emitted from an<br />

induction motor and also electromagnetic interference (EMI). Different approaches<br />

were used in PWM techniques for reduction of these disadvantages. One of the<br />

proposed methods is random pulse width modulation (RPWM) [5, 7, 11, 14, 61, 68,<br />

104].<br />

Previously presented modulation methods were named deterministic pulse width<br />

modulation (DEPWM), because of constant sampling and switching frequency and all<br />

35


2. Voltage Source Inverter Fed Induction Motor Drive<br />

cycles the switching sequence is deterministic. In RPWM methods the switching<br />

frequency or the switching sequence change randomly.<br />

One of the proposed random modulation techniques is a method with randomly<br />

varied lengths of coincident switching and sampling time of the modulator. This method<br />

was named RPWM 1. The sampling and switching cycles in DEPWM with RPWM 1 is<br />

comparable shown in Fig. 2.23. The reference voltage vectors U c , which are calculated<br />

in one sampling time T s and realized in the next switching time T sw are shown. In drive<br />

systems the controller mostly operates in synchronism with modulator and in RPWM 1<br />

arises problems in the control system, when it works with variable sampling frequency.<br />

An additional control algorithm with variable sampling frequency is difficult tin a<br />

digital implementation.<br />

a)<br />

(1)<br />

U c<br />

(2)<br />

U c<br />

(3)<br />

U c<br />

(K)<br />

U c<br />

( n−1)<br />

U c<br />

(n)<br />

U c<br />

( n+1)<br />

U c<br />

sampling cycles 1 2 3 ... n-1 n ...<br />

switching cycles 1 2 3 ... n-1 n ...<br />

T<br />

s<br />

= T sw<br />

b)<br />

(1)<br />

U c<br />

(2)<br />

U c<br />

(3)<br />

U c<br />

(K)<br />

U c<br />

( n−1)<br />

U c<br />

(n)<br />

U c<br />

( n+1)<br />

U c<br />

sampling cycles 1 2 3 ... n-1 n<br />

switching cycles 1 2 3 ... n-1 n ...<br />

T<br />

s<br />

= T sw<br />

...<br />

Fig. 2.23. Sampling and switching cycles a) DEPWM, b) RPWM 1<br />

For elimination of these disadvantages random modulation techniques were<br />

proposed, which operate with a fixed switching and sampling frequency. These methods<br />

randomly change switching sequence in the interval. Three of these methods are shown<br />

in Fig. 2.24 [6].<br />

First of them (Fig. 2.24a) is random lead-lag modulation (RLL). In this method pulse<br />

position is either commencing at the beginning of the switching interval (leading-edge<br />

36


2.4. Pulse Width Modulation (PWM)<br />

modulation) or its tailing edge is aligned with the end of the interval (lagging-edge<br />

modulation). A random number generator controls the choice between leading and<br />

legging edge modulation.<br />

In Fig. 2.24b is shown a random center pulse displacement (RCD) method. In this<br />

technique pulses are generated identically as in the SVPWM method (Fig. 2.15), but<br />

common pulse center is displaced by the amount α Ts<br />

from the middle of the period.<br />

The parameter α is varied randomly within a band limited by the maximum duty cycle.<br />

The last presented method (Fig. 2.24c) is random distribution of the zero voltage<br />

vector (RZD). Additionally distribution of the zero vectors can by different, until only<br />

one zero vector in switching cycle in the discontinuous methods (Fig. 2.15b, c). This<br />

fact is utilized in the random distribution of the zero voltage vector, where the<br />

proportion between the time duration for the two zero vectors U 0 (000) and U 7 (111) is<br />

randomized in the switching cycles.<br />

a)<br />

Lead Lag Lag Lead<br />

S A<br />

S B<br />

S C<br />

T s<br />

T s<br />

T s<br />

T s<br />

b)<br />

αT s<br />

αTs<br />

αTs<br />

αTs<br />

S A<br />

S B<br />

S C<br />

T s<br />

T s<br />

T s<br />

T s<br />

c)<br />

S A<br />

S B<br />

S C<br />

T s<br />

T s<br />

T s<br />

T s<br />

Fig. 2.24. Different fixed switching random modulation schemes a) Random lead-leg modulation (RLL),<br />

b) Random center displacement (RCD), c) Random zero vector distribution (RZD)<br />

37


2. Voltage Source Inverter Fed Induction Motor Drive<br />

The main disadvantage of the RPWM 1 method (Fig. 2.23b) is variable switching<br />

frequency. For elimination of this disadvantage RPWM 2 [119] was proposed, which<br />

operates with fixed sampling frequency and variable switching frequency. The principle<br />

of this method is shown in Fig. 2.25.<br />

T s<br />

(1)<br />

U c<br />

(2)<br />

U c<br />

(3)<br />

U c<br />

(K)<br />

U c<br />

( n−1)<br />

U c<br />

(n)<br />

U c<br />

( n+1)<br />

U c<br />

sampling cycles<br />

switching cycles<br />

∆t<br />

1 2 3 ... n-1 n ...<br />

1 2 3 ... n-1 n<br />

T sw<br />

...<br />

Fig. 2.25. Sampling and switching cycles in RPWM 2 technique<br />

In this method the start of each switching cycles is delayed with respect to that of the<br />

coincident sampling cycle by a random varied time interval<br />

∆ t . It is given as:<br />

∆ t = rT s<br />

(2.39)<br />

where r denotes a random number between 0 and 1. Time interval<br />

the sake of minimum switching time of inverter.<br />

∆ t is limited for<br />

Fig. 2.26. The output line to line voltage harmonics content a) RPWM 1, b) RPWM 2<br />

Corresponding spectra for the RPWM 1 and RPWM 2 techniques are shown in Fig.<br />

2.26a and 2.26b respectively. It can be seen that the harmonic clusters typical for the<br />

determination modulation (compared to Fig. 2.17) are practically eliminated by the<br />

38


2.5. Summary<br />

random modulation techniques. Simulation result presented in both figures (Fig. 2.17<br />

and Fig. 2.26) was done at the same conditions: sampling frequency 5 kHz, output<br />

frequency 50 Hz.<br />

2.5. Summary<br />

In this chapter mathematical description of IM based on complex space vectors was<br />

presented. The complete equations set is the basis of further consideration of control<br />

and estimation methods.<br />

The structure of two levels voltage source inverter was presented. The main features<br />

and voltage forming methods were described. For the sake of dead-time and voltage<br />

drop on the semiconductor devices the inverter has nonlinear characteristic. Therefore,<br />

in control scheme compensation algorithms are needed.<br />

The inverter is controlled by pulse width modulation (PWM) technique. The<br />

modulation methods are divided into two groups: triangular carrier based and space<br />

vector modulation. Between those two groups there are simple relations. All the carrier<br />

based methods have equivalent to the space vector modulation methods. The type of<br />

carrier based method depends on the added ZSS and type of the space vector<br />

modulation depends on the placement of zero vectors in the sampling period. Presented<br />

modulation methods will be used in the final drive.<br />

This chapter contains compete review of the modulation techniques, including some<br />

random modulation methods. Those methods have very interesting advantages and can<br />

be implemented in special application of IM drives. Currently they have not been<br />

implemented in a presented serially produced drive. However, it will be offered as an<br />

option in a near future. Some experimental results for the implemented modulation<br />

methods are shown in Chapter 7.<br />

39


3. <strong>Vector</strong> Control Methods of Induction Motor<br />

3.1. Introduction<br />

In this chapter review of the most significant IM vector control method is presented.<br />

According to the classification presented in Chapter 1. The theoretical basis and short<br />

characteristic for all methods are given. The direct torque control (<strong>DTC</strong>) method creates<br />

a base for further analyze of <strong>DTC</strong>-<strong>SVM</strong> algorithms. Therefore, <strong>DTC</strong> is more detailed<br />

discussed (see section 3.4).<br />

3.2. Field Oriented Control (FOC)<br />

The principle of the field oriented control (FOC) is based on an analogy to the<br />

separately excited dc motor. In this motor flux and torque can be controlled<br />

independently. The control algorithm can be implemented using simple regulators, e.g.<br />

PI-regulators.<br />

In induction motor independent control of flux and torque is possible in the case of<br />

coordinate system is connected with rotor flux vector. A coordinate system<br />

d − q is<br />

rotating with the angular speed equal to rotor flux vector angular speed<br />

Ω = Ω ,<br />

K<br />

sr<br />

which is defined as follows:<br />

dγsr<br />

Ωsr = (3.1)<br />

dt<br />

The rotating coordinate system<br />

d − q is shown in Fig. 3.1.<br />

The voltage, current and flux complex space vector can be resolved into components<br />

d and q.<br />

U = U + jU<br />

(3.2a)<br />

sK<br />

sd<br />

sq<br />

I = I + jI<br />

, I<br />

rK<br />

= I<br />

rd<br />

+ jIrq<br />

(3.2b)<br />

sK<br />

sd<br />

sq<br />

Ψ<br />

sK<br />

= Ψ<br />

sd<br />

+ jΨ<br />

sq<br />

,<br />

rK<br />

= Ψ<br />

rd<br />

= Ψ<br />

r<br />

Ψ (3.2c)


3.2. Field Oriented Control (FOC)<br />

β<br />

q<br />

I sβ<br />

I s<br />

d<br />

Ω sr<br />

I sq<br />

δ<br />

I sd<br />

Ψ r<br />

γ sr<br />

I sα<br />

α<br />

Fig. 3.1. <strong>Vector</strong> diagram of induction motor in stationary<br />

α − β and rotating d − q coordinates<br />

In<br />

d − q coordinate system the induction motor model equations (2.10-2.12) can be<br />

written as follows:<br />

U<br />

sd<br />

dΨ<br />

dt<br />

sd<br />

= RsI<br />

sd<br />

+ − ΩsrΨ<br />

sq<br />

(3.3a)<br />

dΨ<br />

sq<br />

U<br />

sq<br />

= RsI<br />

sq<br />

+ + ΩsrΨ<br />

sd<br />

dt<br />

(3.3b)<br />

dΨ<br />

r<br />

0 = Rr Ird<br />

+<br />

dt<br />

(3.3c)<br />

r<br />

rq<br />

r<br />

( Ω − p Ω )<br />

0 = R I + Ψ<br />

(3.3d)<br />

sr<br />

b<br />

m<br />

Ψ = L I + L I<br />

(3.4a)<br />

sd<br />

sq<br />

s<br />

s<br />

sd<br />

sq<br />

M<br />

M<br />

rd<br />

Ψ = L I + L I<br />

(3.4b)<br />

rq<br />

Ψ = L I + L I<br />

(3.4c)<br />

r<br />

r<br />

r<br />

rq<br />

rd<br />

M<br />

M<br />

sq<br />

sd<br />

0 = L I + L I<br />

(3.4d)<br />

dΩ<br />

dt<br />

m<br />

1 ⎡ m<br />

⎤<br />

s<br />

LM<br />

= ⎢ pb<br />

Ψ<br />

rI<br />

sq<br />

− M<br />

L ⎥<br />

J ⎣ 2 Lr<br />

⎦<br />

(3.5)<br />

The equations 3.3c and 3.4c can be easy transformed to:<br />

41


3. <strong>Vector</strong> Control Methods of Induction Motor<br />

dΨ<br />

dt<br />

r<br />

L<br />

R<br />

R<br />

M r<br />

r<br />

= I<br />

sd<br />

− Ψ<br />

r<br />

(3.6)<br />

Lr<br />

Lr<br />

The motor torque can by expressed by rotor flux magnitude<br />

component I<br />

sq<br />

as follows:<br />

Ψ<br />

r<br />

and stator current<br />

M<br />

e<br />

m<br />

L<br />

s M<br />

= pb<br />

Ψ<br />

rIsq<br />

(3.7)<br />

2 Lr<br />

Equations (3.6) and (3.7) are used to construct a block diagram of the induction<br />

motor in<br />

d − q coordinate system, which is presented in Fig. 3.2.<br />

R<br />

L<br />

r<br />

r<br />

I sd<br />

LM<br />

R<br />

L<br />

r<br />

r<br />

∫<br />

Ψ r<br />

M e<br />

I sq<br />

p<br />

b<br />

ms<br />

2<br />

L<br />

L<br />

M<br />

r<br />

M e<br />

1<br />

J<br />

∫<br />

Ω m<br />

M L<br />

Fig. 3.2. Block diagram of induction motor in<br />

d − q coordinate system<br />

The main feature of the field oriented control (FOC) method is the coordinate<br />

transformation. The current vector is measured in stationary coordinate<br />

α − β .<br />

Therefore, current components<br />

I , I<br />

β<br />

must be transformed to the rotating system<br />

sα<br />

s<br />

d − q . Similarly, the reference stator voltage vector components U<br />

s α c<br />

,<br />

U<br />

s β c<br />

, must be<br />

transformed from the system<br />

d − q to α − β . These transformations requires a rotor<br />

flux angle γ sr<br />

. Depending on calculations of this angle two different kind of field<br />

oriented control methods maybe considered. Those are <strong>Direct</strong> Field Oriented Control<br />

(DFOC) and Indirect Field Oriented Control (IFOC) methods.<br />

42


3.2. Field Oriented Control (FOC)<br />

For DFOC an estimator or observer calculates the rotor flux angle γ<br />

sr<br />

. Inputs to the<br />

estimator or observer are stator voltages and currents. An example of the DFOC system<br />

is presented in Fig. 3.3.<br />

U dc<br />

Ψ rc<br />

M ec<br />

b<br />

s<br />

1<br />

L M<br />

2 Lr<br />

1<br />

p m L Ψ<br />

M<br />

rc<br />

I sdc<br />

I sqc<br />

PI<br />

PI<br />

d − q<br />

α − β<br />

U<br />

s α c<br />

U<br />

s β c<br />

<strong>SVM</strong><br />

S A<br />

S B<br />

S C<br />

γ<br />

sr<br />

Flux<br />

Estimator<br />

U sα<br />

U sβ<br />

Voltage<br />

Calculation<br />

I sd<br />

d − q<br />

I s α<br />

2<br />

I s<br />

I sq<br />

α − β<br />

I sβ<br />

3<br />

Motor<br />

Fig. 3.3. Block diagram of the <strong>Direct</strong> Field Oriented Control (DFOC)<br />

For the IFOC rotor flux angle γ<br />

sr<br />

is obtained from reference I sdc<br />

, I<br />

sqc<br />

currents. The<br />

angular speed of the rotor flux vector speed can be calculated as follows:<br />

where<br />

Ω = Ω + p Ω<br />

(3.8)<br />

rs<br />

sl<br />

b<br />

m<br />

Ω<br />

sl<br />

is a slip angular speed. It can be calculated from (3.3d) and (3.4d).<br />

Ω<br />

sl<br />

1<br />

R<br />

r<br />

= Isqc<br />

(3.9)<br />

Isdc<br />

Lr<br />

In Fig. 3.4 a block diagram of the IFOC is shown.<br />

43


3. <strong>Vector</strong> Control Methods of Induction Motor<br />

U dc<br />

Ψ rc<br />

M ec<br />

2 Lr<br />

p m L<br />

b<br />

s<br />

1<br />

L M<br />

M<br />

1<br />

Ψ<br />

rc<br />

I sdc<br />

I sqc<br />

PI<br />

PI<br />

d − q<br />

α − β<br />

U<br />

s α c<br />

U<br />

s β c<br />

<strong>SVM</strong><br />

S A<br />

S B<br />

S C<br />

γ<br />

sr<br />

R 1<br />

r<br />

L I<br />

r<br />

sdc<br />

I sd<br />

d − q<br />

I s α<br />

2<br />

I s<br />

Ω sr<br />

∫<br />

I sq<br />

α − β<br />

I sβ<br />

3<br />

Motor<br />

Ω sl<br />

p b<br />

Ω m<br />

Fig. 3.4. Block diagram of the Indirect Field Oriented Control (IFOC)<br />

In both presented examples reference currents in rotating coordinate system I<br />

sdc<br />

, I<br />

sqc<br />

are calculated from the reference flux and torque values. Taking into consideration the<br />

equations describing IM in field oriented coordinate system (3.6) and (3.7) at steady<br />

state the formulas for the reference currents can be written as follows:<br />

I<br />

I<br />

sdc<br />

sqc<br />

1<br />

= Ψ<br />

r<br />

(3.10)<br />

L<br />

M<br />

2<br />

L<br />

1<br />

r<br />

= M<br />

ec<br />

(3.11)<br />

pbms<br />

LM<br />

Ψ<br />

rc<br />

The property of the FOC methods can be summarized as follows:<br />

• the method is based on the analogy to control of a DC motor,<br />

• FOC method does not guarantee an exact decoupling of the torque and flux<br />

control in dynamic and steady state operation,<br />

• relationship between regulated value and control variables is linear only for<br />

constant rotor flux amplitude,<br />

44


3.3. Feedback Linearization Control (FLC)<br />

• full information about motor state variable and load torque is required (the<br />

method is very sensitive to rotor time constant),<br />

• current controllers are required,<br />

• coordinate transformations are required,<br />

• a PWM algorithm is required (it guarantees constant switching frequency),<br />

• in the DFOC rotor flux estimator is required,<br />

• in the IFOC mechanical speed is required,<br />

• the stator currents are sinusoidal except of high frequency switching harmonics.<br />

3.3. Feedback Linearization Control (FLC)<br />

The transformation of the induction motor equations in the field coordinates has a<br />

good physical basis because it corresponds to the decoupled torque production in a<br />

separately excited DC motor. However, from the theoretical point of view, other types<br />

of coordinates can be selected to achieve decoupling and linearization of the induction<br />

motor equations.<br />

In [28] it is shown that a nonlinear dynamic model of IM can be considered as<br />

equivalent to two third-order decoupled linear systems. In [70] a controller based on a<br />

multiscalar motor model has been proposed. The new state variables have been chosen.<br />

In result the motor speed is fully decoupled from the rotor flux. In [82] the authors<br />

proposed a nonlinear transformation of the motor states variables, so that in the new<br />

coordinates, the speed and rotor flux amplitude are decoupled by feedback. Others<br />

proposed also modified methods based on Feedback Linearization Control like in [93,<br />

94].<br />

In the example given new quantities for control of rotor flux magnitude and<br />

mechanical speed were chosen [93]. For this purpose the induction motor equations<br />

(2.10-2.12) can be written in the following form:<br />

x& = f (x) + U s<br />

g + U g<br />

(3.12)<br />

α<br />

α<br />

sβ<br />

β<br />

where:<br />

45


3. <strong>Vector</strong> Control Methods of Induction Motor<br />

⎡−<br />

αΨ<br />

rα<br />

− pbΩ<br />

mΨ<br />

rβ<br />

+ αLM<br />

I<br />

sα<br />

⎤<br />

⎢<br />

⎥<br />

⎢<br />

pbΩ<br />

mΨ<br />

rα<br />

− αΨ<br />

rβ<br />

+ αLM<br />

I<br />

sβ<br />

⎥<br />

⎢ +<br />

− ⎥<br />

f ( x ) =<br />

αβ Ψ<br />

rα<br />

βpbΩ<br />

mΨ<br />

rβ<br />

γI<br />

sα<br />

⎢<br />

⎥<br />

(3.13)<br />

⎢ − βpbΩ<br />

mΨ<br />

rα<br />

+ αβ Ψ<br />

rβ<br />

− γI<br />

sβ<br />

⎥<br />

⎢<br />

M<br />

L<br />

⎥<br />

⎢ µ ( Ψ<br />

rα<br />

I<br />

sβ<br />

− Ψ<br />

rβ<br />

I<br />

sα<br />

) −<br />

⎣<br />

J<br />

⎥<br />

⎦<br />

T<br />

⎡ 1 ⎤<br />

gα = ⎢0 , 0,<br />

, 0,<br />

0⎥ (3.14)<br />

⎣ σL s ⎦<br />

and<br />

T<br />

⎡ 1 ⎤<br />

g<br />

β<br />

= ⎢0 , 0,<br />

0,<br />

, 0⎥ (3.15)<br />

⎣ σL s ⎦<br />

x<br />

[ Ψ Ψ , I , I , Ω ] T<br />

= (3.16)<br />

rα<br />

,<br />

rβ<br />

sα<br />

sβ<br />

R<br />

L<br />

r<br />

m<br />

r<br />

α =<br />

(3.17)<br />

L<br />

σL<br />

L<br />

M<br />

β = (3.18)<br />

s<br />

r<br />

γ<br />

2<br />

2<br />

RsLr<br />

+ Rr<br />

LM<br />

= (3.19)<br />

2<br />

σLsLr<br />

msLM<br />

µ = pb<br />

(3.20)<br />

2J<br />

Because<br />

Ω ,<br />

m<br />

, Ψ<br />

rα<br />

Ψ<br />

rβ<br />

are not dependent on<br />

sα<br />

U<br />

sβ<br />

U , it is possible to chose variable<br />

dependent on x:<br />

2 2 2<br />

1<br />

( x)<br />

= Ψ<br />

rα<br />

+ Ψ<br />

rβ<br />

= Ψ<br />

r<br />

φ (3.21)<br />

φ 2(x) = Ω m<br />

(3.22)<br />

If it is assumed that φ (x 1<br />

) , φ (x 2<br />

) are output variables, the full definition of new<br />

coordinates can be given by:<br />

z<br />

1<br />

= φ 1<br />

(x)<br />

(3.23a)<br />

z<br />

2<br />

= L f<br />

φ 1<br />

(x)<br />

(3.23b)<br />

46


3.3. Feedback Linearization Control (FLC)<br />

z<br />

3<br />

= φ 2<br />

(x)<br />

(3.23c)<br />

z<br />

4<br />

= L f<br />

φ 2<br />

(x)<br />

(3.23d)<br />

⎛Ψ<br />

rβ<br />

⎞<br />

z =<br />

⎜<br />

⎟<br />

5<br />

arctan<br />

(3.23e)<br />

⎝Ψ<br />

rα<br />

⎠<br />

It should be mentioned that the goal of the control is to obtain constant flux<br />

amplitude and to follow the reference angular speed.<br />

The fifth variable cannot be fully linearized. Additionally, it is not controllable (the<br />

fifth variable correspond to slip in the motor). Therefore, the last equation is not<br />

considered. Then the dynamics of the system are given by:<br />

2<br />

⎡&&<br />

z ⎤ ⎡ φ ⎤ ⎡U<br />

1<br />

Lf<br />

1<br />

⎢ ⎥ = ⎢ ⎥ + D<br />

2 ⎢<br />

⎣&&<br />

z ⎦ ⎢⎣<br />

Lfφ<br />

⎥⎦<br />

⎣U<br />

3<br />

2<br />

sα<br />

sβ<br />

⎤<br />

⎥<br />

⎦<br />

(3.24)<br />

where<br />

⎡ L<br />

⎤<br />

gα<br />

L<br />

fφ1<br />

Lg<br />

β<br />

Lfφ1<br />

D = ⎢<br />

⎥<br />

(3.25)<br />

⎢⎣<br />

Lgα<br />

L<br />

fφ2<br />

Lg<br />

β<br />

Lfφ2⎥⎦<br />

If φ 0 (the amplitude of flux is not zero) then det( D) ≠ 0 and it is possible to<br />

1 ≠<br />

define the linearization feedback as:<br />

⎡U<br />

⎢<br />

⎣U<br />

sβ<br />

2<br />

⎤ ⎪⎧<br />

⎡ ⎤ ⎪⎫<br />

-<br />

− Lfφ1<br />

⎡v1<br />

⎤<br />

⎥ = D ⎨⎢<br />

⎥ + ⎢ ⎥⎬<br />

(3.26)<br />

2<br />

⎦ ⎪⎩ ⎢⎣<br />

− Lfφ2⎥⎦<br />

⎣v2⎦⎪⎭<br />

sα 1<br />

Then the resulting system is described by the equations:<br />

z &<br />

1<br />

= z 2<br />

(3.27a)<br />

z &<br />

2<br />

= v 1<br />

(3.27b)<br />

z &<br />

3<br />

= z 4<br />

(3.27c)<br />

z &<br />

4<br />

= v 2<br />

(3.27d)<br />

and the final block diagram of the induction motor with the new defined control<br />

signals can be shown as in Fig. 3.5.<br />

47


3. <strong>Vector</strong> Control Methods of Induction Motor<br />

ν 1<br />

∫<br />

z 2<br />

∫<br />

2<br />

Ψ r<br />

Ψ r<br />

∫<br />

Ω m<br />

ν 2<br />

∫<br />

z 4<br />

J<br />

M e<br />

M L<br />

Fig. 3.5. Block diagram of the induction motor with new v<br />

1<br />

and v<br />

2<br />

control signals<br />

The control signals v<br />

1, v<br />

2<br />

are calculated by using linear feedback as follows:<br />

( z1<br />

− z1<br />

) k12<br />

2<br />

v =<br />

ref<br />

−<br />

(3.28)<br />

1<br />

k11<br />

z<br />

( z3<br />

− z3<br />

) k22<br />

4<br />

v =<br />

ref<br />

−<br />

(3.29)<br />

2<br />

k21<br />

z<br />

where coefficients k 11<br />

, k<br />

12<br />

, k<br />

21<br />

, k<br />

22<br />

are chosen to receive reference close loop<br />

system dynamics.<br />

An example of a FLC system for PWM inverter-fed induction motor is presented in<br />

Fig. 3.6.<br />

The property of the FLC can be summarized as follows:<br />

• it guarantees exactly decoupling of the motor speed and rotor flux control in both<br />

dynamic and steady state,<br />

• the method is implemented in a state variable control fashion and needs complex<br />

signal processing,<br />

• full information about motor state variables and load torque is required,<br />

• there are no current controllers,<br />

• a PWM vector modulator is required, what further guarantee constant switching<br />

frequency,<br />

48


3.4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control (<strong>DTC</strong>)<br />

• the stator currents are sinusoidal except of high frequency switching harmonics.<br />

U<br />

dc<br />

2<br />

Ψ rc<br />

Ω mc<br />

Flux<br />

Controller<br />

Speed<br />

Controller<br />

ν 1<br />

ν 2<br />

Control<br />

Signals<br />

Transformation<br />

U<br />

s β c<br />

U<br />

s α c<br />

<strong>Vector</strong><br />

Modulator<br />

S A<br />

S B<br />

S C<br />

Voltage<br />

Calculation<br />

z 1<br />

I sα<br />

z 2<br />

z 3<br />

z 4<br />

z 5<br />

Feedback<br />

Signals<br />

Transformation<br />

I sβ<br />

Ψˆrα<br />

Ψˆrβ<br />

Flux<br />

Estimator<br />

Uˆ s<br />

I s<br />

Motor<br />

Ω m<br />

Fig. 3.6. Block scheme of the feedback linearization control method<br />

3.4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control (<strong>DTC</strong>)<br />

3.4.1. Basics of <strong>Direct</strong> Flux and <strong>Torque</strong> Control<br />

As it was mentioned in section 3.2 in the classical vector control strategy (FOC) the<br />

torque is controlled by the stator current component I sq<br />

in accordance with equation<br />

(3.7). This equation can also be written as:<br />

where:<br />

ms<br />

LM<br />

M<br />

e<br />

= pb<br />

Ψ<br />

rI<br />

s<br />

sinδ<br />

(3.30)<br />

2 L<br />

r<br />

δ - angle between rotor flux vector and stator current vector.<br />

The formula (3.30) can be transformed into the equation:<br />

M<br />

e<br />

m<br />

L<br />

s M<br />

= pb<br />

Ψ<br />

sΨ<br />

r<br />

sin δΨ<br />

(3.31)<br />

2<br />

2 Lr<br />

Ls<br />

− LM<br />

where:<br />

δ<br />

Ψ<br />

- angle between rotor and stator flux vectors.<br />

49


3. <strong>Vector</strong> Control Methods of Induction Motor<br />

It can be noticed that the torque depends on the stator and rotor flux magnitude as<br />

well as the angleδ<br />

Ψ<br />

. The vector diagram of IM is presented in Fig. 3.7. The two angels<br />

δ and δ Ψ<br />

are also shown in Fig. 3.7. The angle δ is important in FOC algorithms,<br />

whereas δ<br />

Ψ<br />

in <strong>DTC</strong> techniques.<br />

β<br />

I s<br />

Ψ s<br />

δ<br />

γ ss<br />

δ Ψ<br />

Ψ r<br />

γ sr<br />

α<br />

Fig. 3.7. <strong>Vector</strong> diagram of induction motor<br />

From the motor voltage equation (2.10a), for the omitted voltage drop on the stator<br />

resistance, the stator flux can by expressed as:<br />

d<br />

Ψ<br />

s = U<br />

s<br />

dt<br />

(3.32)<br />

Taking into consideration the output voltage of the inverter in the above equation it<br />

can be written as:<br />

where:<br />

t<br />

∫<br />

Ψ s<br />

= Uvdt<br />

(3.33)<br />

0<br />

2 3<br />

⎧ j(<br />

v−1)<br />

π<br />

⎪<br />

U<br />

dce<br />

v = 1...6<br />

U = 3<br />

v ⎨<br />

(3.34)<br />

⎪0<br />

v = 0,7<br />

⎩<br />

Equation (3.33) describe eight voltage vectors which correspond to possible inverter<br />

states. These vectors are shown in Fig. 3.8. There are six active vectors U 1 -U 6 and two<br />

zero vectors U 0 , U 7 .<br />

50


3.4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control (<strong>DTC</strong>)<br />

Im<br />

U 3<br />

(010)<br />

U 2<br />

(110)<br />

U 4<br />

(011)<br />

U 0<br />

(000)<br />

U 1<br />

(100)<br />

U 7<br />

(111)<br />

Re<br />

U 5<br />

(001) U 6<br />

(101)<br />

Fig. 3.8. Inverter output voltage represented as space vectors<br />

It can be seen from (3.33), that the stator flux directly depends on the inverter voltage<br />

(3.34).<br />

By using one of the active voltage vectors the stator flux vector moves to the<br />

direction and sense of the voltage vector. It can be observed by simulation of six-step<br />

mode (Fig. 3.9) and PWM operation (Fig. 3.10). In Fig. 3.9 is well shown how stator<br />

flux changes direction for the cycle sequence of the active voltage vectors. Obviously,<br />

the same effect is for the PWM operation (Fig. 3.10). However, in this case the control<br />

algorithm choose correct voltage vectors, thanks to that waveform is close to be<br />

sinusoidal. In this simulation a low sampling frequency is used (0.5kHz) for better<br />

presenting the effect. A zoom part of the flux vector trajectory is shown in Fig. 3.11.<br />

In induction motor the rotor flux is slowly moving but the stator flux can be changed<br />

immediately. In direct torque control methods the angle between stator and rotor flux<br />

δ<br />

Ψ<br />

can be used as a variable of torque control (3.31). Moreover stator flux can be<br />

adjusted by stator voltage in simple way. Therefore, angle δ<br />

Ψ<br />

as well as torque can be<br />

changed thanks to the appropriate selection of voltage vector.<br />

There are the general bases of the direct flux and torque control methods. Those<br />

consideration and above equations can be used in analysis of the classical <strong>DTC</strong><br />

algorithms as well as in new proposed methods. It is also bases of the <strong>DTC</strong>-<strong>SVM</strong><br />

methods, which are presented in Chapter 4.<br />

51


3. <strong>Vector</strong> Control Methods of Induction Motor<br />

a)<br />

b)<br />

Fig. 3.9. IM under six-step mode a) voltage and stator flux waveforms, b) stator flux trajectory<br />

a)<br />

b)<br />

Fig. 3.10. IM under PWM operation a) voltage and stator flux waveforms, b) stator flux trajectory<br />

52


3.4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control (<strong>DTC</strong>)<br />

β<br />

voltage U 4<br />

applied<br />

voltage U 3<br />

applied<br />

voltage U 4<br />

applied<br />

voltage U 3<br />

applied<br />

voltage U 4<br />

applied<br />

U 3<br />

(010)<br />

U 2<br />

(110)<br />

voltage U 3<br />

applied<br />

voltage U 2<br />

applied<br />

U 4<br />

(011)<br />

U 0<br />

(000)<br />

U 1<br />

(100)<br />

voltage U 3<br />

applied<br />

U 7<br />

(111)<br />

α<br />

U 5<br />

(001) U 6<br />

(101)<br />

Fig. 3.11. Forming of the stator flux trajectory by appropriate voltage vectors selection<br />

3.4.2. Classical <strong>Direct</strong> <strong>Torque</strong> Control (<strong>DTC</strong>) <strong>–</strong> Circular Flux Path<br />

The block diagram of classical <strong>DTC</strong> proposed by I. Takahashi and T. Nogouchi [97]<br />

is presented in Fig. 3.12.<br />

Ψ<br />

sc<br />

e Ψ<br />

e M<br />

Flux<br />

Controller<br />

dΨ<br />

d M<br />

M ec<br />

U s<br />

I s<br />

<strong>Vector</strong><br />

Selection<br />

Table<br />

γ ss<br />

(N)<br />

S A<br />

S B<br />

S C<br />

U dc<br />

<strong>Torque</strong><br />

Controller<br />

Sector<br />

Detection<br />

Voltage<br />

Calculation<br />

Mˆ<br />

Ψˆ s<br />

e<br />

Ψˆ sα<br />

Ψˆsβ<br />

Flux and<br />

<strong>Torque</strong><br />

Estimator<br />

Motor<br />

Fig. 3.12. Block scheme of the direct torque control method<br />

53


3. <strong>Vector</strong> Control Methods of Induction Motor<br />

The stator flux amplitude<br />

signals which are compared with the estimated<br />

Ψ<br />

sc<br />

and the electromagnetic torque M<br />

c<br />

are the reference<br />

Ψˆ s<br />

and<br />

Mˆ e<br />

values respectively. The<br />

flux e and torque Ψ<br />

e<br />

M<br />

errors are delivered to the hysteresis controllers. The digitized<br />

output variables d Ψ<br />

,<br />

d and the stator flux position sector ( N)<br />

M<br />

γ selects the<br />

appropriate voltage vector from the switching table. Thus, the selection table generates<br />

pulses S A , S B , S C to control the power switches in the inverter.<br />

For the flux is defined two-level hysteresis controller, for the torque three-level, as it<br />

is shown in Fig. 3.13.<br />

ss<br />

a)<br />

b)<br />

d Ψ<br />

d M<br />

e M<br />

H Ψ<br />

e Ψ<br />

H M<br />

Fig. 3.13. The hysteresis controllers a) two-level, b) three-level<br />

The output signals d<br />

Ψ<br />

, d<br />

M<br />

are defined as:<br />

d<br />

Ψ<br />

=1 for e<br />

Ψ<br />

> HΨ<br />

(3.35a)<br />

d<br />

Ψ<br />

= 0 for eΨ<br />

< −HΨ<br />

(3.35b)<br />

d<br />

M<br />

=1 for e<br />

M<br />

> H<br />

M<br />

(3.36a)<br />

d = 0 for e = 0<br />

(3.36b)<br />

M<br />

M<br />

d<br />

M<br />

= −1 for eM<br />

< −H<br />

M<br />

(3.36c)<br />

In the classical <strong>DTC</strong> method the plane is divided for the six sectors (Fig. 3.14),<br />

which are defined as:<br />

⎛ π π ⎞<br />

Sector 1: γ ss<br />

∈⎜−<br />

, + ⎟ (3.37a)<br />

⎝ 6 6 ⎠<br />

⎛ π π ⎞<br />

Sector 2: γ ∈ ss ⎜ + , ⎟<br />

⎝ 6 2<br />

(3.37b)<br />

⎠<br />

⎛ π 5π<br />

⎞<br />

Sector 3: γ ss<br />

∈⎜+<br />

, + ⎟ (3.37c)<br />

⎝ 2 6 ⎠<br />

54


3.4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control (<strong>DTC</strong>)<br />

⎛ ⎞<br />

Sector 4: ∈⎜+<br />

π 5<br />

γ , −<br />

ss ⎟<br />

⎝ 6 6 ⎠<br />

(3.37d)<br />

⎛ ⎞<br />

Sector 5: ∈⎜−<br />

π<br />

γ , −<br />

ss ⎟<br />

⎝ 6 2 ⎠<br />

(3.37e)<br />

⎛ π π ⎞<br />

Sector 6: γ ss<br />

∈⎜−<br />

, − ⎟<br />

⎝ 2 6 ⎠<br />

(3.37f)<br />

Sector 3<br />

β<br />

U 3<br />

(010)<br />

Sector 2<br />

U 2<br />

(110)<br />

Sector 4<br />

U 4<br />

(011)<br />

U 0<br />

(000) U 1<br />

(100)<br />

U 7<br />

(111)<br />

α<br />

Sector 1<br />

U 5<br />

(001)<br />

U 6<br />

(101)<br />

Sector 5 Sector 6<br />

Fig. 3.14. Sectors in the classical <strong>DTC</strong> method<br />

For the stator flux vector laying in sector 1 (Fig. 3.15) in order to increase its<br />

magnitude the voltage vectors U 1 , U 2 , U 6 can be selected. Conversely, a decrease can be<br />

obtained by selecting U 3 , U 4 , U 5 . By applying one of the zero vectors U 0 or U 7 the<br />

integration in equation (3.33) is stopped. The stator flux vector is not changed.<br />

For the torque control, angle between stator and rotor flux δ Ψ<br />

is used (equation<br />

3.31). Therefore, to increase motor torque the voltage vectors U 2 , U 3 , U 4 can be selected<br />

and to decrease U 1 , U 5 , U 6 .<br />

The above considerations allow construction of the selection table as presented in<br />

Table 3.1.<br />

55


3. <strong>Vector</strong> Control Methods of Induction Motor<br />

β<br />

U 3<br />

U 2<br />

U 1<br />

U 4<br />

U 5<br />

U 6<br />

Ψ s<br />

δ Ψ<br />

Sector 1<br />

α<br />

Ψ r<br />

Fig. 3.15. Selection of the optimum voltage vectors for the stator flux vector in sector 1<br />

Table 3.1. Optimum switching table<br />

dΨ<br />

d Sector 1 Sector 2 Sector 3 Sector 4 Sector 5 Sector 6<br />

M<br />

1<br />

0<br />

1<br />

0<br />

-1<br />

1<br />

0<br />

U 2<br />

U 3<br />

U 4<br />

U 5<br />

U 6<br />

U 1<br />

U 7<br />

U 0<br />

U 7<br />

U 0<br />

U 7<br />

U 0<br />

U 6<br />

U 1<br />

U 2<br />

U 3<br />

U 4<br />

U 5<br />

U 3<br />

U 4<br />

U 5<br />

U 6<br />

U 1<br />

U 2<br />

U 7<br />

U 0<br />

U 7<br />

U 0<br />

U 7<br />

-1<br />

U 0<br />

U 4<br />

U 5<br />

U 6<br />

U 1<br />

U 2<br />

U 3<br />

The signal waveforms for steady state operation of classical <strong>DTC</strong> method are shown<br />

in Fig. 3.16.<br />

The <strong>DTC</strong> was proposed as an analog control method. The implementation of the<br />

hysteresis controller in the analog setup is easy and the control system works properly.<br />

When the hysteresis controller is implemented in a digital signal processor (DSP), its<br />

operation is quite different from that of the analog scheme [19]. The digital<br />

implementation of the hysteresis controller is also called sampled hysteresis.<br />

56


3.4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control (<strong>DTC</strong>)<br />

a)<br />

b)<br />

Fig. 3.16. Steady state operation for the classical <strong>DTC</strong> method ( f s = 40kHz)<br />

a) signals in time domain, b) stator flux trajectory<br />

In Fig. 3.17 are presented typical switching sequences of the torque hysteresis<br />

controller for the analog (Fig. 3.17a) and for the digital (Fig. 3.17b) implementation.<br />

57


3. <strong>Vector</strong> Control Methods of Induction Motor<br />

a) b)<br />

S/H<br />

M<br />

c<br />

+ H m<br />

M c<br />

M<br />

c<br />

− H m<br />

t 1<br />

t 2<br />

t 3<br />

T s<br />

T s<br />

T s<br />

Fig. 3.17. Operating of the torque hysteresis controller a) analog, b) digital<br />

In the analog implementation the torque ripple are kept exactly within the hysteresis<br />

band and the switching instants are not equally spaced. The digital system operates at<br />

fixed sampling time T s<br />

and works like analog only for high sampling frequencies<br />

f<br />

s<br />

1<br />

= .<br />

T<br />

s<br />

For the lower sapling frequency the switching instants are not when the estimated<br />

torque crosses the hysteresis band but on the sampling time. This situation is presented<br />

in Fig. 3.17b. The simulation results illustrated control system behavior at lower<br />

sampling frequency<br />

f s<br />

= 15kHz<br />

are given in Fig. 3.18. It can be seen that current and<br />

torque ripples are bigger compare to this one operate with sampling frequency<br />

f s<br />

= 40kHz (see Fig. 3.16).<br />

The influence of the torque hysteresis band for the torque error and switching<br />

frequency at different sampling frequencies is shown in Fig. 3.19 and Fig. 3.20. At low<br />

sampling frequency f s = 20kHz (Fig. 3.19) the switching frequency and torque error are<br />

not sensitive for hysteresis band. However, at the high sampling frequency f s = 80kHz<br />

(Fig. 3.20) when the hysteresis band is increased the switching frequency decreases and<br />

the torque error increases. Simulated results show that the hysteresis controllers need a<br />

high sampling frequency to obtain a proper operation.<br />

The torque and flux errors are calculated according to equations:<br />

Ψˆ<br />

s<br />

−Ψ<br />

sc<br />

ε<br />

ψ<br />

= 100%<br />

(3.38a)<br />

s<br />

Ψ<br />

sN<br />

58


3.4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control (<strong>DTC</strong>)<br />

where:<br />

Mˆ<br />

e<br />

− M<br />

ec<br />

ε<br />

M<br />

= 100%<br />

(3.38b)<br />

M<br />

eN<br />

Ψ<br />

sN<br />

- nominal stator flux, M<br />

eN<br />

- nominal torque<br />

Fig. 3.18. Steady state operation for the classical <strong>DTC</strong> method operating with lower<br />

f s = 15kHz<br />

sampling frequency ( )<br />

The average value of the flux and torque errors are calculated in a period of the<br />

fundamental frequency.<br />

59


3. <strong>Vector</strong> Control Methods of Induction Motor<br />

a)<br />

f sw [Hz]<br />

25000<br />

20000<br />

15000<br />

10000<br />

5000<br />

0<br />

4792<br />

5400<br />

2750 2208 2367 2333<br />

4567 4333 3508<br />

0,0 0,5 1,0 1,5 2,0 2,5 3,0 3,5 4,0 H m [Nm]<br />

b)<br />

ε Μ _avr [%]<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

11,06<br />

11,97 11,00 9,43 9,93 10,68 12,03<br />

9,65<br />

10,17<br />

0,0 0,5 1,0 1,5 2,0 2,5 3,0 3,5 4,0 H m [Nm]<br />

Fig. 3.19. Simulated results for classical <strong>DTC</strong> a) switching frequency and b) torque error as a function of<br />

the torque hysteresis band at sampling frequency f s = 20kHz<br />

a)<br />

f sw [Hz]<br />

25000<br />

20000<br />

15000<br />

10000<br />

5000<br />

0<br />

19750<br />

13317<br />

8233<br />

6142 5492 5450 5666<br />

7400 6666<br />

0,0 0,5 1,0 1,5 2,0 2,5 3,0 3,5 4,0 H m [Nm]<br />

b)<br />

ε Μ _avr [%]<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

2,43<br />

2,64<br />

3,06<br />

4,21<br />

5,36<br />

6,56<br />

7,77<br />

8,94<br />

10,27<br />

0,0 0,5 1,0 1,5 2,0 2,5 3,0 3,5 4,0 H m [Nm]<br />

Fig. 3.20. Simulated results for classical <strong>DTC</strong> a) switching frequency and b) torque error as a function of<br />

the torque hysteresis band at sampling frequency f s = 80kHz<br />

60


3.4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control (<strong>DTC</strong>)<br />

The classical <strong>DTC</strong> method can be characterized as follows:<br />

Advantages:<br />

• simple structure:<br />

o no coordinate transformation,<br />

o no separate voltage modulation block,<br />

o no current control loops,<br />

• very good flux and torque dynamic performance,<br />

Disadvantages:<br />

• variable switching frequency,<br />

• problems during starting and low speed operation,<br />

• high torque ripples,<br />

• flux and current distortion caused by stator flux vector sector position change<br />

• high sampling frequency is required for digital implementation.<br />

3.4.3. <strong>Direct</strong> Self Control (DSC) <strong>–</strong> Hexagon Flux Path<br />

The block diagram of the direct self control method proposed by M. Depenbrock [31,<br />

32] is presented in Fig. 3.21. This method was mainly applied in high power<br />

applications, which required fast torque dynamic and low switching frequency [96].<br />

Based on the command stator flux<br />

Ψ<br />

sC<br />

, the flux comparators generate digital variables d<br />

A<br />

,<br />

Ψ<br />

sc<br />

and the actual phase components Ψ<br />

sA<br />

, Ψ<br />

sB<br />

,<br />

d<br />

B<br />

, d<br />

C<br />

, which corresponds to<br />

active voltage vectors (U 1 <strong>–</strong> U 6 ). The hysteresis torque controller generates the signal<br />

d<br />

m<br />

, which determines zero states. For the constant flux region, the control algorithm is<br />

as follows:<br />

S<br />

A<br />

= d C<br />

,<br />

B<br />

d<br />

A<br />

S = , S<br />

C<br />

= d<br />

B<br />

for d<br />

m<br />

= 1<br />

(3.39a)<br />

S = 0 , S = 0 , S = 0 for d = 0<br />

(3.39b)<br />

A<br />

B<br />

C<br />

m<br />

61


3. <strong>Vector</strong> Control Methods of Induction Motor<br />

ψ sc<br />

Flux<br />

Comparators<br />

d A<br />

U dc<br />

S B<br />

d B<br />

S C<br />

S A<br />

M ec<br />

d m<br />

<strong>Torque</strong><br />

Controller<br />

d C<br />

Voltage<br />

Calculation<br />

ψˆ<br />

sC<br />

ψˆsB<br />

ψˆ<br />

sA<br />

ABC<br />

α − β<br />

ψˆsα<br />

ψˆsβ<br />

Mˆ<br />

e<br />

Flux and<br />

<strong>Torque</strong><br />

Estimator<br />

U s<br />

I s<br />

Motor<br />

Fig. 3.21. Block diagram of <strong>Direct</strong> Self Control method<br />

The signal waveforms for steady state operation of DSC method are shown in Fig.<br />

3.22. It can be seen that the flux trajectory is identical with that for the six-step mode<br />

(Fig. 3.9). This follows from the fact that the zero voltage vectors stop the flux vector,<br />

but do not affect its trajectory. The dynamic performances of torque control for the DSC<br />

are similar as for the classical <strong>DTC</strong>.<br />

The property of the DSC can be summarized as follows:<br />

• hexagonal trajectory of the stator flux vector for PWM operation,<br />

• block type of PWM (not sinusoidal),<br />

• non-sinusoidal current waveforms,<br />

• switching selection table is not required,<br />

• low (minimum) inverter switching frequency (depended on hysteresis torque<br />

band),<br />

• very good torque and flux control dynamics.<br />

62


3.4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control (<strong>DTC</strong>)<br />

a)<br />

b)<br />

Fig. 3.22. Steady state operation for the DSC method<br />

a) signals in time domain, b) stator flux trajectory<br />

Several solutions have been proposed to improve the conventional DSC. For<br />

instance, reduction of the current distortion has been achieved by introducing 12 stator<br />

flux sectors [110] or by processing not only the stator flux value , but also the stator flux<br />

63


3. <strong>Vector</strong> Control Methods of Induction Motor<br />

angle [109]. Also solutions based on fuzzy logic and neural networks solutions were<br />

proposed [85, 90].<br />

3.5. Summary<br />

In this chapter review of significant vector control methods of IM has been<br />

presented. The characteristic features for all control schemes were described.<br />

The FLC structure guarantees exact decoupling of the motor speed and rotor flux<br />

control in both dynamic and steady states. However, it is complicated and difficult to<br />

implement in practice. This method requires complex computation and additionally it is<br />

sensitive to changes of motor parameters. Because of these features this method was not<br />

chosen for implementation.<br />

Advantages<br />

Table 3.2 Comparison of control methods<br />

FOC <strong>DTC</strong> <strong>DTC</strong>-<strong>SVM</strong><br />

‣ Modulator<br />

‣ Constant switching<br />

frequency<br />

‣ Unipolar inverter<br />

output voltage<br />

‣ Low switching<br />

losses<br />

‣ Low sampling<br />

frequency<br />

‣ Current control<br />

loops<br />

Disadvantages • Coordinate<br />

transformation<br />

• A lot of control<br />

loops<br />

• Control structure<br />

depended on rotor<br />

parameters<br />

Structure<br />

independent on<br />

rotor parameters,<br />

universal for IM<br />

and PMSM<br />

Simple<br />

implementation of<br />

sensorless<br />

operation<br />

No coordinate<br />

transformation<br />

No current control<br />

loops<br />

• No modulator<br />

• Bipolar inverter<br />

output voltage<br />

• Variable switching<br />

frequency<br />

• High switching<br />

losses<br />

• High sampling<br />

frequency<br />

Structure<br />

independent on<br />

rotor parameters,<br />

universal for IM<br />

and PMSM<br />

Simple<br />

implementation of<br />

sensorless<br />

operation<br />

No coordinate<br />

transformation<br />

No current control<br />

loops<br />

‣ Modulator<br />

‣ Constant switching<br />

frequency<br />

‣ Unipolar inverter<br />

output voltage<br />

‣ Low switching<br />

losses<br />

‣ Low sampling<br />

frequency<br />

Due to above mentioned facts the FOC and <strong>DTC</strong> methods were considered next.<br />

Analysis of advantages and disadvantages of FOC and <strong>DTC</strong> methods resulted in a<br />

search for method which will eliminate disadvantages and keep advantages of those<br />

64


3.5. Summary<br />

methods. Table 3.2 summarizes features of analyzed control methods. It can be seen a<br />

combination of <strong>DTC</strong> and FOC leads to the direct torque control with space vector<br />

modulation (<strong>DTC</strong>-<strong>SVM</strong>) method which is an effect of this search. In Table 3.2 also<br />

characteristic performance of <strong>DTC</strong>-<strong>SVM</strong> was given.<br />

The disadvantages of classical <strong>DTC</strong> are caused by hysteresis controllers and<br />

switching table used in a structure. Therefore, new <strong>DTC</strong>-<strong>SVM</strong> method replaces<br />

switching table by space vector modulator and linear PI controllers are used like in the<br />

FOC scheme. However, the current control loops are eliminated. The <strong>DTC</strong>-<strong>SVM</strong><br />

methods are widely discussed in the Chapter 4 where a detailed description of those<br />

features can be found.<br />

65


4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control with <strong>Space</strong> <strong>Vector</strong><br />

Modulation (<strong>DTC</strong>-<strong>SVM</strong>)<br />

4.1. Introduction<br />

<strong>Direct</strong> flux and torque control with space vector modulation (<strong>DTC</strong>-<strong>SVM</strong>) schemes<br />

are proposed in order to improve the classical <strong>DTC</strong>. The <strong>DTC</strong>-<strong>SVM</strong> strategies operate<br />

at a constant switching frequency. In the control structures, space vector modulation<br />

(<strong>SVM</strong>) algorithm is used. The type of <strong>DTC</strong>-<strong>SVM</strong> strategy depends on the applied flux<br />

and torque control algorithm. Basically, the controllers calculate the required stator<br />

voltage vector and then it is realized by space vector modulation technique.<br />

In the <strong>DTC</strong>-<strong>SVM</strong> methods several classes have evolved:<br />

• schemes with PI controllers [111],<br />

• schemes with predictive/dead-beat [74],<br />

• schemes based on fuzzy logic and/or neural networks [40],<br />

• variable-structure control (VSC) [72, 73, 112].<br />

Different structures of <strong>DTC</strong>-<strong>SVM</strong> methods are presented in the next section. For<br />

each of the control structures, different controller design methods are proposed.<br />

The classical <strong>DTC</strong> algorithm is based on the instantaneous values and directly<br />

calculated the digital control signals for the inverter. The control algorithm in <strong>DTC</strong>-<br />

<strong>SVM</strong> methods are based on averaged values whereas the switching signals for the<br />

inverter are calculated by space vector modulator. This is main difference between<br />

classical <strong>DTC</strong> and <strong>DTC</strong>-<strong>SVM</strong> control methods.<br />

4.2. Structures of <strong>DTC</strong>-<strong>SVM</strong> <strong>–</strong> Review<br />

4.2.1. <strong>DTC</strong>-<strong>SVM</strong> Scheme with Closed <strong>–</strong> Loop Flux Control<br />

In the control structure of Fig. 4.1 the rotor flux is assumed as a reference [24]. The<br />

reference stator flux components defined in the rotor flux coordinates<br />

calculated from the following equations:<br />

Ψ<br />

sdc<br />

, Ψ<br />

sqc<br />

can be


4.2. Structures of <strong>DTC</strong>-<strong>SVM</strong> <strong>–</strong> Review<br />

L ⎛<br />

⎞<br />

s<br />

Lr<br />

dΨ<br />

rc<br />

Ψ =<br />

⎜Ψ<br />

+<br />

⎟<br />

sdc rc<br />

σ (4.1a)<br />

LM<br />

⎝ Rr<br />

dt ⎠<br />

Ψ<br />

sqc<br />

2<br />

L<br />

L<br />

M<br />

r<br />

ec<br />

= σ<br />

s<br />

(4.1b)<br />

pbms<br />

LM<br />

Ψ<br />

rc<br />

Formulas (4.1) can be derived from the equations (3.3), (3.4) and (3.7). The<br />

equations (3.3), (3.4) and (3.7) describe the motor model in the rotor flux coordinate<br />

system<br />

d − q .<br />

The amplitude of the reference stator flux, using equations (4.1) can by expressed as:<br />

2<br />

2<br />

⎛ Ls<br />

⎞ ⎛ 2 ⎞ 2 ⎛ L ⎞<br />

r<br />

M<br />

ec<br />

Ψ =<br />

⎜<br />

⎟ +<br />

⎜<br />

⎟ ( )<br />

⎜<br />

⎟<br />

sc<br />

Ψ<br />

rc<br />

σ Ls<br />

(4.2)<br />

⎝ LM<br />

⎠ ⎝ pbms<br />

⎠ ⎝ LM<br />

Ψ<br />

rc ⎠<br />

The commanded value of stator flux<br />

2<br />

Ψ<br />

sdc<br />

, Ψ<br />

sqc<br />

after transformation to stationary<br />

coordinate system α − β are compared with the estimated values Ψˆ sα<br />

, Ψˆ sβ<br />

.<br />

Ψ rc<br />

M ec<br />

Egs (4.1)<br />

Ψ sdc<br />

Ψ sqc<br />

d − q<br />

α − β<br />

Ψ sc<br />

∆Ψ s<br />

1<br />

T s<br />

U sc<br />

<strong>SVM</strong><br />

S A<br />

S B<br />

S C<br />

γˆsr<br />

R s<br />

Rotor<br />

Flux<br />

Estimator<br />

Ψˆ<br />

s<br />

Stator<br />

Flux<br />

Estimator<br />

U s<br />

I s<br />

Voltage<br />

Calculation<br />

α − β<br />

U dc<br />

I A<br />

ABC<br />

I B<br />

Fig. 4.1. <strong>DTC</strong>-<strong>SVM</strong> scheme with closed flux control<br />

The reference voltage vector depends on the increment stator flux<br />

drop on the stator winding resistance<br />

U<br />

sc<br />

s<br />

s<br />

T + R<br />

s<br />

s<br />

R<br />

s<br />

:<br />

= ∆Ψ<br />

I<br />

(4.3)<br />

∆Ψ<br />

s<br />

and voltage<br />

In this <strong>DTC</strong>-<strong>SVM</strong> structure the rotor flux magnitude is regulated. Thanks of them<br />

increase the torque overload capability is possible [19, 24]. However, the drawback of<br />

67


4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control with <strong>Space</strong> <strong>Vector</strong> Modulation (<strong>DTC</strong>-<strong>SVM</strong>)<br />

this algorithm is that it requires all the motor parameters and moreover it is very<br />

sensitive to their variation.<br />

4.2.2. <strong>DTC</strong>-<strong>SVM</strong> Scheme with Closed <strong>–</strong> Loop <strong>Torque</strong> Control<br />

The method with close-loop torque control was originally proposed for the<br />

permanent magnet synchronous motor (PMSM) [35, 36, 37]. However, the <strong>DTC</strong> basics<br />

for both IM and PMSM are identical and therefore the method can also be used for the<br />

IM [126]. The block scheme of the control structure <strong>DTC</strong>-<strong>SVM</strong> with close-loop torque<br />

control is presented in Fig. 4.2.<br />

Ψ sc<br />

M ec<br />

<strong>Torque</strong><br />

Controller<br />

∆<br />

PI<br />

δ ψ<br />

Eg. (4.4)<br />

Ψsc<br />

∆Ψ s<br />

1<br />

T s<br />

U sc<br />

<strong>SVM</strong><br />

S A<br />

S B<br />

S C<br />

Mˆ<br />

e<br />

γˆss<br />

Ψˆ<br />

s<br />

Flux and<br />

<strong>Torque</strong><br />

Estimator<br />

R s<br />

U s<br />

I s<br />

Voltage<br />

Calculation<br />

α − β<br />

U dc<br />

I A<br />

ABC<br />

I B<br />

Fig. 4.2. <strong>DTC</strong>-<strong>SVM</strong> scheme with closed-loop torque control<br />

For the torque regulation a PI controller is applied. Output of this PI controller is an<br />

increment of torque angle ∆δ<br />

Ψ<br />

(Fig. 4.3). In this way the torque is controlled by<br />

changing the angle between stator and rotor fluxes according to the basics of <strong>DTC</strong> (see<br />

section 3.4.2).<br />

The reference stator flux vector is calculated as follows:<br />

( ˆ γ + ∆δ<br />

)<br />

j ss Ψ<br />

Ψ = Ψ sc<br />

e<br />

(4.4)<br />

sc<br />

Next, reference stator flux vector is compared with the estimated value. The error of<br />

the flux<br />

equation (4.3).<br />

∆Ψ<br />

s<br />

is used, for calculation of the reference voltage vector, according to the<br />

68


4.2. Structures of <strong>DTC</strong>-<strong>SVM</strong> <strong>–</strong> Review<br />

β<br />

∆δΨ<br />

Ψ sc<br />

Ψˆ<br />

s<br />

γˆss<br />

δˆΨ<br />

Ψˆ<br />

r<br />

γˆsr<br />

α<br />

Fig. 4.3. <strong>Vector</strong> diagram<br />

The presented method has simple structure and only one PI torque controller. It<br />

makes the tuning procedure easier. The flux is adjusted in open-loop fashion.<br />

4.2.3. <strong>DTC</strong>-<strong>SVM</strong> Scheme with Close <strong>–</strong> Loop <strong>Torque</strong> and Flux Control<br />

Operating in Polar Coordinates<br />

When both torque and flux magnitudes are controlled in a closed-loop way, the<br />

strategies provide further improvement. The method operating in polar coordinates is<br />

shown in Fig. 4.4 [49].<br />

Flux<br />

Controller<br />

Ψ sc<br />

M ec<br />

P<br />

PI<br />

∆γ<br />

sd<br />

k Ψ<br />

∆γ<br />

s<br />

Eg. (4.7)<br />

∆Ψ s<br />

1<br />

T s<br />

U sc<br />

<strong>SVM</strong><br />

S A<br />

S B<br />

S C<br />

<strong>Torque</strong><br />

Controller<br />

∆γ ss<br />

γˆss<br />

R s<br />

Ψˆ s<br />

Mˆ<br />

e<br />

Flux and<br />

<strong>Torque</strong><br />

Estimator<br />

U s<br />

Voltage<br />

Calculation<br />

U dc<br />

I s<br />

α − β<br />

I A<br />

ABC<br />

I B<br />

Fig. 4.4. <strong>DTC</strong>-<strong>SVM</strong> scheme operated in stator flux polar coordinates<br />

The error of the stator flux vector<br />

of the flux and torque controllers as follows:<br />

∆Ψ<br />

s<br />

is calculated from the outputs k Ψ<br />

and ∆γ<br />

s<br />

69


4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control with <strong>Space</strong> <strong>Vector</strong> Modulation (<strong>DTC</strong>-<strong>SVM</strong>)<br />

∆Ψ<br />

s<br />

( k) = Ψ ( k) − Ψ ( k −1)<br />

s<br />

s<br />

j∆γ<br />

s ( k )<br />

([ 1+<br />

k ( k)<br />

] ⋅e<br />

−1) ⋅ ( −1)<br />

= k<br />

Ψ<br />

Ψ s<br />

(4.5)<br />

With the approximation<br />

e<br />

j∆γ<br />

s<br />

( k )<br />

+ j∆γ<br />

( k)<br />

≅ 1 (4.6)<br />

The equation (4.5) can be written in the form<br />

s<br />

s<br />

( k) = [ k ( k) + j∆ ( k)<br />

] ⋅ Ψ ( k −1)<br />

∆Ψ γ (4.7)<br />

Ψ<br />

s<br />

s<br />

The commanded stator voltage vector is calculated according to equation (4.3). To<br />

improve the dynamic performance of the torque control, the angle increment<br />

composed of two parts: the dynamic part<br />

the stationary part ∆γ<br />

ss<br />

generated by a feedforward loop.<br />

∆γ<br />

s<br />

is<br />

∆γ<br />

sd<br />

delivered by the torque controller and<br />

4.2.4. <strong>DTC</strong>-<strong>SVM</strong> Scheme with Close <strong>–</strong> Loop <strong>Torque</strong> and Flux Control<br />

in Stator Flux Coordinates<br />

A block diagram of the method with close-loop torque and flux control in stator flux<br />

coordinate system [111] is presented in Fig. 4.5. The output of the PI flux and torque<br />

controllers can be interpreted as the reference stator voltage components<br />

the stator flux oriented coordinates ( x − y ).<br />

U<br />

sxc<br />

, U<br />

syc<br />

in<br />

Flux<br />

Controller<br />

Ψ sc<br />

PI<br />

U sxc<br />

x − y<br />

S A<br />

M ec<br />

PI<br />

<strong>Torque</strong><br />

Controller<br />

U syc<br />

Ψˆ s<br />

Mˆ<br />

e<br />

α − β<br />

γˆss<br />

Flux and<br />

<strong>Torque</strong><br />

Estimator<br />

U sc<br />

U s<br />

I s<br />

<strong>SVM</strong><br />

Voltage<br />

Calculation<br />

α − β<br />

S B<br />

S C<br />

U dc<br />

I A<br />

ABC<br />

I B<br />

Fig. 4.5. <strong>DTC</strong>-<strong>SVM</strong> scheme operated in stator flux cartesian coordinates<br />

70


4.3. Analysis and Controller Design for <strong>DTC</strong>-<strong>SVM</strong> Method<br />

These dc voltage commands are then transformed into stationary frame ( α − β ), the<br />

commanded values U<br />

s α c<br />

, U<br />

s β c<br />

are delivered to <strong>SVM</strong>.<br />

4.2.5. Conclusions from Review of the <strong>DTC</strong>-<strong>SVM</strong> Structures<br />

In the three first presented structures (Fig. 4.1, Fig. 4.2 and Fig. 4.4) the calculation<br />

of reference voltage vector is based on demanded<br />

∆Ψ<br />

s<br />

according to equation (4.3).<br />

This differentiation algorithm is very sensitive to disturbances. In case of errors in the<br />

feedback signals the differentiation algorithm may not be stable. This is very serious<br />

drawback of these methods.<br />

The methods presented in Fig. 4.1 and Fig. 4.2 do not have close-loop flux control.<br />

In these methods stator flux magnitude is only adjusted.<br />

The last presented method (Fig. 4.5) eliminates problems with differentiation<br />

algorithm. Moreover, this method controls torque and flux in close-loop fashion.<br />

Therefore, this scheme will be selected for experimental realization. In the next subsection<br />

controller design for flux and torque closed loops will be discussed.<br />

4.3. Analysis and Controller Design for <strong>DTC</strong>-<strong>SVM</strong> Method with Close <strong>–</strong> Loop<br />

<strong>Torque</strong> and Flux Control in Stator Flux Coordinates<br />

The compete set of motor model equations can be written in stator flux coordinate<br />

system<br />

x − y . This system of coordinates x − y rotates with the stator flux angular<br />

speed<br />

Ω = Ω . This angular speed is defined as follows:<br />

K<br />

ss<br />

Ω<br />

ss<br />

dγ<br />

ss<br />

= (4.8)<br />

dt<br />

where: γ<br />

ss<br />

is a stator flux vector angle.<br />

The complex space vector can be resolved into components x and y .<br />

U = U + jU<br />

(4.9a)<br />

sK<br />

sx<br />

sy<br />

I = I + j I<br />

s K sx sy<br />

,<br />

rK<br />

rx ry<br />

I = I + jI<br />

(4.9b)<br />

71


4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control with <strong>Space</strong> <strong>Vector</strong> Modulation (<strong>DTC</strong>-<strong>SVM</strong>)<br />

Ψ<br />

sK<br />

= Ψ<br />

sx<br />

= Ψ<br />

s<br />

,<br />

rK<br />

= Ψ<br />

rx<br />

+ jΨ<br />

ry<br />

Ψ (4.9c)<br />

The motor model equations (2.10-2.12) in<br />

U<br />

x − y coordinate system can be written as:<br />

dΨ<br />

s<br />

= RsI<br />

sx<br />

(4.10a)<br />

dt<br />

sx<br />

+<br />

U = R I + Ω Ψ<br />

(4.10b)<br />

sy<br />

s<br />

sy<br />

ss<br />

s<br />

dΨ<br />

dt<br />

( p Ω − Ω )<br />

rx<br />

0 = Rr<br />

I<br />

rx<br />

+ + Ψ<br />

ry b m ss<br />

(4.11a)<br />

dΨ<br />

ry<br />

0 = Rr<br />

I<br />

ry<br />

+ + Ψ<br />

rx<br />

( Ωss<br />

− pbΩm<br />

)<br />

(4.11b)<br />

dt<br />

Ψ = L I + L I<br />

(4.12a)<br />

s<br />

s<br />

s<br />

sy<br />

sx<br />

M<br />

M<br />

ry<br />

rx<br />

0 = L I + L I<br />

(4.12b)<br />

Ψ = L I + L I<br />

(4.12c)<br />

rx<br />

ry<br />

r<br />

r<br />

rx<br />

ry<br />

M<br />

M<br />

sx<br />

Ψ = L I + L I<br />

(4.12d)<br />

sy<br />

dΩ<br />

dt<br />

m<br />

1 ⎡ ms<br />

⎤<br />

= ⎢ pb<br />

Ψ<br />

sI<br />

sy<br />

− M<br />

L<br />

J<br />

⎥<br />

⎣ 2 ⎦<br />

(4.13)<br />

The electromagnetic torque can be expressed by the following formula:<br />

M<br />

e<br />

m<br />

2<br />

s<br />

= pb<br />

Ψ<br />

sI<br />

sy<br />

(4.14)<br />

Based on the equations (4.10-4.14) the block diagram of induction motor can be<br />

constructed (Fig. 4.6).<br />

The block scheme presented in Fig. 4.6 is a full model of an induction motor. As can<br />

be seen, this model is quite complicated and therefore difficult to analyze. However,<br />

taking into consideration the stator voltage equations (4.10) and torque equation (4.14),<br />

the motor can be described as follows:<br />

dΨ<br />

dt<br />

s<br />

= U − R I<br />

(4.15)<br />

sx<br />

s<br />

sx<br />

M<br />

e<br />

=<br />

1<br />

R<br />

s<br />

p<br />

b<br />

m<br />

2<br />

s<br />

Ψ<br />

s<br />

( U − Ω Ψ )<br />

sy<br />

ss<br />

s<br />

(4.16)<br />

72


4.3. Analysis and Controller Design for <strong>DTC</strong>-<strong>SVM</strong> Method<br />

M L<br />

R s<br />

I sx<br />

s<br />

r<br />

1<br />

2<br />

m<br />

L L − L<br />

U sx<br />

∫<br />

Ψ s<br />

L r<br />

Ψ s<br />

L<br />

L<br />

M<br />

s<br />

ms<br />

pb<br />

2<br />

M e<br />

1<br />

J<br />

∫<br />

Ω m<br />

U sy<br />

Ω<br />

÷<br />

ss<br />

LM<br />

I sy<br />

R s<br />

R r<br />

I rx<br />

1<br />

σL r<br />

LM<br />

L L − L<br />

s<br />

r<br />

2<br />

m<br />

∫<br />

Ψ rx<br />

p b<br />

∫<br />

R r<br />

Ψ ry<br />

I ry<br />

1<br />

σL r<br />

Fig. 4.6. Complete block diagram of an induction motor in the stator flux oriented coordinates<br />

x − y<br />

The block diagram of induction motor based on equations (4.15) and (4.16) is shown<br />

in Fig. 4.7.<br />

R s I sx<br />

U sx<br />

∫<br />

Ψ s<br />

Ω ss<br />

U<br />

sy<br />

p<br />

b<br />

ms<br />

2<br />

1<br />

R<br />

s<br />

M<br />

e<br />

Fig. 4.7. Simplified (rotor equation omitted) induction motor block diagram in the stator flux oriented<br />

coordinates x − y<br />

73


4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control with <strong>Space</strong> <strong>Vector</strong> Modulation (<strong>DTC</strong>-<strong>SVM</strong>)<br />

Different control structures based on the above induction motor model are proposed<br />

in literature [73, 111, 112]. One of them is a method with two PI controllers [111],<br />

which is presented in Fig. 4.5.<br />

Considering a simple model of IM (Fig. 4.7), Fig. 4.8 shows the flux and torque<br />

control loops for the method shown in Fig. 4.5. In Fig. 4.8 the dashed line represents the<br />

IM model.<br />

R s I sx<br />

Ψ sc<br />

PI<br />

U sx<br />

∫<br />

Ψ s<br />

Ω ss<br />

M ec<br />

PI<br />

U sy<br />

ms<br />

pb<br />

2<br />

1<br />

R<br />

s<br />

M e<br />

Fig. 4.8. Control loops with two PI controllers and simplified IM model of Fig. 4.7<br />

In the next parts two approaches to a controller design will be presented and<br />

compared. Both of them are based o the assumption that control loop can be considered<br />

as quasi-continuous (fast sampling). The first method is based on simple symmetric<br />

criterion [66], the second one uses root locus technique [34, 86].<br />

PI Controllers<br />

The transfer function of PI controllers is given as follows:<br />

where:<br />

G<br />

R<br />

() s<br />

( s)<br />

⎛ 1 ⎞ +<br />

= K<br />

p<br />

= K<br />

p<br />

() s<br />

⎜ +<br />

sT<br />

⎟<br />

i<br />

sTi<br />

U<br />

1<br />

=<br />

E ⎝ ⎠<br />

sT<br />

i<br />

1 (4.17)<br />

K<br />

p<br />

- controller gain, T<br />

i<br />

- controller integrating time.<br />

The PI controller scheme is presented in Fig. 4.9.<br />

74


4.3. Analysis and Controller Design for <strong>DTC</strong>-<strong>SVM</strong> Method<br />

E()<br />

s<br />

K p<br />

1<br />

U ( s)<br />

T i<br />

1s<br />

Fig. 4.9. Block diagram of PI controller<br />

Presented above model of the controller was used in <strong>DTC</strong>-<strong>SVM</strong> control method with<br />

two PI controllers.<br />

4.3.1. <strong>Torque</strong> and Flux Controllers Design <strong>–</strong> Symmetry Criterion Method<br />

Flux Controller Design<br />

The block diagram of the flux control loop is shown in Fig. 4.10. This control loops<br />

is based on the model presented in Fig. 4.8. The voltage drop on the stator resistance is<br />

neglected. In the stator flux control loop the inverter delay is taken into consideration.<br />

Ψ sc<br />

PI<br />

U sx<br />

1<br />

1+ sT<br />

1<br />

1 Ψ<br />

s<br />

s<br />

Fig. 4.10. Stator flux magnitude control loops<br />

For the flux controller parameter design the symmetry criterion can by applied [66].<br />

In accordance with the symmetry criterion the plant transfer function can be written as:<br />

G<br />

() s<br />

=<br />

K e<br />

sT<br />

−sτ<br />

0<br />

c<br />

2<br />

1<br />

1<br />

( + sT )<br />

(4.18)<br />

where: K<br />

c<br />

= 1 is the inverter gain, τ 0<br />

is dead time of the inverter ( τ<br />

0<br />

= 0 ideal<br />

converter), T = 2<br />

1 , and T<br />

1<br />

= Ts<br />

is a sum of small time constants, which includes<br />

statistical delay of the PWM generation and signal processing delay. The optimal<br />

controller parameters can be calculated as:<br />

75


4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control with <strong>Space</strong> <strong>Vector</strong> Modulation (<strong>DTC</strong>-<strong>SVM</strong>)<br />

K<br />

pΨ<br />

=<br />

K<br />

T<br />

T + τ<br />

2<br />

=<br />

2<br />

1 0<br />

1<br />

2T<br />

c<br />

( )<br />

s<br />

(4.19)<br />

iΨ<br />

( T + ) Ts<br />

T = τ 4<br />

(4.20)<br />

4<br />

1 0<br />

=<br />

In Table 4.1 are shown flux controller parameters calculated according to equations<br />

(4.19) and (4.20). The considered range of the sampling frequency was form 2.5kHz to<br />

10kHz. In Table 4.1 are also shown parameters of the step flux response obtained in<br />

simulation, t nΨ<br />

- time when the actual flux is first time equal reference value and p<br />

Ψ<br />

-<br />

overshoot. The results of simulation are presented in Fig. 4.11.<br />

Table 4.1. Flux controller parameters calculated according to symmetric optimum criterion<br />

f s K p Ψ T i Ψ t n Ψ p Ψ<br />

10.0 kHz 5000 0.00040 0.00150 s 1.60 %<br />

5.0 kHz 2500 0.00080 0.00180 s 2.37 %<br />

2.5 kHz 1250 0.00160 0.00200 s 9.33 %<br />

a)<br />

b)<br />

c)<br />

Fig. 4.11. Simulated flux response for controller parameters calculated according to symmetric optimum<br />

criterion at different sampling frequency a) f s = 10kHz<br />

, b) f s = 5kHz<br />

, c) f s = 2. 5kHz<br />

76


4.3. Analysis and Controller Design for <strong>DTC</strong>-<strong>SVM</strong> Method<br />

Presented in Fig. 4.11 simulation results confirm proper operation of the flux<br />

controller for the different sampling frequency. The symmetric optimum criterion can<br />

be apply to tune flux controller in analyzed <strong>DTC</strong>-<strong>SVM</strong> structure.<br />

<strong>Torque</strong> Controller Design<br />

The block diagram of the torque control loop is shown in Fig. 4.12. The same like for<br />

flux this control loops is based on the model presented in Fig. 4.8. However, coupling<br />

between torque and flux is omitted. Because of that very simple model is obtained and<br />

for this model any criterion cannot be applied.<br />

M ec<br />

PI<br />

U sy<br />

1<br />

1+ sT s<br />

ms<br />

pb<br />

2<br />

1<br />

Ψ<br />

R<br />

s<br />

s<br />

M e<br />

Fig. 4.12. Block diagram of the torque control loops<br />

In this case the simple (practical) way to design torque controller can be used.<br />

Starting from the initial values e.g. K<br />

pM<br />

= 1, TiM<br />

= 4Ts<br />

the proportional gain K<br />

pM<br />

is<br />

increasing cyclically as it is shown in Fig. 4.13. From these oscillograms the best value<br />

of<br />

K<br />

pM<br />

for the fast torque response without oscillation and small overshoot can be<br />

selected. In Fig. 4.13 the chosen simulation results for 5kHz and 10kHz sampling<br />

frequencies are shown. For the sampling frequency 5kHz the best value of proportional<br />

gain is K = 17 and for 10kHz K = 24 .<br />

pM<br />

pM<br />

The finally obtained in this way parameters of the torque controller are shown in<br />

Table 4.2. There are also shown parameters of the step torque response obtained in<br />

simulation, t<br />

nM<br />

- time when the actual torque achieves first time reference value and<br />

p<br />

M<br />

- overshoot.<br />

Table 4.2. <strong>Torque</strong> controller parameters<br />

f s K pM T iM t nM p Μ<br />

10.0 kHz 24 0.0004 0.0007 s 8.39 %<br />

5.0 kHz 17 0.0008 0.0008 s 18.53 %<br />

77


4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control with <strong>Space</strong> <strong>Vector</strong> Modulation (<strong>DTC</strong>-<strong>SVM</strong>)<br />

a)<br />

b)<br />

K pM<br />

= 4<br />

K pM<br />

= 4<br />

K = 10<br />

= 10<br />

pM<br />

K pM<br />

K pM<br />

= 17<br />

K pM<br />

= 24<br />

Fig. 4.13. <strong>Torque</strong> response for selected controller gain K pM values, at different sampling frequency<br />

a) f s 5kHz<br />

T iM = 800µs<br />

, b) f s 10kHz<br />

T iM = 400µs<br />

= ( )<br />

= ( )<br />

4.3.2. <strong>Torque</strong> and Flux Controllers Design <strong>–</strong> Root Locus Method<br />

A root-locus analysis is used for tuning the flux and torque controllers. This<br />

technique shows how the changes in the system’s open-loop characteristics influences<br />

the closed-loop dynamic characteristics. This method allows to plot the locus of the<br />

closed-loop roots in s-plane as an open-loop parameters varies, thus producing a root<br />

locus.<br />

The damping factor, overshoot and settling time [106] limit the allowable area of<br />

existence of the close-loop roots. The border of each of these parameters can be<br />

represented in s-plane as a straight line.<br />

The allowable area of existence for the close-loop roots limited by dumping and<br />

settling time is shown in Fig. 4.14.<br />

78


4.3. Analysis and Controller Design for <strong>DTC</strong>-<strong>SVM</strong> Method<br />

damping<br />

settling<br />

time<br />

Im<br />

α<br />

α<br />

Re<br />

damping<br />

Fig. 4.14. Allowable area of existence for the close-loop roots in s-plane<br />

To plot and analyze the locus of the root in s-plane SISO Design Tool Control<br />

System Toolbox v 5.0 the MathWorks, Inc. was used [84].<br />

The SISO Design Tool is a Graphical User Interface (GUI) that allows to analyze<br />

and tune the Single Input Single Output (SISO) feedback control systems. Using the<br />

SISO Design Tool, it is possible to graphically tune the gains and dynamics of the<br />

compensator (C) and prefilter (F), using a mix of root locus and loop shaping<br />

techniques. The example window of the SISO Design Tool is shown in Fig. 4.15. In the<br />

upper right area of the window, the currently tested control structure is displayed. More<br />

on the left the values of the compensator parameters are visible, and below them the<br />

resulting root-locus of the system is shown. In the root locus diagram, two lines<br />

corresponding to the inserted values of settling time and the overshoot are also visible.<br />

79


4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control with <strong>Space</strong> <strong>Vector</strong> Modulation (<strong>DTC</strong>-<strong>SVM</strong>)<br />

Fig. 4.15. SISO Design Tool<br />

Configuration of the system structure is possible by importing transfer functions of<br />

each block from the workspace. This is shown in Fig. 4.16.<br />

Fig. 4.16. Import system data<br />

80


4.3. Analysis and Controller Design for <strong>DTC</strong>-<strong>SVM</strong> Method<br />

The plant (G) is a transfer function of the motor torque or flux and the compensator<br />

(C) is a transfer function of the PI controller.<br />

In the cases of flux and torque control, the open-loop consists of a PI controller and<br />

plant transfer function, according to scheme (Fig. 4.8). The plant transfer function for<br />

the flux and the torque are calculated separately based on the motor model equation in<br />

the stator flux reference frame (4.10 - 4.12).<br />

Flux Controller Design<br />

Based on the motor model equations (4.10 - 4.12), the following equation can be<br />

obtained:<br />

⎛<br />

⎜ Rr<br />

Ls<br />

+ σ LsLr<br />

⎝<br />

d<br />

dt<br />

⎞<br />

⎟U<br />

⎠<br />

sx<br />

⎡<br />

= ⎢RsR<br />

⎢⎣<br />

r<br />

+<br />

d<br />

dt<br />

2<br />

⎛ d ⎞ ⎤<br />

⎝ dt ⎠ ⎥⎦<br />

( Rr<br />

Ls<br />

+ RsLr<br />

) + σLsLr<br />

⎜ ⎟ ⎥Ψ<br />

s<br />

s<br />

sy<br />

s<br />

r<br />

( Ω − p Ω )<br />

+ R I σ L L<br />

(4.21)<br />

ss<br />

b<br />

m<br />

L 2<br />

M<br />

where: σ = 1−<br />

L L<br />

s<br />

r<br />

Under the assumption that the last term in the equation (4.21) is very small:<br />

s<br />

sy<br />

s<br />

r<br />

( Ω − p Ω ) ≈ 0<br />

R I σ L L<br />

(4.22)<br />

the equation (4.21) becomes:<br />

ss<br />

b<br />

m<br />

2<br />

⎛<br />

d ⎞ ⎡ d<br />

⎛ d ⎞ ⎤<br />

⎜ Rr<br />

Ls<br />

+ σ LsLr<br />

⎟U<br />

sx<br />

= ⎢RsRr<br />

+ ( Rr<br />

Ls<br />

+ RsLr<br />

) + σLsLr<br />

⎜ ⎟ ⎥Ψ<br />

s<br />

(4.23)<br />

⎝<br />

dt ⎠ ⎢⎣<br />

dt<br />

⎝ dt ⎠ ⎥⎦<br />

Based on the equations (4.23) the open-loop flux transfer function can be obtained as<br />

follows:<br />

G<br />

Ψ<br />

() s<br />

Ψ<br />

A<br />

+ s<br />

s<br />

Ψ<br />

= =<br />

(4.24)<br />

2<br />

U<br />

sx<br />

s + BΨ<br />

s + CΨ<br />

where:<br />

A<br />

Ψ<br />

R<br />

σL<br />

r<br />

= ;<br />

r<br />

B<br />

Ψ<br />

R<br />

L<br />

r s s r<br />

= ;<br />

σL<br />

+ R<br />

s<br />

L<br />

r<br />

L<br />

C<br />

Ψ<br />

Rs<br />

Rr<br />

=<br />

σL<br />

L<br />

s<br />

r<br />

The flux control loop is shown in Fig. 4.17, where G RΨ<br />

( s)<br />

is a transfer function of<br />

the PI controller given by equation (4.17).<br />

81


4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control with <strong>Space</strong> <strong>Vector</strong> Modulation (<strong>DTC</strong>-<strong>SVM</strong>)<br />

Ψ sc<br />

G RΨ<br />

() s<br />

U sx<br />

G Ψ<br />

( s)<br />

Ψ<br />

s<br />

Fig. 4.17. Flux control loop<br />

The input data to the SISO Design Tool are obtained based on equations (4.17) and<br />

(4.24). The parameter values are calculated for a 3 kW motor. The motor data are given<br />

in appendix A.3. Required control parameters are set as follows: settling time < 0.003,<br />

overshoot < 4.33%. For these parameters a root loci of the close-loop is obtained, see<br />

Fig. 4.18.<br />

Root Locus Editor (C)<br />

1500<br />

0.93<br />

0.87<br />

0.78<br />

0.64<br />

0.46<br />

0.24<br />

0.97<br />

1000<br />

0.992<br />

500<br />

Imag Axis<br />

0<br />

4e+003<br />

3e+003<br />

2e+003 1e+003<br />

-500<br />

0.992<br />

-1000<br />

0.97<br />

-1500<br />

0.93 0.87<br />

0.78<br />

0.64 0.46<br />

0.24<br />

-4500 -4000 -3500 -3000 -2500 -2000 -1500 -1000 -500 0<br />

Real Axis<br />

Fig. 4.18. Root loci of the close-loop stator flux control system<br />

From the position of the poles, the parameters of the PI flux controller are obtained:<br />

K = 2531, T = 0. 00074 .<br />

pΨ<br />

iΨ<br />

The behaviour of the flux control loop with parameters like above was tested using<br />

SABER simulation package. The model created in SABER takes into account the full<br />

control system, including the models of inverter and induction motor (see appendix<br />

A.2). The flux step response is presented in Fig. 4.19. The simulation result confirms a<br />

good dynamics of the flux and proper operation in the steady state.<br />

82


4.3. Analysis and Controller Design for <strong>DTC</strong>-<strong>SVM</strong> Method<br />

Fig. 4.19. Simulated (SABER) flux response for controller parameters designed<br />

according to root locus method<br />

<strong>Torque</strong> Controller Design<br />

Based on motor model equations (4.10 - 4.12), the following equation can be<br />

obtained:<br />

⎡<br />

⎢<br />

⎣<br />

( R L + R L ) + σL<br />

L I = L U − L Ψ p Ω + I σL<br />

L ( Ω − p Ω )<br />

s<br />

r<br />

r<br />

s<br />

s<br />

r<br />

d ⎤<br />

dt ⎥<br />

⎦<br />

sy<br />

r<br />

sy<br />

r<br />

s<br />

b<br />

m<br />

sx<br />

s<br />

r<br />

ss<br />

b<br />

m<br />

L 2<br />

M<br />

where: σ = 1−<br />

L L<br />

s<br />

r<br />

Under the assumption that the last term in equation (4.25) is very small:<br />

sx<br />

s<br />

r<br />

( Ω − p Ω ) ≈ 0<br />

ss<br />

b<br />

m<br />

(4.25)<br />

I σ L L<br />

(4.26)<br />

the equation (4.25) becomes:<br />

⎡<br />

⎢<br />

⎣<br />

d ⎤<br />

+ σ<br />

dt ⎥<br />

(4.27)<br />

⎦<br />

( Rs<br />

Lr<br />

Rr<br />

Ls<br />

) + Ls<br />

Lr<br />

I<br />

sy<br />

= LrU<br />

sy<br />

− LrΨ<br />

s<br />

pbΩm<br />

The additional assumption is that the motor is not loaded M = 0 .<br />

Under those assumptions the rotor speed can be expressed:<br />

L<br />

83


4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control with <strong>Space</strong> <strong>Vector</strong> Modulation (<strong>DTC</strong>-<strong>SVM</strong>)<br />

dΩ<br />

dt<br />

m<br />

1<br />

J<br />

m<br />

s<br />

= pb<br />

Ψ<br />

s<br />

I<br />

sy<br />

(4.28)<br />

2<br />

From equation (4.14) current I<br />

sy<br />

can be expressed as follows:<br />

I<br />

sy<br />

2<br />

= M<br />

e<br />

(4.29)<br />

p m Ψ<br />

b<br />

s<br />

s<br />

If both sides of equation (4.27) are differentiated, this equation becomes:<br />

⎡<br />

⎢<br />

⎢⎣<br />

( R L + R L )<br />

s<br />

r<br />

r<br />

s<br />

d<br />

dt<br />

⎛<br />

+ σ Ls<br />

Lr<br />

⎜<br />

⎝<br />

d<br />

dt<br />

2<br />

⎞<br />

⎟<br />

⎠<br />

⎤<br />

⎥I<br />

⎥⎦<br />

sy<br />

= L<br />

r<br />

dU<br />

dt<br />

sy<br />

− L Ψ<br />

r<br />

s<br />

p<br />

b<br />

dΩ<br />

dt<br />

m<br />

(4.30)<br />

Based on the equations (4.30), (4.28) and (4.29) the open-loop torque transfer<br />

function can be obtained as follows:<br />

G<br />

M<br />

() s<br />

M<br />

U<br />

e<br />

M<br />

= =<br />

(4.31)<br />

2<br />

sy<br />

s<br />

A s<br />

+ B s + C<br />

M<br />

M<br />

where:<br />

A<br />

M<br />

p m Ψ<br />

2σL<br />

b s s<br />

= ;<br />

s<br />

B<br />

M<br />

R L + R L<br />

σL<br />

L<br />

s r r s<br />

= ;<br />

s<br />

r<br />

C<br />

M<br />

=<br />

2<br />

pb<br />

msΨ<br />

2σL<br />

J<br />

s<br />

2<br />

s<br />

The torque control loop is shown in Fig. 4.20, where G RM<br />

( s)<br />

is a transfer function of<br />

the PI controller given by equation (4.17).<br />

M ec<br />

G RM<br />

() s<br />

U<br />

sy<br />

G M<br />

( s)<br />

M<br />

e<br />

Fig. 4.20. <strong>Torque</strong> control loop<br />

The input data to the SISO Design Tool are obtained in the same way like for the<br />

flux. The transfer functions are calculated for the 3 kW motor from the equation (4.17)<br />

and (4.31). The required control parameters are set as follows: settling time < 0.0015,<br />

overshoot < 2%. For these parameters a root loci of the close-loop is obtained, see Fig.<br />

4.21. From the position of the poles (Fig. 4.21), the parameters of the PI torque<br />

controller are obtained: K = 33. 21, T = 0. 00045 .<br />

pM<br />

iM<br />

84


4.3. Analysis and Controller Design for <strong>DTC</strong>-<strong>SVM</strong> Method<br />

Root Locus Editor (C)<br />

2500<br />

0.93<br />

0.87<br />

0.78<br />

0.66<br />

0.48<br />

0.24<br />

2000<br />

0.97<br />

1500<br />

1000<br />

0.992<br />

500<br />

Imag Axis<br />

0<br />

7e+003 6e+003<br />

5e+003<br />

4e+003<br />

3e+003 2e+003<br />

1e+003<br />

-500<br />

-1000<br />

0.992<br />

-1500<br />

-2000<br />

0.97<br />

-2500<br />

0.93 0.87<br />

0.78<br />

0.66<br />

0.48<br />

0.24<br />

-7000 -6000 -5000 -4000 -3000 -2000 -1000 0<br />

Real Axis<br />

Fig. 4.21. Root loci of the close-loop torque control system<br />

The transfer function of the close loop torque control shown in Fig. 4.20 is given as:<br />

G<br />

Mc<br />

() s<br />

M<br />

=<br />

M<br />

e<br />

ec<br />

=<br />

s<br />

2<br />

+<br />

( A K + B )<br />

M<br />

A<br />

M<br />

pM<br />

T<br />

K<br />

iM<br />

pM<br />

M<br />

( T s + 1)<br />

iM<br />

s + C<br />

M<br />

+<br />

A<br />

M<br />

T<br />

K<br />

iM<br />

pM<br />

(4.32)<br />

The SISO Design Tool enables to observe the step response of the investigated<br />

control system. In the Fig. 4.22 is shown the step response of the torque control system<br />

from Fig. 4.20 described by equation (4.32), with the PI controller parameters setting as:<br />

K = 33.21, T = 0. 00045 .<br />

pM<br />

iM<br />

1.4<br />

Step Response<br />

From: r<br />

1.2<br />

1<br />

Amplitude<br />

To: y<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 0.5 1 1.5 2 2.5 3 3.5 4<br />

Time (sec)<br />

x 10 -3<br />

Fig. 4.22. Simulated (Matlab) step response of the system from Fig. 4.20 described by transfer<br />

function given by equation (4.32)<br />

85


4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control with <strong>Space</strong> <strong>Vector</strong> Modulation (<strong>DTC</strong>-<strong>SVM</strong>)<br />

It should be note that moment of inertia J can change during drive operation (for<br />

example in still industry systems). However, the value of coefficient<br />

(4.32) normally is several order lower in comparison with ( A K T )<br />

it’s influence on torque close loop dynamic can be neglected.<br />

M<br />

pM<br />

C<br />

M<br />

iM<br />

, in equation<br />

. Therefore,<br />

Because of the forcing element in transfer function (4.32) the step response presented<br />

in Fig. 4.22 characterized much higher overshoot then the assumed 2%.<br />

To compensate the forcing element in the numerator (4.32) a prefilter is inserted into<br />

the reference channel of the torque controller. The transfer function of the prefilter is<br />

given as:<br />

1<br />

GFM () s =<br />

(4.33)<br />

T s + 1<br />

F<br />

The time constant of the prefilter is equal time constant of the torque controller<br />

T<br />

F<br />

= T iM<br />

.<br />

The full control loop of torque with prefilter is shown in Fig. 4.23. The step response<br />

of this control loop is presented in Fig. 4.24.<br />

M<br />

ec<br />

G FM<br />

() s<br />

G RM<br />

( s)<br />

U sy<br />

G M<br />

( s)<br />

M<br />

e<br />

Fig. 4.23. <strong>Torque</strong> control loop with prefilter<br />

1.4<br />

Step Response<br />

From: r<br />

1.2<br />

1<br />

Amplitude<br />

To: y<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5<br />

Time (sec)<br />

x 10 -3<br />

Fig. 4.24. Simulated (Matlab) step response of the system from Fig. 4.23<br />

86


4.3. Analysis and Controller Design for <strong>DTC</strong>-<strong>SVM</strong> Method<br />

Figure 4.24 shows that the torque control loop with a prefilter incorporated into the<br />

reference channel reduces considerably the overshoot.<br />

The behaviour of the torque control loop with the same settings of the parameters<br />

was also tested in SABER simulation model. The torque step response is presented in<br />

Fig. 4.25. The result of simulation confirms a good dynamics of the torque and proper<br />

operation in the steady state.<br />

Fig. 4.25. Simulated (SABER) torque response<br />

<strong>Torque</strong> Controller Design for High Power Motor<br />

The same method of tuning the controllers was used for a 90 kW motor. The<br />

parameters of this motor can be found in appendix A.3. The required control parameters<br />

are set as follows: for the flux settling time < 0.003, overshoot < 4.33% and for the<br />

torque settling time < 0.0015, overshoot < 2%. The parameters of the controllers are<br />

obtained as follows: flux controller K = 2592 , T = 0. 00076 and torque controller<br />

K = 1.8492 , T = 0. 00046 .<br />

pM<br />

iM<br />

pΨ<br />

The simulation model of drive with a 90 kW motor was also build in the SABER<br />

package.<br />

The flux step response is presented in Fig. 4.26. The control loop of the flux is<br />

identical for both motors (Fig. 4.8) and does not depend on the motor parameters.<br />

Therefore, the parameters of the flux controller and the result of simulation (Fig. 4.26)<br />

is very similar to the result for the 3 kW motor (Fig. 4.19).<br />

iΨ<br />

87


4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control with <strong>Space</strong> <strong>Vector</strong> Modulation (<strong>DTC</strong>-<strong>SVM</strong>)<br />

The torque response for the 90 kW motor is presented in Fig. 4.27. The results of the<br />

simulations (Fig. 4.26, 4.27), similarly like in the case of the small power ratting motor,<br />

confirm a good dynamics of the torque and a proper operation in the steady state.<br />

Fig. 4.26. Simulated (SABER) flux response for 90 kW motor<br />

Fig. 4.27. Simulated (SABER) torque response for 90 kW motor<br />

4.3.3. Summary of Flux and <strong>Torque</strong> Controllers Design<br />

In the Fig. 4.28 a full control structure of the <strong>DTC</strong>-<strong>SVM</strong> scheme is shown. This<br />

scheme is completed on the prefilter, compared to the basic scheme form Fig. 4.5.<br />

88


4.3. Analysis and Controller Design for <strong>DTC</strong>-<strong>SVM</strong> Method<br />

The presented above controller tuning algorithm is based on the open-loop transfer<br />

function for the flux (equation 4.24) and for the torque (equation 4.31). These transfer<br />

functions are obtained under the assumptions (4.22) and (4.26) respectively. Because of<br />

the assumed simplifications, the results of full model simulations are slightly differ form<br />

the initially expected values.<br />

Flux<br />

Controller<br />

Ψ sc<br />

PI<br />

U sxc<br />

x − y<br />

S A<br />

U sc<br />

<strong>SVM</strong><br />

S B<br />

M ec<br />

F<br />

PI<br />

U syc<br />

α − β<br />

S C<br />

Prefilter<br />

<strong>Torque</strong><br />

Controller<br />

Ψˆ s<br />

Mˆ<br />

e<br />

γˆss<br />

Flux and<br />

<strong>Torque</strong><br />

Estimator<br />

U s<br />

I s<br />

Voltage<br />

Calculation<br />

α − β<br />

U dc<br />

I A<br />

ABC<br />

I B<br />

Fig. 4.28. Full scheme of the <strong>DTC</strong>-<strong>SVM</strong> control method<br />

Additional assumption for the torque controller analysis is that the stator flux<br />

magnitude is constant. Therefore, decoupling between flux and torque control loops is<br />

important. In Fig. 4.29 the torque step response (Fig. 4.29a) and magnitude stator flux<br />

step response (Fig. 4.29b) are shown. From Fig. 4.29 can be seen that both controllers<br />

are very fast and decoupling between flux and torque is correct.<br />

The full control structure (Fig. 4.28) is different from the basic scheme, which can be<br />

seen in Fig. 4.8. In the torque reference channel a prefilter is incorporated. The basic<br />

structure assumed four controllers parameters:<br />

K<br />

pΨ<br />

, T<br />

iΨ<br />

, K<br />

pM<br />

and T iM<br />

. The addition<br />

of the prefilter does not introduce any additional parameters, because the time constant<br />

of the prefilter is equal to the torque controller integrating time T<br />

iM<br />

(see equation 4.33).<br />

Thus the control methods needs only four parameters.<br />

Additionally, if a very fast torque response is not required, the prefilter time constant<br />

can be increased independently from the torque controller parameters in order to<br />

improve the stability of the system.<br />

89


4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control with <strong>Space</strong> <strong>Vector</strong> Modulation (<strong>DTC</strong>-<strong>SVM</strong>)<br />

a)<br />

b)<br />

Fig. 4.29. Dynamic tests a) torque step change, b) flux step change. From the top: reference and estimated<br />

torque, reference and estimated stator flux<br />

In section 4.3 two methods of flux and torque controller design for <strong>DTC</strong>-<strong>SVM</strong> are<br />

presented. The comparison of the result obtained in two methods is summarized in<br />

Table 4.3. The summary is done for the 3kW motor and sampling frequency<br />

f s<br />

= 10kHz . The first method uses simplified IM model and is based on symmetric<br />

optimum criterion. However, this approach gives good results only for flux control loop.<br />

The second approach uses dynamic model of IM including rotor parameters and is<br />

based on root locus method. The results obtained in simulation are good for both flux<br />

and torque controllers. However, it is much more complicated than first method.<br />

The dynamic of the flux control loop is very similar in both cases. Therefore, to tune<br />

flux controller symmetry criterion should be used because it is simpler.<br />

90


4.3. Analysis and Controller Design for <strong>DTC</strong>-<strong>SVM</strong> Method<br />

Symmetry<br />

Criterion<br />

Method<br />

Table 4.3. Summary of controller design<br />

Model parameters<br />

Controller parameters<br />

Flux <strong>Torque</strong> Flux <strong>Torque</strong><br />

T s<br />

T =1<br />

pure integrator<br />

R<br />

r<br />

Root Locus A<br />

Ψ<br />

=<br />

Method<br />

σ L<br />

r<br />

R<br />

r<br />

B =<br />

C<br />

Ψ<br />

Ψ<br />

=<br />

R<br />

σ L<br />

s<br />

L<br />

s<br />

σ L<br />

s<br />

R<br />

L<br />

r<br />

+<br />

r<br />

s<br />

R<br />

L<br />

r<br />

s<br />

L<br />

r<br />

T s<br />

p ,<br />

A<br />

B<br />

C<br />

b<br />

, m<br />

s<br />

, Ψ<br />

s<br />

R<br />

M<br />

M<br />

M<br />

=<br />

=<br />

=<br />

L<br />

R<br />

p<br />

p<br />

b<br />

m<br />

sΨ<br />

2σ L<br />

s<br />

L<br />

r<br />

L<br />

r<br />

+ R<br />

σ L L<br />

r<br />

s<br />

s<br />

m<br />

sΨ<br />

2σ L L<br />

b<br />

2<br />

s<br />

s<br />

r<br />

s<br />

r<br />

r<br />

2<br />

s<br />

J<br />

L<br />

L<br />

s<br />

r<br />

K pΨ<br />

T iΨ<br />

K pΨ<br />

T iΨ<br />

= 5000<br />

= 0.00040<br />

= 2531<br />

= 0.00074<br />

K pM<br />

T iM<br />

K pM<br />

T iM<br />

= 24.00<br />

= 0.00080<br />

= 33.21<br />

= 0.00045<br />

Dynamic parameters<br />

Flux<br />

<strong>Torque</strong><br />

t nΨ<br />

= 0. 0015s<br />

t nM<br />

= 0. 0007s<br />

p Ψ<br />

=1.6%<br />

p M<br />

= 8.39%<br />

t nΨ<br />

= 0. 0019s<br />

t nM<br />

= 0. 0009s<br />

p Ψ<br />

= 1.49%<br />

p M<br />

= 1.04%<br />

91


4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control with <strong>Space</strong> <strong>Vector</strong> Modulation (<strong>DTC</strong>-<strong>SVM</strong>)<br />

All simulation results for root locus method presented in section 4.3.2 were done at<br />

sampling frequency<br />

f s<br />

= 10kHz<br />

. However, presented controller design method<br />

provides to obtain controller parameters for different sampling frequency. This aspect<br />

will be presented for the torque controller. When the sampling frequency is changed the<br />

input parameters: settling time and overshoot must be modified. For lower sampling<br />

frequency the dynamic of control loop is decreasing [34]. Thus, for the continuous<br />

analysis, which is used in root locus method, the settling time should be increased and<br />

overshoot reduced.<br />

Table 4.4 shows torque controller parameters calculated for three sampling frequency<br />

values:<br />

f s<br />

= 10kHz<br />

, f s<br />

= 5kHz<br />

and f s<br />

= 2. 5kHz<br />

.<br />

Table 4.4. <strong>Torque</strong> controller parameters for different sampling frequency<br />

f s settling time overshoot K p Μ T i Μ<br />

10.0 kHz 0.0015 2% 33.21 0.00045<br />

5.0 kHz 0.0030 1% 15.88 0.00098<br />

2.5 kHz 0.0060 1% 7.12 0.00180<br />

Simulated results obtained for parameters presented in Table 4.4 are shown in Fig.<br />

4.30. The result of simulation confirms a good behavior of the system for all three<br />

sampling frequencies.<br />

The root locus method gives proper results for different motor type. It confirms<br />

results obtained for the 90 kW motor.<br />

The very important features of the <strong>DTC</strong>-<strong>SVM</strong> in comparison with classical <strong>DTC</strong> are<br />

performance in steady state. In the Fig. 4.31 the steady state operation of the <strong>DTC</strong>-<strong>SVM</strong><br />

control system is shown. It can be seen that the line current is sinusoidal and voltage has<br />

an unipolar waveform. Presented in Fig. 4.31 can be compared with simulation results<br />

for classical <strong>DTC</strong> from Fig. 3.16, where controller just select voltage vectors to reduce<br />

instantaneous flux and torque errors, and does not implement the true PWM. Therefore,<br />

inverter output voltage is not unipolar. This increase switching losses of the<br />

semiconductor power devices.<br />

92


4.3. Analysis and Controller Design for <strong>DTC</strong>-<strong>SVM</strong> Method<br />

a)<br />

b)<br />

c)<br />

Fig. 4.30. 3 kW motor torque response for controller parameters calculated according to root locus<br />

method at different sampling frequency a) f s = 10kHz<br />

, b) f s = 5kHz<br />

, c) f s = 2. 5kHz<br />

93


4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control with <strong>Space</strong> <strong>Vector</strong> Modulation (<strong>DTC</strong>-<strong>SVM</strong>)<br />

Fig. 4.31. Steady state operation. From the top: line to line voltage, line current<br />

The features of the <strong>DTC</strong>-<strong>SVM</strong> method can be summarized as follows:<br />

• good dynamic control of flux and torque,<br />

• constant switching frequency,<br />

• unipolar voltage thanks to use of PWM block (<strong>SVM</strong>),<br />

• low flux and torque ripple,<br />

• sinusoidal stator currents.<br />

4.4. Speed Controller Design<br />

If the stator flux is assumed constant, Ψ s<br />

= const.<br />

, that based on the equations (4.13)<br />

and (4.14) dynamic of IM can be described as:<br />

dΩ<br />

dt<br />

m<br />

[ M − M ]<br />

= 1 e L<br />

(4.34)<br />

J<br />

A block diagram of the speed control loop is shown in Fig. 4.32, where G RS<br />

() s is a<br />

'<br />

transfer function of PI controller (see equation 4.17) and G M<br />

( s)<br />

is a transfer function of<br />

full torque control loop. In the speed controller design process the filter for the<br />

measured value should be taken into consideration. T<br />

f<br />

is a time constant of the filter.<br />

The low pass filter is necessary in hardware setup.<br />

94


4.4. Speed Controller Design<br />

M L<br />

Ω mc<br />

() s<br />

M<br />

ec<br />

'<br />

G RS G M<br />

( s)<br />

M<br />

e<br />

1<br />

J<br />

1 Ωm<br />

s<br />

1<br />

T f<br />

s + 1<br />

Fig. 4.32. Block diagram of the speed control loop<br />

The transfer function of the full torque control loop (Fig. 4.23) can be calculated as:<br />

G<br />

'<br />

M<br />

M<br />

M<br />

e<br />

() s = = G () s ⋅G<br />

() s<br />

ec<br />

FM<br />

Mc<br />

(4.35)<br />

where: G Mc<br />

( s)<br />

- torque control loop transfer function given by equation (4.32),<br />

G FM<br />

() s - prefilter transfer function given by equation (4.33).<br />

'<br />

The transfer function G M<br />

( s)<br />

can by expressed as:<br />

where:<br />

'<br />

' AM<br />

GM () s =<br />

(4.36)<br />

' 2 '<br />

B s + C s + 1<br />

A<br />

'<br />

M<br />

=<br />

C<br />

M<br />

T<br />

A<br />

iM<br />

M<br />

M<br />

K<br />

pM<br />

+ A<br />

M<br />

K<br />

M<br />

pM<br />

;<br />

B<br />

'<br />

M<br />

=<br />

C<br />

M<br />

T<br />

iM<br />

TiM<br />

+ A<br />

M<br />

K<br />

pM<br />

; C<br />

'<br />

M<br />

T<br />

=<br />

C<br />

iM<br />

( A K + B )<br />

M<br />

T<br />

M<br />

iM<br />

pM<br />

+ A<br />

M<br />

K<br />

M<br />

pM<br />

The torque control loop can be approximate by first order integrating part, because<br />

of:<br />

'<br />

M<br />

B ≈ 0<br />

(4.37)<br />

The simplified transfer function can be written as:<br />

'<br />

' AM<br />

GM () s =<br />

(4.38)<br />

'<br />

C s + 1<br />

M<br />

For the torque controller parameters K =15. 87 , T = 0. 00087 obtained in section<br />

pM<br />

4.3.3 at the sampling frequency f s<br />

= 5kHz<br />

the transfer function parameters have values:<br />

iM<br />

95


4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control with <strong>Space</strong> <strong>Vector</strong> Modulation (<strong>DTC</strong>-<strong>SVM</strong>)<br />

'<br />

M<br />

'<br />

'<br />

M<br />

A = 0.9944 , B M<br />

= 3.563e − 007 , C = 0. 0009329. Those parameters confirm that<br />

assumption (4.37) is correct.<br />

The step response of the full and simplified transfer function are shown in Fig. 4.33.<br />

25<br />

20<br />

15<br />

full transfer<br />

function<br />

simplified<br />

transfer function<br />

10<br />

5<br />

0<br />

-5<br />

0 0.005 0.01 0.015 0.02 0.025 0.03<br />

Time<br />

Fig. 4.33. <strong>Torque</strong> response for full and simplified transfer function<br />

For the speed controller parameter design the symmetry criterion can by applied [66].<br />

In accordance with the symmetry criterion the plant transfer function can be written as:<br />

G<br />

() s<br />

=<br />

K e<br />

sT<br />

−sτ<br />

0<br />

c<br />

2<br />

1<br />

1<br />

( + sT )<br />

(4.39)<br />

where:<br />

converter),<br />

K<br />

c<br />

= A M<br />

' is gain of the plan, τ 0<br />

is dead time of the inverter ( τ 0<br />

= 0 ideal<br />

T<br />

2<br />

= J , and T<br />

1<br />

= C + Tf<br />

is a sum of small time constants. The optimal<br />

controller parameters can be calculated as:<br />

K<br />

ps<br />

=<br />

K<br />

T<br />

2<br />

( T + τ ) 2( + T )<br />

2<br />

c 1 0<br />

=<br />

J<br />

C<br />

( T + ) = ( C )<br />

is<br />

= 4<br />

0<br />

4 T f<br />

f<br />

(4.40)<br />

T<br />

1<br />

τ +<br />

(4.41)<br />

For the filter frequency<br />

f f<br />

= 25Hz<br />

where:<br />

T<br />

f<br />

1<br />

= (4.42)<br />

2πf<br />

f<br />

96


4.4. Speed Controller Design<br />

the speed controller parameters are obtained as follows: K = 1. 33; T = 0. 0292 .<br />

Fig. 4.34, 4.35 and 4.36 show simulation and experimental results for the system<br />

operated with speed controller parameters obtained above. The speed reversals are<br />

presented in Fig. 4.34 and 4.35 for high and small reference speed differences<br />

respectively. The step change of the load torque at constant speed is presented in Fig.<br />

4.36. All presented in Fig. 4.34, 4.35 and 4.36 results confirm proper operation of the<br />

speed control loop.<br />

ps<br />

is<br />

a) b)<br />

Fig. 4.34. Speed reversal Ω m = ±100rad<br />

/ s a) simulated (SABER), b) experimental 1) reference speed<br />

(75 (rad/s)/div), 2) actual speed (75 (rad/s)/div), 3) reference torque (20 Nm/div)<br />

a) b)<br />

Fig. 4.35. Speed reversal - small signal Ω m = ±5rad<br />

/ s a) simulated (SABER), b) experimental 1)<br />

reference speed (7.5 (rad/s)/div), 2) actual speed (7.5 (rad/s)/div), 3) reference torque (20 Nm/div)<br />

97


4. <strong>Direct</strong> Flux and <strong>Torque</strong> Control with <strong>Space</strong> <strong>Vector</strong> Modulation (<strong>DTC</strong>-<strong>SVM</strong>)<br />

a) b)<br />

Fig. 4.36. Load torque step change at Ω m<br />

= 100rad<br />

/ s a) simulated (SABER), b) experimental<br />

1) reference speed (30 (rad/s)/div), 2) actual speed (30 (rad/s)/div), 3) estimated torque (20 Nm/div)<br />

4.5. Summary<br />

This chapter gives review of <strong>DTC</strong>-<strong>SVM</strong> control methods. To analysis and<br />

implementation was chosen <strong>DTC</strong>-<strong>SVM</strong> method with close-loop torque and flux control<br />

in stator flux coordinates. Full mathematical analysis of IM drive working with this<br />

control method is presented. Two different flux and torque controllers design algorithm<br />

are analyzed and discussed. Furthermore, speed controller tuning methods is shown.<br />

The flux and torque controller design methods for sampling frequency changes and<br />

different motor power are discussed. The analysis presented in this chapter give<br />

complex knowledge about control structure and controller design methods. Obtained<br />

parameters provide good dynamic and steady state operation of a drive. It is confirmed<br />

by simulation and experimental results presented in this chapter and in Chapter 7.<br />

98


5. Estimation in Induction Motor Drives<br />

5.1. Introduction<br />

The vector control methods of induction motor require feedback signals. This is an<br />

information about flux, torque and mechanical speed in drives operated without<br />

mechanical sensor (sensorless operation mode).<br />

There are many different method to obtain these state variables of induction motor.<br />

Basic methods can be divided into three main group [87]:<br />

• physical methods <strong>–</strong> based on nonlinear construction of IM [60, 77, 113],<br />

• mathematical models <strong>–</strong> used mathematical description of IM and control theory,<br />

• neural network methods <strong>–</strong> based on the artificial intelligence techniques [9, 91,<br />

95].<br />

The general classification of the state variables calculation methods is presented in<br />

Fig. 5.1 [87].<br />

Induction motor state variables<br />

calculation methods<br />

Physical<br />

methods<br />

Mathematical<br />

models<br />

Neural network<br />

methods<br />

Estimators of<br />

state variables<br />

Observer of<br />

state variables<br />

Kalman Filter<br />

Fig. 5.1. Classification of induction state variables calculation methods<br />

The mathematical models is based on the space vector equations, which describe<br />

induction motors. Fig. 5.1 shows division of these methods into three groups:<br />

• estimators of state variables,<br />

• observer of state variables,


5. Estimation in Induction Motor Drives<br />

• Kalman filter.<br />

The <strong>DTC</strong>-<strong>SVM</strong> method is based on the information about stator flux vector (see<br />

section 4.3). Therefore, it is the most important variable of the motor. Measurement of<br />

flux in motor is difficult and demands special sensor. This solution is very expensive<br />

and complicated. Because of that a method of calculation motor flux was developed.<br />

In vector control methods this part of algorithm is especially important. Estimation<br />

algorithm uses as input signals values, which are simple to measure. There are current<br />

and voltage signals. Obviously new methods aim at reducing number of sensors for<br />

more reliable operation and lower price of a drive.<br />

The motor flux is the main component to calculate torque and speed. Therefore,<br />

accuracy of the estimation flux is very important. Flux estimation is a significant task in<br />

implementing of high-performance motor drives.<br />

The advanced state variables calculation algorithm is characterized by:<br />

• accuracy in steady and dynamic states,<br />

• robustness for motor parameters variation,<br />

• minimal number of sensor,<br />

• operation in whole speed range,<br />

• low calculation demanded.<br />

All estimation algorithms based on the motor parameters. These parameters change<br />

in time work of the drive. For instance, with change the temperature. Therefore,<br />

estimation algorithm have to be less sensitive to the parameters variations.<br />

All presented flux estimation algorithms are shown as stator flux estimators, because<br />

of these algorithms work with <strong>DTC</strong>-<strong>SVM</strong> structure. In some algorithm rotor flux<br />

estimation is required, but in this case it is convert on stator flux.<br />

5.2. Estimation of Inverter Output Voltage<br />

Input signals for the estimators are measurements of stator currents and voltages<br />

which are recreated from the switching signals. Switch signals for the each inverter<br />

phase are obtained by control algorithm. The reference voltage vector is realized by<br />

100


5.2. Estimation of Inverter Output Voltage<br />

modulator (see section 2.4). However, duty times are modified by dead-time, which is<br />

requisite for correct inverter operation (see section 2.3). Because of this modification<br />

delivered to the motor voltage is different from reference. To eliminate dead-time effect<br />

there is a special part for compensation of dead-time in control algorithms. Obtained by<br />

vector modulator duty cycles, represented by switching signals S A , S B , S C are modified<br />

to S ' A , S ' '<br />

B , S C (Fig. 5.2). This modification depends on the phase current direction and is<br />

realized for each phase. Many different dead-time compensation methods are presented<br />

in literature [2, 3, 8, 29, 64, 76]. Thanks to this modification after change signals by<br />

dead-time, a correct voltage vector obtained by controller is delivered to the motor.<br />

Because of that signals S A , S B , S C are used to recreate voltage values. The voltage is<br />

calculated form the equations:<br />

2<br />

U<br />

sα<br />

= U<br />

dc A<br />

5<br />

B<br />

+<br />

3<br />

( D − 0. ( D D ))<br />

C<br />

(5.1a)<br />

U<br />

( D − D )<br />

s<br />

U<br />

(5.1b)<br />

β dc B C<br />

= 3<br />

3<br />

and<br />

where D A , D B , D C are duty cycles corresponding to the switching signals S A , S B , S C<br />

U<br />

dc<br />

is the voltage of inverter dc-link.<br />

U dc<br />

U<br />

s β c<br />

U<br />

s α c<br />

<strong>Vector</strong><br />

Modulator<br />

S A<br />

S B<br />

S C<br />

Dead<br />

Time<br />

&<br />

Voltage<br />

Drop<br />

Compensation<br />

S A<br />

'<br />

S B<br />

'<br />

S C<br />

'<br />

Dead<br />

Time<br />

S A<br />

+<br />

S A<br />

-<br />

S B<br />

+<br />

S B<br />

-<br />

S C<br />

+<br />

S C<br />

-<br />

U sα<br />

U sβ<br />

Voltage<br />

Calculation<br />

U dc<br />

Is<br />

I s<br />

Motor<br />

Fig. 5.2. Input signals for the estimators<br />

101


5. Estimation in Induction Motor Drives<br />

In Fig. 5.2 voltage calculation block diagram is shown. Simultaneously with deadtime<br />

compensation a voltage drop compensation algorithm is realized. It is especially<br />

important for low speed operation range, when voltage is very low.<br />

The main assumption in voltage calculation method is that identical voltage vector,<br />

which is calculated by a controller is delivered to the motor. It means, proper<br />

information about voltage depends on correct implementation dead-time and voltage<br />

drop compensation algorithms.<br />

Dead <strong>–</strong> Time Compensation<br />

In order to prevent shortcircuiting an inverter leg, there should be a dead-time (T D )<br />

between the turn-off one switch (IGBT) and the turn-on of the next one (from the same<br />

leg). T D should be larger than the maximum storage time of the switching device. The<br />

effect of the dead-time is a voltage distortion delivered to the motor. The voltage<br />

distortion ∆U is depending on current sign, as can be seen in Fig. 5.3.<br />

a)<br />

T 1<br />

b)<br />

T 1<br />

U dc<br />

2<br />

0<br />

C<br />

S A +<br />

T 2<br />

A<br />

D 1<br />

I A<br />

> 0<br />

U dc<br />

2<br />

0<br />

C<br />

S A +<br />

T 2<br />

A<br />

D 1<br />

< 0 I A<br />

U dc<br />

2<br />

C<br />

S A -<br />

D 2<br />

U dc<br />

2<br />

C<br />

S A -<br />

D 2<br />

I > 0<br />

< 0<br />

A<br />

I A<br />

S A<br />

T D<br />

T D<br />

S A<br />

T D<br />

T D<br />

t<br />

t<br />

S A<br />

+<br />

S A<br />

+<br />

t<br />

t<br />

S A<br />

-<br />

S A<br />

-<br />

U A0<br />

dc<br />

t<br />

t<br />

1<br />

U<br />

2<br />

U A0<br />

dc<br />

1<br />

U<br />

2<br />

0<br />

t<br />

0<br />

t<br />

1<br />

−<br />

2<br />

U dc<br />

1<br />

−<br />

2<br />

U dc<br />

Fig. 5.3. Dead-time effect for different current sing a) I > 0 , b) I < 0<br />

A<br />

A<br />

102


5.2. Estimation of Inverter Output Voltage<br />

So the real voltage vector across the motor can be expressed as:<br />

U<br />

mot<br />

= U − ∆U<br />

(5.2)<br />

sc<br />

The voltage distortion ∆U can be written as:<br />

( )<br />

∆U = T f U sign<br />

(5.3)<br />

D<br />

s<br />

dc<br />

I s<br />

where: f<br />

s<br />

- sampling frequency,<br />

sign ()- signum function.<br />

The dead-time compensation can be implemented by adjusting the phase duty cycles<br />

as following:<br />

'<br />

k<br />

D = D + T<br />

k<br />

D<br />

f<br />

s<br />

sign<br />

( I )<br />

k<br />

(5.4)<br />

where:<br />

k = A,<br />

B,<br />

C .<br />

This means that the on-time of the upper bridge arm switch is shortened by T D and<br />

for positive current it is increased by the same amount for negative current.<br />

Because of the current has ripple around zero-crossing the algorithm should be<br />

modified. One of the possible solutions is method with current level. In this method the<br />

current level ( I<br />

level<br />

) is defined, which describes zone around the zero current as:<br />

− I > I > I<br />

(5.5)<br />

level<br />

k<br />

level<br />

If the condition (5.6) is performed the duty cycles are modified as follows:<br />

'<br />

k<br />

D = D +<br />

k<br />

I<br />

I<br />

k<br />

level<br />

T<br />

D<br />

f<br />

s<br />

sign<br />

( I )<br />

k<br />

(5.6)<br />

In the other cases the duty cycles are modified according to the equation (5.4).<br />

The value of the current level ( I<br />

level<br />

) depends on the motor power and can be<br />

deducted experimentally. For 3kW drive the optimal value of current level was<br />

I level<br />

= 0. 1 A .<br />

The simulated results for the dead-time compensation algorithms are presented in<br />

Fig. 5.4. In this test drive operates with scalar control (U/f=const.) algorithm at<br />

fundamental frequency<br />

f = 2 Hz .<br />

103


5. Estimation in Induction Motor Drives<br />

a)<br />

b)<br />

Fig. 5.4. Simulated U/f=const. control method at frequency f = 2Hz<br />

a) without dead-time compensation,<br />

b) with dead-time compensation<br />

From Fig. 5.4a it can be seen that without dead-time compensation the output<br />

currents are considerably distorted and has reduced value. Fig. 5.4b shown simulated<br />

result with dead-time compensation algorithm. Thanks of the compensation proper<br />

voltage is delivered to the motor. Therefore, currents have correct value and currents<br />

waveforms are sinusoidal.<br />

Presented dead-time compensation algorithm was implemented in final control<br />

system.<br />

5.3. Stator Flux <strong>Vector</strong> Estimators<br />

The flux vector estimator algorithms can be divided into two groups in terms of the<br />

input signal. The currents and voltages are the input signals to the voltage models (VM),<br />

while the currents and speed or position information are input signals to the current<br />

models (CM). Obviously, for sensorless control structures general voltage models with<br />

many different modifications and improvements are used.<br />

The stator flux can be directly obtained from the motor model equation (2.10a) as<br />

follows:<br />

∫<br />

( U − I )<br />

Ψˆ = dt<br />

(5.7)<br />

s<br />

s<br />

R s<br />

s<br />

104


5.3. Stator Flux <strong>Vector</strong> Estimators<br />

This is a classical voltage model of stator flux vector estimation, which obtain flux<br />

by integrating the motor back electromagnetic force (EMF). The block diagram of this<br />

estimator is shown in the Fig. 5.5.<br />

I s<br />

R s<br />

U s<br />

∫<br />

Ψˆ<br />

s<br />

Fig. 5.5. Voltage model based estimator with pure integrators<br />

This method is sensitive for only one motor parameter, stator resistance. However,<br />

the implementation of pure integrator is difficult because of dc drift and initial value<br />

problems. Moreover, when estimator based on pure integrator in control structure are<br />

additional disadvantages. Using a pure integrator to estimate the stator flux it is not<br />

possible to magnetize the machine if a zero torque command is applied [25]. Moreover,<br />

the dynamic performance is lower and torque oscillations are bigger than in another<br />

stator flux estimation method. Because of that many different stator flux estimation<br />

algorithms based on the voltage model were proposed, which does not approach to the<br />

pure integrator [15, 53, 54, 57, 58].<br />

Voltage Model with Low <strong>–</strong> Pass Filter (VM-LPF)<br />

The simplest method, which eliminates problems with initial conditions and dc drift,<br />

which appear in pure integrator, is a method with low-pass filter. In this case the<br />

equation (5.7) can be transformed as follows:<br />

dΨˆ<br />

dt<br />

s<br />

( Uˆ<br />

1<br />

s<br />

− R I<br />

s<br />

) Ψˆ<br />

s<br />

−<br />

s<br />

= (5.8)<br />

T<br />

F<br />

The block diagram of the method with low-pass filter is presented in Fig. 5.6.<br />

I s<br />

R s<br />

1<br />

T F<br />

U s<br />

1<br />

s<br />

Ψˆ<br />

s<br />

Fig. 5.6. Flux estimator based on voltage model with low-pass filter<br />

105


5. Estimation in Induction Motor Drives<br />

The estimator stabilization time depends on the low-pass filter time constant T F .<br />

Obviously, the low-pass filter produces some errors in phase angle and a magnitude of<br />

stator flux, especially when the motor frequency is lower than the cutoff frequency of<br />

the filter. Therefore, flux estimator with low-pass filter can be used successfully only in<br />

a limited speed range.<br />

Voltage Model with Compensated Low <strong>–</strong> Pass Filter (VM-CLPF)<br />

One way to overcome the errors introduced by low-pass filter is compensated<br />

algorithm [48]. The block diagram of flux estimator based on a voltage model with<br />

compensated low-pass filter is presented in Fig. 5.7.<br />

U s<br />

1 −<br />

jλsign(<br />

Ωˆ<br />

ss<br />

)<br />

s +<br />

1<br />

ˆ<br />

Ω ss<br />

λ<br />

Ψˆ<br />

s<br />

Ψˆ<br />

γˆ<br />

ss<br />

s<br />

ˆΩ ss<br />

s<br />

γˆ<br />

ss<br />

Fig. 5.7. Flux estimator based on voltage model with compensated low-pass filter<br />

In presented method the compensation is carried out before low-pass filtering. The<br />

stator flux is given by equation:<br />

Ψˆ<br />

E<br />

s<br />

s<br />

1−<br />

jλsign(<br />

Ωˆ<br />

=<br />

s + λ Ωˆ<br />

ss<br />

ss<br />

)<br />

(5.9)<br />

where: λ is a positive constant.<br />

The complex-valued gain, instead of calculating the phase error and the gain error, is<br />

used to compensation. Moreover, due to shifting the poles of pure integration from the<br />

origin to − λ Ωˆ<br />

ss<br />

, the drift problems are avoided. The λ factor can be selected in range<br />

from 0.1 to 0.5. For lower λ the transient performance is better, but a higher value of λ<br />

allows bigger system inexactness.<br />

106


5.3. Stator Flux <strong>Vector</strong> Estimators<br />

Voltage Model with Reference Flux (VM-RF)<br />

The block diagram of the estimator based on voltage model with reference flux is<br />

presented in Fig. 5.8 [25].<br />

U s<br />

I s<br />

s<br />

R s<br />

L s<br />

σ<br />

L<br />

L<br />

r<br />

M<br />

τ<br />

1+ sτ<br />

Ψˆ<br />

r<br />

L<br />

L<br />

M<br />

r<br />

I s<br />

L s<br />

σ<br />

Ψˆ<br />

s<br />

Ψ rc<br />

j<br />

e γˆ<br />

sr<br />

1<br />

1+ sτ<br />

γˆ<br />

sr<br />

Fig. 5.8. Flux estimator based on voltage model with rotor flux assumed as reference<br />

This estimator calculates rotor and stator flux vector on the basis of stator voltages<br />

and currents, and simultaneously the difference between reference and estimated rotor<br />

flux magnitude is utilizing to correction estimated values.<br />

In this estimator first a rotor flux vector is calculated based on the equation:<br />

dΨˆ<br />

dt<br />

r<br />

E<br />

+ K(<br />

Ψˆ<br />

=<br />

r<br />

r<br />

−Ψ rc<br />

e<br />

γ<br />

j ˆsr<br />

)<br />

(5.10)<br />

where K is the gain factor and<br />

E<br />

r<br />

is the rotor back EMF defined as:<br />

E<br />

r<br />

Lr<br />

dI<br />

s<br />

= ( U<br />

s<br />

− RsI<br />

s<br />

− σLs<br />

)<br />

(5.11)<br />

L<br />

dt<br />

m<br />

Then assuming<br />

1<br />

K = − the equation (5.10) can be rewritten yielding:<br />

τ<br />

ˆ<br />

τ<br />

1+<br />

sτ<br />

1<br />

+<br />

1+<br />

sτ<br />

j ˆ γ sr<br />

Ψ E Ψ rc<br />

e<br />

(5.12)<br />

= r<br />

r<br />

where:<br />

d<br />

s = (5.13)<br />

dt<br />

107


5. Estimation in Induction Motor Drives<br />

From the equation describing the IM in α − β coordinate system (2.15) formulas for<br />

calculation stator flux vector<br />

Ψ<br />

s<br />

are obtained.<br />

Ψ ˆ Lm<br />

s<br />

= Ψˆ<br />

r<br />

+ σLsI<br />

s<br />

L<br />

(5.14)<br />

r<br />

This estimator works correctly for a wide speed range, ensures good dynamic<br />

performance, eliminates influence of non correct initial values of the flux level.<br />

Moreover, in this algorithm rotor flux is calculated, which is necessary for rotor speed<br />

calculation (see section 5.5). It is important advantage of this estimator.<br />

The flux estimator based on voltage model with reference flux was selected for the<br />

implementation <strong>DTC</strong>-<strong>SVM</strong> control structure in sensorless operation mode (see section<br />

6.2). Presented algorithm is compromise between precision of rotor and stator flux<br />

estimation and computing demand.<br />

Current Model in Rotor Coordinated (CM-RC)<br />

The measured currents and mechanical speed are the input signals for the flux<br />

estimator based on the current model in rotor coordinate.<br />

Coordinate system<br />

d ′ − q′<br />

rotates with the angular speed of the motor shaft Ω<br />

m<br />

,<br />

which can be defined as follows:<br />

dγm<br />

Ωm = (5.15)<br />

dt<br />

Taking into consideration number of pole pairs<br />

system d ′ − q′<br />

is equal Ω<br />

K<br />

= pbΩm<br />

.<br />

p<br />

b<br />

angular speed of the coordinate<br />

The voltage, currents and fluxes complex space vector can be resolved into<br />

components d′ and q′ .<br />

U = j<br />

(5.16a)<br />

sK<br />

U<br />

sd ′ + U<br />

sq′<br />

I = j , I<br />

rK<br />

= I<br />

rd ′ + jI<br />

rq′<br />

(5.16b)<br />

sK<br />

I<br />

sd ′ + I<br />

sq′<br />

Ψ<br />

sK<br />

= Ψ<br />

sd′ + jΨ<br />

sq′<br />

,<br />

rK<br />

= Ψ<br />

rd ′ + jΨ<br />

rq′<br />

Ψ (5.16c)<br />

108


5.3. Stator Flux <strong>Vector</strong> Estimators<br />

The complete set of equations for IM (2.10-2.12) can be transformed to the<br />

d ′ − q′<br />

coordinate system. In this coordinate system the motor model equation can be written as<br />

follows:<br />

dΨ<br />

= (5.17a)<br />

dt<br />

sd′<br />

U<br />

sd′ RsI<br />

sd ′ + − pbΩmΨ<br />

sq′<br />

dΨ<br />

= (5.17b)<br />

dt<br />

sq′<br />

U<br />

sq′ RsI<br />

sq′<br />

+ + pbΩmΨ<br />

sd′<br />

dΨ<br />

rd ′<br />

0 = Rr<br />

I<br />

rd ′ +<br />

dt<br />

(5.17c)<br />

dΨ<br />

rq′<br />

0 = Rr<br />

I<br />

rq′ +<br />

dt<br />

(5.17d)<br />

Ψ<br />

sd<br />

′ LsI<br />

sd′<br />

+ LM<br />

I<br />

rd ′<br />

= (5.18a)<br />

Ψ<br />

sq′ LsI<br />

sq′<br />

+ LM<br />

I<br />

rq′<br />

= (5.18b)<br />

Ψ<br />

rd<br />

′ Lr<br />

I<br />

rd′<br />

+ LM<br />

I<br />

sd ′<br />

= (5.18c)<br />

Ψ<br />

rq′ Lr<br />

I<br />

rq′<br />

+ LM<br />

I<br />

sq′<br />

= (5.18d)<br />

dΩ<br />

dt<br />

m<br />

1 ⎡<br />

⎢ p<br />

J ⎣<br />

m<br />

2<br />

⎤<br />

( Ψ I −Ψ<br />

I ) − M ⎥ ⎦<br />

s<br />

=<br />

b sd ′ sq′<br />

sq′<br />

sd ′ L<br />

(5.19)<br />

From the equations (5.17-5.17) formulas for the estimated rotor flux can be obtained<br />

[66].<br />

dΨˆ<br />

dt<br />

dΨˆ<br />

dt<br />

rd<br />

′<br />

rq′<br />

r<br />

( L I ′ −Ψ<br />

′)<br />

1<br />

= ˆ<br />

M sd<br />

T<br />

r<br />

rd<br />

( L I ′ −Ψ<br />

′ )<br />

1<br />

= ˆ<br />

M sq<br />

T<br />

rq<br />

(5.20a)<br />

(5.20b)<br />

where:<br />

T =<br />

r<br />

L<br />

R<br />

r<br />

r<br />

The current vector is measured in stationary coordinate α − β . Therefore, current<br />

components<br />

I<br />

sα<br />

,<br />

sβ<br />

I must be transformed to the system d ′ − q′<br />

. Similarly, the<br />

estimated rotor flux vector<br />

Ψ<br />

r<br />

, must be transformed from the system<br />

d ′ − q′<br />

to α − β .<br />

109


5. Estimation in Induction Motor Drives<br />

Stator flux vector<br />

Ψ<br />

s<br />

is calculated from the equation (5.14).<br />

Block diagram of the whole stator flux estimator is shown in Fig. 5.9.<br />

I sα<br />

L s<br />

σ<br />

I sα<br />

α − β<br />

I<br />

sd ′<br />

LM<br />

1<br />

T r<br />

∫<br />

Ψ ˆ<br />

rd ′<br />

d ′ − q′<br />

Ψˆ rα<br />

L<br />

L<br />

M<br />

r<br />

Ψˆ sα<br />

I sβ<br />

I<br />

sq ′<br />

∫<br />

Ψ ˆ<br />

rq ′<br />

1<br />

L M<br />

d ′ − q′<br />

T r<br />

α − β<br />

Ψˆ rβ<br />

L<br />

L<br />

M<br />

r<br />

Ψˆ sβ<br />

γ m<br />

I sβ<br />

L s<br />

σ<br />

Fig. 5.9. Block diagram of the current model flux estimator in rotor coordinates<br />

This flux estimator model ensures good accuracy over the entire frequency range. It<br />

has a very good behavior in steady and dynamic state. Also it has resistant to wrong<br />

initial conditions. Its disadvantage is sensitive on change motor parameters.<br />

This estimator was selected for the implementation <strong>DTC</strong>-<strong>SVM</strong> control structure in<br />

sensor operation mode (see section 6.2).<br />

5.4. <strong>Torque</strong> Estimation<br />

The induction motor output torque is calculated based on the equation (2.9), which<br />

for stationary coordinate system α − β can be written as follows:<br />

M<br />

e<br />

m<br />

2<br />

s<br />

= pb<br />

Im<br />

s<br />

Is<br />

( ˆ * ms<br />

) = p ( Ψˆ<br />

I −Ψˆ<br />

I )<br />

Ψ<br />

b sα<br />

sβ<br />

sβ<br />

sα<br />

(5.21)<br />

2<br />

It can be seen that the calculated torque is depended on the current measurement<br />

accuracy and stator flux estimation method.<br />

5.5. Rotor Speed Estimation<br />

If a flux estimator works properly and rotor flux is accurately calculated mechanical<br />

speed can be obtained from simple motor model equation [87]. If in control structure the<br />

110


5.5. Rotor Speed Estimation<br />

stator flux estimator is applied rotor flux can be calculated based on the equations<br />

(5.14).<br />

In the IM mechanical speed is defined as difference between synchronous speed and<br />

sleep frequency:<br />

Ω<br />

m<br />

b<br />

( Ω − Ω )<br />

= 1 sr sl<br />

(5.22)<br />

p<br />

where:<br />

Ω<br />

sr<br />

- rotor synchronous speed,<br />

Ω<br />

sl<br />

- slip frequency,<br />

p<br />

b<br />

- number of pole pairs.<br />

The rotor synchronous speed is equal angular speed of the rotor flux vector and can<br />

be calculated as:<br />

Ω<br />

sr<br />

dγ<br />

sr<br />

= (5.23)<br />

dt<br />

The slip frequency of induction motor is defined as follows [66]:<br />

Ω<br />

sl<br />

= Ω − p Ω<br />

(5.24)<br />

sr<br />

b<br />

m<br />

Based on the equations (3.3d) and (3.4d) in rotor flux coordinate system the slip<br />

frequency can be expressed:<br />

Ω<br />

sl<br />

L<br />

1<br />

M<br />

= Rr<br />

Isq<br />

(5.25)<br />

Lr<br />

Ψ<br />

r<br />

Taking into consideration the torque equations (3.7) and (5.25) the estimated sleep<br />

frequency can be calculated as follows:<br />

Ω<br />

sl<br />

r<br />

ˆ 2<br />

r<br />

( Ψˆ<br />

I −Ψˆ<br />

I )<br />

R<br />

=<br />

sα<br />

sβ<br />

sβ<br />

sα<br />

(5.26)<br />

Ψ<br />

Finally mechanical motor speed is calculated from the equation (5.22).<br />

111


5. Estimation in Induction Motor Drives<br />

5.6. Summary<br />

In this chapter estimation algorithms of flux, torque and rotor speed are presented.<br />

The estimators provide feedback signals for <strong>DTC</strong>-<strong>SVM</strong> control scheme. Algorithms<br />

selected to the implementation in final structure are described and discussed.<br />

The speed estimator is based on the estimated stator and rotor fluxes. The mechanical<br />

speed can be calculated in a simple way if motor flux is properly estimated. Therefore,<br />

flux estimation algorithm is the most important part of sensorless control scheme.<br />

Selected flux estimator for the sensorless mode is based on the voltage model. Thus<br />

algorithm is sensitive on accuracy of inverter output voltage calculation. The voltages<br />

are reconstructed from switching signals. In this method dead-time compensation<br />

algorithm is significant. The dead-time effect and compensation algorithm was<br />

presented.<br />

The presented estimation methods are implemented in final <strong>DTC</strong>-<strong>SVM</strong> control<br />

structure. The experimental results, presented in Chapter 7 confirm proper operation of<br />

selected estimation methods.<br />

112


6. Configuration of the Developed IM Drive Based on<br />

<strong>DTC</strong>-<strong>SVM</strong><br />

6.1. Introduction<br />

In this chapter a whole implemented control system will be presented. In the first<br />

part, the configuration of the system and operation modes are described. In the next<br />

parts, two hardware setups, which were used to verify <strong>DTC</strong>-<strong>SVM</strong> control structure are<br />

presented. To development work was used laboratory setup based on dSPACE company<br />

control board DS1103 PPC. This board has powerful microprocessor and special inputoutput<br />

interface. The laboratory setup and control board DS1103 will be widely<br />

described in section 6.3. The control algorithm was also implemented in a setup based<br />

on a microcontroller TMS320LF2406 from Texas Instruments company. The<br />

TMS320LF2406 is a 16-bits, fixed point microcontroller devoted for drive application<br />

(see section 6.4).<br />

6.2. Block Scheme of Implemented Control System<br />

The IM drive based on <strong>DTC</strong>-<strong>SVM</strong> control structure can operate in three modes:<br />

• scalar control,<br />

• sensor vector control,<br />

• sensorless vector control.<br />

The inverter operate in a mode which is required by application. The system<br />

configuration depends on the switches position, see Fig. 6.1. The most advanced is the<br />

sensorless vector control mode.<br />

In the scalar control mode algorithm obtains command voltage vector based on the<br />

reference frequency. The command voltage vector is realized by space vector modulator<br />

(<strong>SVM</strong>).<br />

The reference speed in the command signal in the vector control modes. Depending<br />

on mode the reference speed is compared with measured (sensor vector control mode)<br />

or estimated (sensorless vector control mode) speed signal.


6. Configuration of the Developed IM Drive Based on <strong>DTC</strong>-<strong>SVM</strong><br />

Reference<br />

Frequency<br />

Scalar<br />

Control<br />

Switch 1<br />

Reference<br />

Speed<br />

Speed<br />

Controller<br />

References<br />

Value<br />

<strong>Torque</strong><br />

and Flux<br />

Controller<br />

<strong>SVM</strong><br />

Measurements<br />

Signals<br />

Inverter<br />

Estimations<br />

Value<br />

Switch 2<br />

Estimation<br />

Speed<br />

Measurment<br />

Speed<br />

<strong>Torque</strong><br />

and Flux<br />

Estimator<br />

Speed<br />

Estimator<br />

Speed<br />

Sensor<br />

Motor<br />

Fig. 6.1. Block scheme of implemented control algorithm<br />

Based on the speed error speed controller calculates reference torque value. The<br />

commanded flux is obtained from the reference speed and selected characteristic, which<br />

depends on the application. The reference values of torque and flux are compared with<br />

estimated values. Based on the errors flux and torque controllers calculate command<br />

voltage vector. The command voltage vector is realized by the same space vector<br />

modulator (<strong>SVM</strong>) algorithm, which is used in scalar control mode. Therefore, depended<br />

on application requirements change between scalar and vector mode is simple.<br />

The measured current and reconstructed voltage are input signals for the estimation<br />

algorithms (see Chapter 5).<br />

An inverter control structure presented in Fig. 6.1 was implemented for IM.<br />

However, this structure can be also used for Permanent Magnet Synchronous Motor<br />

(PMSM) [129].<br />

All presented in Fig. 6.1 blocks are described in previous chapter of the thesis. The<br />

torque, flux and speed controllers are discussed in Chapter 4. The estimation algorithms<br />

are shown in Chapter 5 and different modulation techniques are presented in Chapter 2.<br />

The experimental results for all three operating modes are presented in Chapter 7.<br />

114


6.3. Laboratory Setup Based on DS1103<br />

6.3. Laboratory Setup Based on DS1103<br />

The basic structure of the laboratory setup is depicted in Fig. 6.1. The motor setup<br />

consist of induction motor and DC motor, which is used for the loading. The induction<br />

motor is fed by the frequency inverter controlled directly by the DS1103 board. The<br />

dSPACE DS1103 PPC is plugged in the host PC. The DC motor is supplied by a torque<br />

controlled rectifier. The encoder is used for the measure mechanical speed. The DSP<br />

Interface <strong>–</strong> a set of eurocards mounted in a 19” rack with the main purpose to provide<br />

galvanic isolation to all signals connected to the DS1103 PPC controller.<br />

grid<br />

3 2<br />

3<br />

Rectifier<br />

Inverter<br />

Rectifier<br />

measured<br />

DC line voltage<br />

S A<br />

S B<br />

S C<br />

measured<br />

phase<br />

current<br />

Measurement<br />

PC<br />

DSP<br />

Interface<br />

DS1103 dSPACE<br />

Master : PowerPC 604e<br />

Slave: DSP TMS320F240<br />

encoder<br />

AC motor<br />

DC motor<br />

Fig. 6.2. Structure of the laboratory setup<br />

Fig. 6.3. Laboratory setup<br />

115


6. Configuration of the Developed IM Drive Based on <strong>DTC</strong>-<strong>SVM</strong><br />

In Fig. 6.3 view of the laboratory setup is shown. All parts of the laboratory setup<br />

can be seen in this picture.<br />

dSPACE DS1103 PPC Board<br />

The dSPACE DS1103 PPC is a mixed RISC/DSP digital controller providing a very<br />

powerful processor for floating point calculations as well as comprehensive I/O<br />

capabilities. Here are the most relevant features of the controller:<br />

• Motorola PowerPC 604e running at 333 MHz,<br />

• Slave DSP TI's TMS320F240 Subsystem,<br />

• 16 channels (4 x 4ch) ADC, 16 bit , 4 µs, ±10 V,<br />

• 4 channels ADC, 12 bit , 800 ns, ± 10V,<br />

• 8 channels (2 x 4ch) DAC, 14 bit , ±10 V,6 µs,<br />

• Incremental Encoder Interface -7 channels<br />

• 32 digital I/O lines, programmable in 8-bit groups,<br />

• Software development tools (Matlab/Simulink, RTI, RTW, TDE, Control Desk)<br />

The DS1103 PPC card is pluged in one of the ISA slot of the motherboard of a host<br />

computer of the type PIII/900MHz, 512 MBRAM, 40GB HDD, Windows 2000. All the<br />

connections are made through six flat cables (50 wires each) available at the backside of<br />

the desktop computer.<br />

The DS1103 PPC is a very flexible and powerful system featuring both high<br />

computational capability and comprenhensive I/O periphery. The board can be<br />

programmed in C language. Additionally, it features a software SIMULINK interface<br />

that allows all applications to be developed in the Matlab/Simulink user friendly<br />

environment. All compiling and downloading processes are carried out automatically in<br />

the background. An experimenting software called Control Desk, allow real-time<br />

management of the running process by providing a virtual control panel with<br />

instruments and scopes.<br />

The detailed parameters of the dSPACE DS1103 PPC board are given in Appendix<br />

A5.<br />

116


6.3. Laboratory Setup Based on DS1103<br />

Experimenting Software <strong>–</strong> Control Desk<br />

Control Desk experiment software provides all the functions for controlling,<br />

monitoring, and automation of real-time experiments and makes the development of<br />

controllers more effective. A Control Desk experiment layout for controlling an<br />

induction motor with <strong>DTC</strong>-<strong>SVM</strong> control methods is shown in Fig. 6.5.<br />

Fig. 6.4. Control Desk experiment layout<br />

Control Desk package consists of the following modules:<br />

• The Experiment Management - assures a consistent data management controlling<br />

all the data relevant for an experiment. The experiment can be loaded as a<br />

complete set of data with a single operation. The content of the experiment can<br />

be defined by the user.<br />

• The Hardware Management - allows you to configure the dSPACE hardware and<br />

to handle real-time applications with a graphical user interface.<br />

• The Instrumentation Kits - offer a variety of virtual instruments to build and<br />

configure virtual instrument panels according to your special needs.<br />

117


6. Configuration of the Developed IM Drive Based on <strong>DTC</strong>-<strong>SVM</strong><br />

Using data acquisition instruments you can capture data from the model running on<br />

the real-time hardware. Changing parameter values is performed by operating input<br />

instruments. The integrated Parameter Editor allows you to read the current parameter<br />

values from the hardware and to change a parameter set in one step.<br />

6.4. Drive Based on TMS320LF2406<br />

<strong>DTC</strong>-<strong>SVM</strong> control algorithm was implemented in the drive based on microcontroller<br />

TMS320LF2406. Setup consists of 18 kVA IGBT inverter and 15 kW induction motor.<br />

The view of inverter is shown in Fig. 6.5. In this picture main control board of the<br />

inverter with microprocessor module can be seen.<br />

Fig. 6.5. 18 kVA inverter controlled by TMS320FL2406 processor<br />

118


6.4. Drive Based on TMS320LF2406<br />

The motor set (Fig. 6.6), which was used in tests consists of 15 kW induction motor<br />

and 22 kW DC motor. The induction motor data are given in appendix A.3. The DC<br />

motor works as a load and it is supply from the controlled rectifier.<br />

Fig. 6.6. Motor set. From the left 22 kW DC motor and 15 kW IM motor.<br />

Fig. 6.7. TMS320LF2406 microprocessor board<br />

119


6. Configuration of the Developed IM Drive Based on <strong>DTC</strong>-<strong>SVM</strong><br />

The microprocessor board shown in the Fig. 6.7 was used to control the inverter. The<br />

sizes of the processor module are 53x56mm. This board contains microcontroller<br />

TMS320LF2406 and required equipment. The communication with main inverter board<br />

by three connectors (2x20pins and 1x26pins) is provided.<br />

The TMS320Lx240xA series of devices are members of the TMS320 family of<br />

digital signal processors (DSPs) designed to meet a wide range of digital motor control<br />

(DMC) and other embedded control applications [99, 100]. This series is based on the<br />

C2xLP 16-bit, fixed-point, low-power DSP CPU, and is complemented with a wide<br />

range of on-chip peripherals and on-chip ROM or flash program memory, plus on-chip<br />

dual-access RAM (DARAM).<br />

The TMS320 family consists of fixed-point, floating-point, multiprocessor digital<br />

signal processors (DSPs), and fixed-point DSP controllers. TMS320 DSPs have an<br />

architecture designed specifically for real-time signal processing. The 240xA series of<br />

DSP controllers combine this real-time processing capability with controller peripherals<br />

to create an ideal solution for control system applications. There are short characteristics<br />

of the TMS320 family:<br />

• flexible instruction set,<br />

• operational flexibility,<br />

• high-speed performance<br />

• Innovative parallel architecture,<br />

• cost effectiveness.<br />

Devices within a generation of a TMS320 platform have the same CPU structure but<br />

different on-chip memory and peripheral configurations. Spin-off devices use new<br />

combinations of on-chip memory and peripherals to satisfy a wide range of needs in the<br />

worldwide electronics market. By integrating memory and peripherals onto a single<br />

chip, TMS320 devices reduce system costs and save circuit board space.<br />

The detailed parameters of the TMS320FL2406 microprocessor are given in<br />

Appendix A6.<br />

The important feature of the TMS320FL246 microprocessor is the bootloader.<br />

Thanks to that it is possible to program the device using Serial Communications<br />

120


6.4. Drive Based on TMS320LF2406<br />

Interface (SCI) or Serial Peripheral Interface (SPI). Therefore, program can be loaded<br />

from the PC via standard serial port (RS232).<br />

This way of programming was used during the implementation of <strong>DTC</strong>-<strong>SVM</strong> control<br />

algorithm. Thus it was possible to work with the processor without using the expensive<br />

tools like JTAG.<br />

121


7. Experimental Results<br />

7.1. Introduction<br />

In this chapter selected experimental results obtained in the system described in<br />

Chapter 6 are shown. All tests was done for 3 kW induction motor, which parameters<br />

are given in Appendix A3.<br />

7.2. Pulse Width Modulation<br />

In Fig. 7.1 <strong>–</strong> 7.5 different modulation method are presented. All test was measured at<br />

frequency f = 40 Hz .<br />

In Fig. 7.1 space vector modulation method with symmetrical zero vectors placement<br />

<strong>–</strong> SVPWM is shown (see section 2.4.3).<br />

Fig. 7.1. <strong>Space</strong> vector modulation (SVPWM) at frequency f = 40 Hz 1) switching signal S A ,<br />

2) pole voltage U A0 (150 V/div), 3) phase voltage U A (150 V/div), 4) output current I A (5 A/div)<br />

In Fig. 7.2 discontinuous pulse width modulation <strong>–</strong> DPWM2 is shown (see section<br />

2.4.3). It can be observe differences in pole voltage waveforms and switching signal in<br />

Fig. 7.1 and 7.2. DPWM2 modulation method has 60º no switch sectors. However,<br />

phase voltage and output current have sinusoidal waveforms.


7.2. Pulse Width Modulation<br />

Fig. 7.2. Discontinuous modulation (DPWM2) at frequency f = 40 Hz 1) switching signal S A ,<br />

2) pole voltage U A0 (150 V/div), 3) phase voltage U A (150 V/div), 4) output current I A (5 A/div)<br />

In Fig. 7.3 and 7.4 overmodulation (OM) algorithm is shown (see section 2.4.5).<br />

Fig. 7.3. Overmodulation mode I at frequency f = 40 Hz 1) switching signal S A , 2) pole voltage<br />

U A0 (150 V/div), 3) phase voltage U A (150 V/div), 4) output current I A (5 A/div)<br />

123


7. Experimental Results<br />

Fig. 7.4. Overmodulation mode II at frequency f = 40 Hz 1) switching signal S A , 2) pole voltage<br />

U A0 (150 V/div), 3) phase voltage U A (150 V/div), 4) output current I A (5 A/div)<br />

The results for six-step mode are presented in Fig. 7.5.<br />

Fig. 7.5. Six-step mode at frequency f = 40 Hz 1) switching signal S A , 2) pole voltage U A0 (150 V/div),<br />

3) phase voltage U A (150 V/div), 4) output current I A (10 A/div)<br />

Results presented in Fig. 7.3 <strong>–</strong> 7.5 ware obtained at decreased dc-link voltage.<br />

Therefore, overmodulation and six-step operation modes can be shown with frequency<br />

124


7.3. Flux and <strong>Torque</strong> Controllers<br />

f = 40 Hz like the other results. Thanks to it, current and voltage waveforms can be<br />

better compared.<br />

Experimental results presented in Fig. 7.1 <strong>–</strong> 7.5 confirm proper operation all type<br />

modulation algorithms.<br />

7.3. Flux and <strong>Torque</strong> Controllers<br />

Dynamic tests for the flux and torque controller were done for different sampling<br />

frequencies values and the same condition like for simulation presented in section 4.3<br />

(motor speed Ω = 0 ). The flux controller parameters were calculated according to<br />

m<br />

symmetric optimum criterion (see section 4.3.1) and torque controller parameters were<br />

calculated according to root locus method (see section 4.3.2).<br />

In Fig. 7.6 <strong>–</strong> 7.8 are presented stator flux step response at sampling frequency<br />

f s<br />

= 10 kHz , f s<br />

= 5 kHz , f s<br />

= 2. 5 kHz respectively. Those results can be compared<br />

with simulation results presented in Fig. 4.11.<br />

Fig. 7.6. Stator flux response at sampling frequency f s = 10 kHz 1) reference flux (0.15 Wb/div),<br />

2) estimated flux (0.15 Wb/div)<br />

125


7. Experimental Results<br />

Fig. 7.7. Stator flux response at sampling frequency f s = 5 kHz 1) reference flux (0.15 Wb/div),<br />

2) estimated flux (0.15 Wb/div)<br />

Fig. 7.8. Stator flux response at sampling frequency f s = 2. 5 kHz 1) reference flux (0.15 Wb/div),<br />

2) estimated flux (0.15 Wb/div)<br />

Presented in Fig. 7.6 <strong>–</strong> 7.8 experimental results confirm proper operation of the flux<br />

control loop at different sampling frequency.<br />

126


7.3. Flux and <strong>Torque</strong> Controllers<br />

The experimental results of torque controller dynamic test are shown in Fig. 7.9 <strong>–</strong><br />

7.11. Presented results were obtain at sampling frequency f s<br />

= 10 kHz (Fig. 7.9),<br />

f s<br />

= 5 kHz (Fig. 7.10), f s<br />

= 2. 5 kHz (Fig. 7.11).<br />

Fig. 7.9. <strong>Torque</strong> response at sampling frequency f s = 10 kHz 1) reference torque (4.5 Nm/div),<br />

3) estimated torque (4.5 Nm/div)<br />

Fig. 7.10. <strong>Torque</strong> response at sampling frequency f s = 5 kHz 1) reference torque (4.5 Nm/div),<br />

3) estimated torque (4.5 Nm/div)<br />

127


7. Experimental Results<br />

Fig. 7.11. <strong>Torque</strong> response at sampling frequency f s = 2. 5 kHz 1) reference torque (4.5 Nm/div),<br />

3) estimated torque (4.5 Nm/div)<br />

The result from Fig. 7.9 <strong>–</strong> 7.11 can be compared with simulation results presented in<br />

Fig. 4.30. Experimental results presented in Fig. 7.9 <strong>–</strong> 7.11 confirm proper operation of<br />

the torque control loop at different sampling frequency.<br />

The decoupling between flux and torque control loops is presented in Fig. 7.12. The<br />

torque step response (Fig. 7.12a) and magnitude stator flux step response (Fig. 7.12b)<br />

are shown.<br />

a)<br />

128


7.4. <strong>DTC</strong>-<strong>SVM</strong> Control System<br />

b)<br />

Fig. 7.12. Dynamic tests a) torque step change, b) flux step change<br />

1) reference torque (9 Nm/div), 2) estimated torque (9 Nm/div),<br />

3) reference flux (0.3 Wb/div), 4) estimated flux (0.3 Wb/div)<br />

The results from Fig. 7.12 can be compared with simulation results presented in Fig.<br />

4.29. From Fig. 7.12 can be seen that decoupling between flux and torque is correct.<br />

7.4. <strong>DTC</strong>-<strong>SVM</strong> Control System<br />

In this section the experimental result for three possible drive operation modes,<br />

which are described in Chapter 6 are shown. Therefore, comparison of a system<br />

behavior in different modes is possible.<br />

In Fig. 7.13 <strong>–</strong> 7.16 results for scalar control mode are presented. Fig. 7.13 gives<br />

result for system startup to frequency<br />

f = 40Hz<br />

(motor speed Ω m<br />

= 125rad<br />

/ s ).<br />

129


7. Experimental Results<br />

Fig. 7.13. Scalar control mode - Startup from 0 to f = 40Hz<br />

1) reference frequency (25 Hz/div),<br />

2) actual speed (30 (rad/s)/div, 4) phase current (10 A/div)<br />

The load torque step change at frequency<br />

f = 25Hz<br />

is shown in Fig. 7.14.<br />

Fig. 7.14. Scalar control mode - Load torque step change from 0 to M L = M N at frequency f = 25Hz<br />

1) reference frequency (25 Hz/div), 2) actual speed (30 (rad/s)/div), 3) torque (20 Nm/div),<br />

4) phase current (10 A/div)<br />

In Fig. 7.15 and 7.16 result of speed reverses are shown (<br />

f = ± 25Hz<br />

). The reverse<br />

time is 0.5s (Fig. 7.15) and 5s (Fig. 7.16).<br />

130


7.4. <strong>DTC</strong>-<strong>SVM</strong> Control System<br />

Fig. 7.15. Scalar control mode - Speed reversal f = ± 25Hz<br />

(reverse time 0.5s) 1) reference frequency<br />

(25 Hz/div), 2) actual speed (30 (rad/s)/div), 4) phase current (10 A/div)<br />

Fig. 7.16. Scalar control mode - Speed reversal f = ± 25Hz<br />

(reverse time 5s) 1) reference frequency<br />

(25 Hz/div), 2) actual speed (30 (rad/s)/div), 4) phase current (10 A/div)<br />

In Fig. 7.17 <strong>–</strong> 7.20 results for sensor vector control mode are presented. Fig. 7.17<br />

gives result for system startup to speed<br />

Ω m<br />

= 120 rad / s .<br />

131


7. Experimental Results<br />

Fig. 7.17. <strong>Vector</strong> control mode with speed sensor - Startup from 0 to Ω m = 120 rad / s 1) reference speed<br />

(30 (rad/s)/div), 2) actual speed (30 (rad/s)/div, 4) phase current (10 A/div)<br />

The load torque step change at speed<br />

Ω m<br />

= 75 rad / s is shown in Fig. 7.18.<br />

Fig. 7.18. <strong>Vector</strong> control mode with speed sensor - Load torque step change from 0 to<br />

M = M at<br />

speed Ω m = 75 rad / s 1) reference speed (30 (rad/s)/div), 2) actual speed (30 (rad/s)/div),<br />

3) torque (20 Nm/div), 4) phase current (10 A/div)<br />

L<br />

N<br />

In Fig. 7.19 and 7.20 result of speed reverses are shown ( Ω m<br />

= ±75rad<br />

/ s ). The<br />

reverse time is 0.5s (Fig. 7.19) and 5s (Fig. 7.20).<br />

132


7.4. <strong>DTC</strong>-<strong>SVM</strong> Control System<br />

Fig. 7.19. <strong>Vector</strong> control mode with speed sensor - Speed reversal Ω m<br />

= ±75rad<br />

/ s (reverse time 0.5s)<br />

1) reference speed (30 (rad/s)/div), 2) actual speed (30 (rad/s)/div), 4) phase current (10 A/div)<br />

Fig. 7.20. <strong>Vector</strong> control mode with speed sensor - Speed reversal Ω m<br />

= ±75rad<br />

/ s (reverse time 5s)<br />

1) reference speed (30 (rad/s)/div), 2) actual speed (30 (rad/s)/div), 4) phase current (10 A/div)<br />

In sensorless vector control mode the accuracy of the speed estimation algorithm<br />

is important. Therefore, static and dynamic error of estimated speed were<br />

investigated. The error of estimated speed can be written as:<br />

133


7. Experimental Results<br />

εΩ<br />

m<br />

Ω ˆ<br />

m<br />

− Ωm<br />

= 100%<br />

(7.1)<br />

Ω<br />

m<br />

where:<br />

Ω<br />

m<br />

- actual speed,<br />

Ωˆ m<br />

- estimated speed.<br />

In Fig. 7.21 speed estimation error as the function of mechanical speed in steady<br />

state is presented.<br />

ε<br />

m<br />

[%]<br />

Ω<br />

50<br />

45<br />

40<br />

35<br />

error_omega [%]<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 5 10 15 20 25 30 35 40 45 50<br />

omega_m [rad/s]<br />

[rad/s] Ω m<br />

Fig. 7.21. Estimated speed error as the function of mechanical speed in steady state.<br />

The results of speed estimator dynamic test are presented in Fig. 22. In this test speed<br />

controller operates with the sensor and speed estimator work in open loop fashion.<br />

134


7.4. <strong>DTC</strong>-<strong>SVM</strong> Control System<br />

Fig. 7.22. Dynamic test of the speed estimation - Speed reversal Ω m = ±50rad<br />

/ s 1) reference speed<br />

(30 (rad/s)/div), 2) actual speed (30 (rad/s)/div), 3) estimated speed (30 (rad/s)/div),<br />

4) error of estimated speed (25 %/div)<br />

In Fig. 7.23 <strong>–</strong> 7.26 results for sensorless vector control mode are presented. Fig. 7.23<br />

gives result for system startup to speed<br />

Ω m<br />

= 120 rad / s .<br />

Fig. 7.23. Sensorless vector control mode - Startup from 0 to Ω m = 120 rad / s 1) reference speed<br />

(30 (rad/s)/div), 2) actual speed (30 (rad/s)/div, 4) phase current (10 A/div)<br />

The load torque step change at speed<br />

Ω m<br />

= 75 rad / s is shown in Fig. 7.24.<br />

135


7. Experimental Results<br />

Fig. 7.24. Sensorless vector control mode - Load torque step change from 0 to M = M at speed<br />

Ω m = 75 rad / s 1) reference speed (30 (rad/s)/div), 2) actual speed (30 (rad/s)/div),<br />

3) torque (20 Nm/div), 4) phase current (10 A/div)<br />

L<br />

N<br />

In Fig. 7.25 and 7.26 result of speed reverses are shown ( Ω m<br />

= ±75rad<br />

/ s ). The<br />

reverse time is 0.5s (Fig. 7.25) and 5s (Fig. 7.26).<br />

Fig. 7.25. Sensorless vector control mode - Speed reverse Ω m<br />

= ±75rad<br />

/ s (reverse time 0.5s)<br />

1) reference speed (30 (rad/s)/div), 2) actual speed (30 (rad/s)/div), 4) phase current (10 A/div)<br />

136


7.4. <strong>DTC</strong>-<strong>SVM</strong> Control System<br />

Fig. 7.26. Sensorless vector control mode - Speed reverse Ω m<br />

= ±75rad<br />

/ s (reverse time 5s)<br />

1) reference speed (30 (rad/s)/div), 2) actual speed (30 (rad/s)/div), 4) phase current (10 A/div)<br />

137


8. Summary and Conclusions<br />

In this thesis the most convenient industrial control scheme for voltage source<br />

inverter-fed induction motor drives was searched for, based on the existing control<br />

methods. This method should provide: operation in wide power range, guarantee good<br />

and repeatable parameters of drive. It is required by a serial production of a drive. To<br />

achieve a low costs the control system should be implemented in simple<br />

microprocessor. The analysis of existing methods were done in order to chose the<br />

industrial oriented universal scheme.<br />

The most important control techniques of IM were presented in Chapter 3: Field<br />

Oriented Control (FOC), Feedback Linearization Control (FLC) and <strong>Direct</strong> <strong>Torque</strong><br />

Control (<strong>DTC</strong>). The FLC structure guarantees exact decoupling of the motor speed and<br />

rotor flux control in both dynamic and steady states. However, it is complicated and<br />

difficult to implement in practice. This method requires complex computation and<br />

additionally it is sensitive to changes of motor parameters. Because of these features<br />

this method was not chosen for implementation. In next step FOC and <strong>DTC</strong> methods<br />

were analyzed. Characteristics of those methods were done on the basis of the literature,<br />

simulation and experimental investigation. The conclusions of those consideration were<br />

shown in section 3.5.<br />

Analysis of advantages and disadvantages of FOC and <strong>DTC</strong> methods resulted in a<br />

search for method which will eliminate disadvantages and keep advantages of those<br />

methods. The direct torque control with space vector modulation (<strong>DTC</strong>-<strong>SVM</strong>) is an<br />

effect of this search. The main features of this method can be summarized as:<br />

• <strong>Space</strong> vector modulator,<br />

• Constant switching frequency,<br />

• Unipolar voltage thanks to use of PWM block (<strong>SVM</strong>),<br />

• Sinusoidal waveform of stator currents,<br />

• Algorithm operates with torque and flux value <strong>–</strong> implementation in<br />

manufacturing process is easier,<br />

• Good dynamic control of flux and torque. The step responses are slower than in<br />

classical <strong>DTC</strong>, because PI controllers are slower than hysteresis controllers,


8. Summary and Conclusions<br />

which are used in classical <strong>DTC</strong>. However, obtained dynamic (response time for<br />

the torque 1.5-2ms) is sufficient for general purpose drives.<br />

• High sampling frequency is not required. The <strong>DTC</strong>-<strong>SVM</strong> algorithm works<br />

properly at sampling frequency<br />

frequency at least<br />

25 − 40kHz<br />

.<br />

f s<br />

= 5kHz<br />

whereas <strong>DTC</strong> requires sampling<br />

• Low flux and torque ripple than in classical <strong>DTC</strong>. The torque ripples in <strong>DTC</strong>-<strong>SVM</strong><br />

at sampling frequency<br />

f s<br />

= 5kHz<br />

are ten times lower than presented in section<br />

3.4.2 torque ripples for classical <strong>DTC</strong> at sampling frequency f s<br />

= 40kHz<br />

.<br />

The <strong>DTC</strong>-<strong>SVM</strong> scheme is based only on the analysis of stator equations like classical<br />

<strong>DTC</strong>, therefore control algorithm is not sensitive to rotor parameters changes. This<br />

method can be applied also for surface mounted permanent magnet (PM) synchronous<br />

motors [129]. The PM synchronous motors of this type are more frequently used in<br />

standard speed drives as interior PM. Hence, <strong>DTC</strong>-<strong>SVM</strong> method allows universal drive<br />

building for both types of AC motors.<br />

The very important part of <strong>DTC</strong>-<strong>SVM</strong> scheme is a space vector modulator. The<br />

different modulation techniques can be applied in the system. Therefore, a drive has<br />

additional advantages. The most important is full range of voltage control and reduction<br />

of switching losses. For instance, reduction of switching losses can be obtained by<br />

implementation of discontinuous PWM methods. These modulation techniques were<br />

described and characterized in section 2.4. The experimental results for the<br />

implemented modulation methods were shown in Chapter 7.<br />

The short review of <strong>DTC</strong>-<strong>SVM</strong> methods proposed in literature were given in section<br />

4.2. For further consideration the <strong>DTC</strong>-<strong>SVM</strong> method with close-loop torque and flux<br />

control in stator flux Cartesian coordinates have been chosen. In author opinion this<br />

method is best suited for commercial manufactured drives. For chosen scheme two<br />

controller design procedures were proposed. Those analysis were presented in Chapter 4.<br />

Also correction of controllers parameters for sampling frequency changes was discussed.<br />

In adjustable speed drive superior speed controller is used. The analysis of speed<br />

control loop and controller tuning were presented in section 4.4. Correctness of used<br />

method was confirmed by simulation and experimental results.<br />

139


8. Summary and Conclusions<br />

The quality of regulation process depends on an accuracy of feedback signals. In the<br />

vector control of induction motor those signals are provided by flux and torque<br />

estimators and, in sensorless operation mode, by a speed estimator. The precision of<br />

estimated signals depends on:<br />

• exact knowledge of motor parameters,<br />

• good dead-time and voltage drop compensation algorithms,<br />

• well realized measurements,<br />

• implementation of on-line adaptation of motor parameters.<br />

Those features are common for all vector control methods. Therefore, if feedback<br />

signals are estimated accurately, the control scheme should be as simple as possible.<br />

The <strong>DTC</strong>-<strong>SVM</strong> has a simple structure and it can be analyzed and implemented in a<br />

simple way. It is very important feature of <strong>DTC</strong>-<strong>SVM</strong>.<br />

Estimation problems in a drive with induction motor were discussed in Chapter 5.<br />

Following estimation algorithms, selected for implementation, were presented: voltage<br />

estimator with dead-time compensation algorithm, stator flux estimator, torque<br />

estimator and mechanical speed estimator.<br />

All parts of control scheme were verified in simulation and experiment. The whole<br />

scheme consists of: flux and torque controllers, speed controller, estimation of flux,<br />

torque and speed and compensation algorithms. Those complete structure was presented<br />

in Chapter 6. Proposed solution was implemented in 3 kW experimental and 15 kW<br />

industrial drives. The laboratory setups were also presented in Chapter 6.<br />

Presented in Chapter 7 experimental results confirm proper operation of developed<br />

control system.<br />

Thus, thesis shows the process to select and develop the most convenient control<br />

scheme for voltage source inverter-fed induction motor drives. Whole problems of<br />

direct flux and torque control with space vector modulation (<strong>DTC</strong>-<strong>SVM</strong>) were analyzed<br />

and investigated in simulation and experiment.<br />

Finally, it should be stressed that the developed system was brought into serial<br />

production. Presented algorithm has been used in new family of inverter drives<br />

produced by Polish company Power Electronic Manufacture <strong>–</strong> „TWERD”, Toruń.<br />

140


References<br />

[1] V. Ambrozic, G.S. Buja, R. Menis, "Band-Constrained Technique for <strong>Direct</strong> <strong>Torque</strong> Control of<br />

Induction Motor", IEEE Transactions on Industrial Electronics, Vol. 51, Issue: 4, Aug. 2004,<br />

pp.776 - 784.<br />

[2] C. Attaianese, D. Capraro, G. Tomasso, "A low cost digital <strong>SVM</strong> modulator with dead time<br />

compensation", Power Electronics Specialists Conference, PESC. 2001 IEEE 32nd Annual, Vol. 1,<br />

17-21 June 2001, pp.158-163.<br />

[3] C. Attaianese, D. Capraro, G. Tomasso, "Hardware dead time compensation for VSI based<br />

electrical drives", IEEE International Symposium on Industrial Electronics, Proceedings ISIE<br />

2001, Vol. 2, 12-16 June 2001, pp.759-764.<br />

[4] U. Baader, M. Depenbrock, G. Gierse, "<strong>Direct</strong> Self Control (DSC) of Inverter-Fed-Inducktion<br />

Machine - A Basis for Speed Control Without Speed Measurement", IEEE Trans. of Industry<br />

Applications, Vol. 28, No. 3 May/June 1992, pp.581-588.<br />

[5] M. M. Bech, "Analysis of Random Pulse-Width Modulation Techniques for Power Electronic<br />

Converters", Alborg University, Denmark Institute of Energy Technology, August 2000.<br />

[6] M.M. Bech, F. Blaabjerg, J.K. Pedersen, "Random modulation techniques with fixed switching<br />

frequency for three-phase power converters", IEEE Transactions on Power Electronics, Vol. 15,<br />

Issue: 4, July 2000, pp.753-761.<br />

[7] M.M. Bech, J.K. Pedersen, F. Blaabjerg, A.M. Trzynadlowski, "A methodology for true<br />

comparison of analytical and measured frequency domain spectra in random PWM converters",<br />

IEEE Transactions on Power Electronics, Vol. 14, Issue: 3, May 1999, pp.578-586.<br />

[8] L. Ben-Brahim, "The analysis and compensation of dead-time effects in three phase PWM<br />

inverters", Industrial Electronics Society, 1998. IECON '98. Proceedings of the 24th Annual<br />

Conference of the IEEE, Vol. 2d, 31 Aug.-4 Sept. 1998, pp.792-797.<br />

[9] L. Ben-Brahim, R. Kurosawa, "Identification of induction motor speed using neural networks",<br />

Record of the Power Conversion Conference, Yokohama 1993, 19-21 April 1993, pp.689-694.<br />

[10] M. Bertoluzzo, G. Buja, R. Menis, "Analytical formulation of the direct control of induction motor<br />

drives", Proceedings of the IEEE International Symposium on Industrial Electronics, ISIE '99, Vol.<br />

1, 12-16 July 1999, pp.PS14-PS20.<br />

[11] T. Biskup, J. Teluk, "Modulacja stochastyczna. Badania eksperymentalne wpływu rozkładu<br />

prawdopodobieństwa generatora losowego na efekty akustyczne", SENE'99, Łódź-Arturówek, 17-<br />

19 Nov. 1999, pp.65-70.<br />

[12] F. Blaschke, "The principle of fiels-orientation as applied to the Transvector closed-loop control<br />

system for rotating-field machines", in Siemens Reviev 34, 1972, pp.217-220.


References<br />

[13] V. Blasko, "Analysis of a hybrid PWM based on modified space-vector and triangle-comparison<br />

methods", IEEE Transactions on Industry Applications, Vol. 33, Issue: 3, May-June 1997, pp.756-<br />

764.<br />

[14] S. Bolognani, A. Di Bella, M. Zigliotto, "Random modulation and acoustic noise reduction in IM<br />

drives: a case study", Ninth International Conference on Electrical Machines and Drives, (Conf.<br />

Publ. No. 468), 1-3 Sept. 1999, pp.137 - 141.<br />

[15] C.J. Bonanno, Li Zhen, Longya Xu, "A direct field oriented induction machine drive with robust<br />

flux estimator for position sensorless control", Industry Applications Conference, 1995. Thirtieth<br />

IAS Annual Meeting, IAS '95., Conference Record of the 1995 IEEE, Vol. 1, 8-12 Oct. 1995,<br />

pp.166-173.<br />

[16] B. K. Bose, "Modern Power Electronics and AC drives", Prentice-Hall, 2002.<br />

[17] G. Buja, D. Casadei, G. Serra, "<strong>Direct</strong> stator flux and torque control of an induction motor:<br />

theoretical analysis and experimental results", Proceedings of the 24th Annual Conference of the<br />

IEEE Industrial Electronics Society, IECON '98, Vol. 1, 31 Aug.-4 Sept. 1998, pp.T50-T64.<br />

[18] G. Buja, D. Casadei, G. Serra, "<strong>DTC</strong>-based strategies for induction motor drives", 23rd<br />

International Conference on Industrial Electronics, Control and Instrumentation, IECON 97, Vol.<br />

4, 9-14 Nov. 1997, pp.1506-1516.<br />

[19] G.S. Buja, M.P. Kazmierkowski, "<strong>Direct</strong> <strong>Torque</strong> Control of PWM Inverter-Fed AC Motors-A<br />

Survey", IEEE Transactions on Industrial Electronics, Vol. 51, Issue: 4, Aug. 2004, pp.744-757.<br />

[20] D. Casadei, F. Profumo, G. Serra, A. Tani, "FOC and <strong>DTC</strong>: two viable schemes for induction<br />

motors torque control", IEEE Transactions on Power Electronics, Vol. 17, Issue: 5, Sept. 2002,<br />

pp.779-787.<br />

[21] D. Casadei, G. Grandi, G. Serra, A. Tani, "Effects of flux and torque hysteresis band amplitude in<br />

direct torque control of induction machines", 20th International Conference on Industrial<br />

Electronics, Control and Instrumentation, IECON '94, Vol. 1, 5-9 Sept. 1994, pp.299-304.<br />

[22] D. Casadei, G. Grandi, G. Serra, A. Tani, "Switching Strategies in <strong>Direct</strong> <strong>Torque</strong> Control of<br />

Induction Machines", Proc. of ICEM Conf., D8.11, 1994, pp.204-209.<br />

[23] D. Casadei, G. Serra, A. Tani, "Analytical Investigation of <strong>Torque</strong> and Flux Ripple in <strong>DTC</strong><br />

Schemes for Induction Motors", Proceedings of the IECON Conference, 1997, pp. 552-556.<br />

[24] D. Casadei, G. Serra, A. Tani, "Constant frequency operation of a <strong>DTC</strong> induction motor drive for<br />

electric vehicle", Proc. of ICEM Conf., Vol. 3, 1996, pp. 224-229.<br />

[25] D. Casadei, G. Serra, A. Tani, "Steady-state and transient performance evaluation of a <strong>DTC</strong><br />

scheme in the low speed range", IEEE Transactions on Power Electronics, Vol. 16, Issue: 6, Nov.<br />

2001, pp.846-851.<br />

142


References<br />

[26] D. Casadei, G. Serra, K. Tani, "Implementation of a direct control algorithm for induction motors<br />

based on discrete space vector modulation", IEEE Transactions on Power Electronics, Vol. 15,<br />

Issue: 4, July 2000, pp.769 - 777.<br />

[27] Tsung-Po Chen, Yen-Shin Lai, Chang-Huan Liu, "A new space vector modulation technique for<br />

inverter control" Power Electronics Specialists Conference, 1999. PESC 99. 30th Annual IEEE,<br />

Vol. 2, 27 June-1 July 1999, pp.777-782.<br />

[28] J. Chiasson, "Dynamic feedback linearization of the induction motor", IEEE Transactions on<br />

Automatic Control, Vol. 38, Issue: 10, Oct. 1993, pp.1588-1594.<br />

[29] Jong-Woo Choi; Seung-Ki Sul, "Inverter output voltage synthesis using novel dead time<br />

compensation", IEEE Transactions on Power Electronics, Vol. 11, Issue: 2, March 1996, pp.221-<br />

227.<br />

[30] Dae-Woong Chung, Joohn-Sheok Kim, Seung-Ki Sul, "Unified voltage modulation technique for<br />

real-time three-phase power conversion", IEEE Transactions on Industry Applications, Vol. 34,<br />

Issue: 2, March-April 1998, pp.374-380.<br />

[31] M. Depenbrock, "<strong>Direct</strong> Self Control of Inverter-Fed Induction Machines", IEEE Trans. on Power<br />

Electronics, Vol. PE-3, no.4, Oct. 1988, pp.420-429.<br />

[32] M. Depenbrock, "<strong>Direct</strong> self-control of the flux and rotary moment of a rotary-field machine", U.S.<br />

Patent 4,678,248.<br />

[33] A. Diaz, E.G. Strangas, "A novel wide range pulse width overmodulation method [for voltage<br />

source inverters]", Applied Power Electronics Conference and Exposition, APEC 2000, Fifteenth<br />

Annual IEEE, Vol. 1, 6-10 Feb. 2000, pp.556-561.<br />

[34] G.F. Franklin, J.D. Powell, M. Workman, "Digital control of dynamic systems", Addison Wesley<br />

Longman, Inc., USA, 1998.<br />

[35] Minghua Fu, Longya Xu, "A novel sensorless control technique for permanent magnet synchronous<br />

motor (PMSM) using digital signal processor (DSP)", Proceedings of the IEEE 1997 National<br />

Aerospace and Electronics Conference, NAECON 1997, Vol. 1, 14-17 July 1997, pp.403-408.<br />

[36] Minghua Fu, Ling Xu, "A sensorless direct torque control technique for permanent magnet<br />

synchronous motors", Power Electronics in Transportation, 22-23 Oct. 1998, pp.21-28.<br />

[37] Minghua Fu, Longya Xu, "A sensorless direct torque control technique for permanent magnet<br />

synchronous motors", Thirty-Fourth IAS Annual Meeting. Conference Record of the 1999 IEEE<br />

Industry Applications Conference, Vol. 1, 3-7 Oct. 1999, pp.159-164.<br />

[38] S. Fukuda, K. Suzuki, "Using harmonic distortion determining factor for harmonic evaluation of<br />

carrier-based PWM methods", Industry Applications Conference, Thirty-Second IAS Annual<br />

Meeting, IAS '97, Conference Record of the 1997 IEEE, Vol. 2, 5-9 Oct. 1997, pp.1534-1541.<br />

[39] P. Grabowski "<strong>Direct</strong> Flux and <strong>Torque</strong> Neuro-Fuzzy Control of Inverter Fed Induction Motor<br />

Drives", PhD Thesis, Warsaw University of Technology, 1999.<br />

143


References<br />

[40] P.Z. Grabowski, M.P. Kazmierkowski, B.K. Bose, F. Blaabjerg, "A simple direct-torque neurofuzzy<br />

control of PWM-inverter-fed induction motor drive", IEEE Transactions on Industrial<br />

Electronics, Vol. 47, Issue: 4, Aug. 2000, pp.863 - 870.<br />

[41] Z. Grunwald, M. P. Kaźmierkowski, W. Koczara, J. Łastowiecki, G. Przywara, "Napęd<br />

elektryczny", Wydawnictwa Naukowo-Techniczne WNT, Warszawa, 1978.<br />

[42] T.G. Habetler, D.M. Divan, "Control strategies for direct torque control using discrete pulse<br />

modulation", IEEE Transactions on Industry Applications, Vol. 27, Issue: 5, Sept.-Oct. 1991,<br />

pp.893-901.<br />

[43] T.G. Habetler, F. Profumo, M. Pastorelli, "<strong>Direct</strong> torque control of induction machines over a wide<br />

speed range", Conference Record of the 1992 IEEE Industry Applications Society Annual<br />

Meeting, Vol.14-9 Oct. 1992, pp.600-606.<br />

[44] T.G. Habetler, F. Profumo, M. Pastorelli, L.M. Tolbert, "<strong>Direct</strong> torque control of induction<br />

machines using space vector modulation", Conference Record of the 1991 IEEE Industry<br />

Applications Society Annual Meeting, Vol.1, 28 Sept.-4 Oct. 1991, pp.428-436.<br />

[45] K. Hasse, "Drehzahlregelverfahren fur schnelle Umkehrantriebe mit stromrichtergespeisten<br />

Asynchron-Kurzschlusslaufermotoren", in Regelungstechnik 20, 1972, pp.60-66.<br />

[46] A.M. Hava, R.J. Kerkman, T.A. Lipo, "A high performance generalized discontinuous PWM<br />

algorithm", Applied Power Electronics Conference and Exposition, APEC '97 Conference<br />

Proceedings 1997, Twelfth Annual, Vol. 2, 23-27 Feb. 1997, pp.886-894.<br />

[47] A.M. Hava, R.J. Kerkman, T.A. Lipo, "Simple analytical and graphical tools for carrier based<br />

PWM methods", Power Electronics Specialists Conference, PESC '97 Record, 28th Annual IEEE,<br />

Vol. 2, 22-27 June 1997, pp.1462-1471.<br />

[48] M. Hinkkanen, J. Luomi, "Modified integrator for voltage model flux estimation of induction<br />

motors", IEEE Transactions on Industrial Electronics, Vol. 50, Issue: 4, Aug. 2003, pp.818-820.<br />

[49] F. Hoffman, M. Janecke, "Fast <strong>Torque</strong> Control of an IGBT-Inverter-Fed Tree-Phase A.C. Drive in<br />

the Whole Speed Range - Experimental Result", Proc. EPE Conf., 1995, pp.3.399-3.404.<br />

[50] D.G. Holmes, "A general analytical method for determining the theoretical harmonic components<br />

of carrier based PWM strategies", Industry Applications Conference, Thirty-Third IAS Annual<br />

Meeting. The 1998 IEEE, Vol. 2, 12-15 Oct. 1998, pp.1207-1214.<br />

[51] D.G. Holmes, "The significance of zero space vector placement for carrier-based PWM schemes",<br />

IEEE Transactions on Industry Applications, Vol. 32, Issue: 5, Sept.-Oct. 1996, pp.1122-1129.<br />

[52] J. Holtz, "Pulsewidth modulation for electronic power conversion", Proceedings of the IEEE, Vol.<br />

82, Issue: 8, Aug. 1994, pp.1194-1214.<br />

[53] J. Holtz, "Sensorless control of induction motor drives", Proceedings of the IEEE, Vol. 90, Issue:<br />

8, Aug. 2002, pp.1359-1394.<br />

144


References<br />

[54] J. Holtz, Juntao Quan, "Sensorless vector control of induction motors at very low speed using a<br />

nonlinear inverter model and parameter identification", IEEE Transactions on Industry<br />

Applications, Vol. 38, Issue: 4, July-Aug. 2002, pp.1087-1095.<br />

[55] J. Holtz, W. Lotzkat, A.M. Khambadkone, "On continuous control of PWM inverters in the<br />

overmodulation range including the six-step mode", IEEE Transactions on Power Electronics, Vol.<br />

8, Issue: 4, Oct. 1993, pp.546-553.<br />

[56] K. B. Howell, "Principles of Fourier analysis", CHAPMAN & HALL/CRC, Boca Raton London<br />

New York Washington, D.C., 2001.<br />

[57] Jun Hu, Bin Wu, "New integration algorithms for estimating motor flux over a wide speed range",<br />

IEEE Transactions on Power Electronics, Vol. 13, Issue: 5, Sept. 1998, pp.969-977.<br />

[58] Hu Hu, Yong Dong Li, Yi Zeng, "<strong>Direct</strong> torque control of induction motor for railway traction in<br />

whole speed range", IECON 02, Industrial Electronics Society, IEEE 2002 28th Annual<br />

Conference, Vol. 3, 5-8 Nov. 2002, pp.2161-2166.<br />

[59] K. D. Hurst, T. G. Habetler, "A Simple, Tacho-Less, I.M. Drive with <strong>Direct</strong> <strong>Torque</strong> Control Down<br />

to Zero Speed", Proceedings of the IECON Conference, Vol.2, 1997, pp.563-568.<br />

[60] K.D. Hurst, T.G. Habetler, G. Griva, F. Profumo, "Speed sensorless field-oriented control of<br />

induction machines using current harmonic spectral estimation", Conference Record of the 1994<br />

IEEE Industry Applications Society Annual Meeting, Vol.1, 2-6 Oct. 1994, pp.601-607.<br />

[61] C.B. Jacobina, A.M.N. Lima, E.R.C. da Silva, A.M. Trzynadlowski, "Current control for induction<br />

motor drives using random PWM", IEEE Transactions on Industrial Electronics, Vol. 45, Issue: 5,<br />

Oct. 1998, pp.704-712.<br />

[62] M. Jayne, I. Ludtke, Liang Yiqiang, T. Arias, "Evaluation of vector and direct torque controlled<br />

strategies for cage rotor induction motor drives", The Third International Power Electronics and<br />

Motion Control Conference, Proceedings PIEMC 2000, Vol. 1, 15-18 Aug. 2000, pp.452-457.<br />

[63] F. Jenni, D. Wueest, "The optimization parameters of <strong>Space</strong> <strong>Vector</strong> Modulation", proc. EPE Conf.,<br />

Vol. 4, 1993, pp. 376-381.<br />

[64] Jong-Lick Lin, "A new approach of dead-time compensation for PWM voltage inverters", IEEE<br />

Transactions on Circuits and Systems I: Fundamental Theory and Applications, Vol. 49, Issue: 4,<br />

April 2002, pp.476-483.<br />

[65] M.P. Kazmierkowski, A.B. Kasprowicz, "Improved direct torque and flux vector control of PWM<br />

inverter-fed induction motor drives", IEEE Transactions on Industrial Electronics, Vol. 42, Issue:<br />

4, Aug. 1995, pp.344-350.<br />

[66] M. P. Kazmierkowski, H. Tunia, "Automatic Control of Converter Fed Drives", ELSEVIER<br />

Amsterdam-London-New York-Tokyo, 1994.<br />

[67] M.P. Kazmierkowski, R. Krishnan, F. Blaabjerg, "Control in Power Electronics Selected<br />

Problems", Academic Press, 2002.<br />

145


References<br />

[68] R.L. Kirlin, A.M. Trzynadlowski, "A unified approach to analysis and design of random<br />

pulsewidth modulation in voltage-source inverters", IEEE Transactions on Circuits and Systems-I:<br />

Fundamental Theory and Applications, Vol. 44, Issue: 8, Aug. 1997, pp.763-766.<br />

[69] Z. Krzemiński, "Cyfrowe sterowanie maszynami asynchronicznymi", Wydawnictwo Politechniki<br />

Gdańskiej, Gdańsk 2001.<br />

[70] Z. Krzemiński, "Nonlinear Control of Induction Motors", in Proc. of 10th IFAC World Congress,<br />

Munich, Germany, 1997, pp.349-354.<br />

[71] Y.-S. Lai, W.-K. Wang, Y.-C. Chen, "Novel Switching Techniques for Reducing the Speed Ripple<br />

of AC Drives With <strong>Direct</strong> <strong>Torque</strong> Control", IEEE Transactions on Industrial Electronics, Vol. 51,<br />

Issue: 4, Aug. 2004, pp.768-775.<br />

[72] C. Lascu, A.M. Trzynadlowski, "Combining the principles of sliding mode, direct torque control,<br />

and space-vector modulation in a high-performance sensorless AC drive", IEEE Transactions on<br />

Industry Applications, Vol. 40, Issue: 1, Jan.-Feb. 2004, pp.170-177.<br />

[73] C. Lascu, I. Boldea, F. Blaabjerg, "Variable-Structure <strong>Direct</strong> <strong>Torque</strong> Control-A Class of Fast and<br />

Robust Controllers for Induction Machine Drives", IEEE Transactions on Industrial Electronics,<br />

Vol. 51, Issue: 4, Aug. 2004, pp.785-792.<br />

[74] Joong-Hui Lee, Chang-Gyun Kim, Myung-Joong Youn, "A dead-beat type digital controller for<br />

the direct torque control of an induction motor", IEEE Transactions on Power Electronics, Vol. 17,<br />

Issue: 5, Sept. 2002, pp.739-746.<br />

[75] Dong-Choon Lee, G-Myoung Lee, "A novel overmodulation technique for space-vector PWM<br />

inverters", IEEE Transactions on Power Electronics, Vol. 13, Issue: 6, Nov. 1998, pp.1144-1151.<br />

[76] D. Leggate, R.J. Kerkman, "Pulse based dead time compensator for PWM voltage inverters",<br />

Proceedings of the 1995 IEEE IECON 21st International Conference on Industrial Electronics,<br />

Control, and Instrumentation, Vol. 1, 6-10 Nov. 1995, pp.474-481.<br />

[77] R.D. Lorenz "Sensorless, drive control methods for stable, high performance, zero speed<br />

operation", proc. EPE-PEMC Conf., Kosice, 2000, pp. 1.1-1.11.<br />

[78] I. Ludtke, M.G. Jayne, "A new direct torque control strategy", IEE Colloquium on Advances in<br />

Control Systems for Electric Drives, 24 May 1995, pp.5/1-5/4.<br />

[79] M. Malinowski "Adaptive modulator for three-phase PWM rectifier/inverter", in proc. EPE-PEMC<br />

Conf., Kosice, 2000, pp.1.35-1.41.<br />

[80] M. Malinowski, "Sensorless Control Strategies for Three-Phase PWM Rectifiers", PhD Thesis,<br />

Warsaw University of Technology, 2001.<br />

[81] R. Marino, P. Valigi, "Nonlinear control of induction motors: a simulation study", in European<br />

Control Conference, Grenoble, France, 1991, pp.1057-1062.<br />

[82] R. Marino, S. Peresada, P. Valigi, "Adaptive input-output linearizing control of induction motors",<br />

IEEE Transactions on Automatic Control, Vol. 38, Issue: 2, Feb. 1993, pp.208 - 221.<br />

146


References<br />

[83] R. Marino, S. Peresada, P. Valigi, "Adaptive partial feedback linearization of induction motors", in<br />

Proc. of the 29th Conference on Decision and Control, Honolulu, Hawaii, Dec. 1990, pp.3313-<br />

3318.<br />

[84] MathWorks, Inc, "Matlab® The Language of Technical Computing", Release 12, 2000.<br />

[85] S.A. Mir, M.E. Elbuluk, D.S. Zinger, "Fuzzy implementation of direct self-control of induction<br />

machines", IEEE Transactions on Industry Applications, Vol. 30, Issue: 3, May-June 1994,<br />

pp.729-735.<br />

[86] K. Ogata, "Modern control engineering", Prentice-Hall, Inc., Englewood Cliffs, New Jersey, 1990.<br />

[87] T. Orłowska-Kowalska, "Bezczujnikowe układy napędowe z silnikami indukcyjnymi", Officyna<br />

Wydawnicza Politechniki Wrocławskiej, Wrocław 2003.<br />

[88] R. Ortega, A. Loria, P. J. Nicklasson, H. Sira-Ramirez, "Passivity-based Control of Euler-Lagrange<br />

Systems", Springer Verlag, London, 1998.<br />

[89] J.O.P. Pinto, B.K. Bose, L.E.B. Da Silva, M.P. Kazmierkowski, "A neural-network-based spacevector<br />

PWM controller for voltage-fed inverter induction motor drive", IEEE Transactions on<br />

Industry Applications, Vol. 36, Issue: 6, Nov.-Dec. 2000, pp.1628-1636.<br />

[90] K.L. Shi, T.F. Chan, Y.K, Wong, S.L. Ho, "<strong>Direct</strong> self control of induction motor based on neural<br />

network", IEEE Transactions on Industry Applications, Vol. 37, Issue: 5, Sept.-Oct. 2001,<br />

pp.1290-1298.<br />

[91] M.G. Simoes, B.K. Bose, "Neural network based estimation of feedback signals for a vector<br />

controlled induction motor drive", IEEE Transactions on Industry Applications, Vol. 31, Issue: 3,<br />

May-June 1995, pp.620-629.<br />

[92] D.L. Sobczuk, "Application of ANN for control of PWM inverter fed induction motor drives", PhD<br />

Thesis, Warsaw University of Technology, 1999.<br />

[93] D.L. Sobczuk, "Feedback linearization control of inverter fed induction motor-DSP<br />

implementation", Proceedings of the 2002 IEEE International Symposium on Industrial<br />

Electronics, ISIE 2002, Vol. 2, 8-11 July 2002, pp.678-682.<br />

[94] D.L. Sobczuk, "Feedback linearization control of inverter fed induction motor-with sliding mode<br />

flux observer", Electrical Drives and Power Electronics International Conference, Slovakia 2003,<br />

pp.465-469.<br />

[95] D.L. Sobczuk, P.Z. Grabowski, "DSP implementation of neural network speed estimator for<br />

inverter fed induction motor", Proceedings of the 24th Annual Conference of the IEEE Industrial<br />

Electronics Society, IECON '98, Vol. 2, 31 Aug.-4 Sept. 1998, pp.981-985.<br />

[96] A. Steimel, "<strong>Direct</strong> Self-Control and Synchronous Pulse Techniques for High-Power Traction<br />

Inverters in Comparison", IEEE Transactions on Industrial Electronics, Vol. 51, Issue: 4, Aug.<br />

2004, pp.810-820.<br />

147


References<br />

[97] I. Takahashi, T. Noguchi, "A new quick-response and high efficiency control strategy of an<br />

induction machine", IEEE Trans. on Industrial Application, Vol. IA-22, no.5, Sept./Oct. 1986,<br />

pp.820-827.<br />

[98] I. Takahashi, T. Noguchi, "Take a Look Back upon the Past Decade of <strong>Direct</strong> <strong>Torque</strong> Control",<br />

Proc. of IECON Conf., Vol. 2, 1997, pp.546-551.<br />

[99] Texas Instruments Incorporated, "TMS320F/C24x DSP Controllers Reference Guide, CPU and<br />

Instruction Set", Literature Number: SPRU160C, 1999.<br />

[100] Texas Instruments Incorporated, "TMS320LF/LC240xA DSP Controllers Reference Guide,<br />

System and Peripherals", Literature Number: SPRU357B, 2002.<br />

[101] Texas Instruments Incorporated, "TMS320LF2407A, TMS320LF2406A, TMS320LF2403A,<br />

TMS320LF2402A TMS320LC2406A, TMS320LC2404A, TMS320LC2402A DSP Controllers",<br />

Literature Number: SPRS145I, 2003.<br />

[102] P. Tiitinen, "The Next Generation Motor Control Method, <strong>Direct</strong> <strong>Torque</strong> Control, <strong>DTC</strong>", PEDES -<br />

New Delhi Conf. Rec., 1996, pp.37-43.<br />

[103] P. Tiitinen, M. Surandra, "The next generation motor control method, <strong>DTC</strong> direct torque control",<br />

Power Electronics, Proceedings of the 1996 International Conference on Drives and Energy<br />

Systems for Industrial Growth, Vol. 1, 8-11 Jan. 1996, pp.37-43.<br />

[104] A.M. Trzynadlowski, F. Blaabjerg, J.K. Pedersen, R.L. Kirlin, S. Legowski, "Random pulse width<br />

modulation techniques for converter-fed drive systems-a review", IEEE Transactions on Industry<br />

Applications, Vol. 30, Issue: 5, Sept.-Oct. 1994, pp.1166-1175.<br />

[105] A.M. Trzynadlowski, S. Legowski, "Minimum-loss vector PWM strategy for three-phase<br />

inverters", IEEE Transactions on Power Electronics, Vol. 9, Issue: 1, Jan. 1994, pp.26-34.<br />

[106] H. Tunia, M. P. Kazmierkowski, "Automatyka napędu przekształtnikowego", Warszawa PWN<br />

1987.<br />

[107] J.W. Umland, M. Safiuddin, "Magnitude and Symmetric Optimum Criterion for the Design of<br />

Linear Control Systems: What Is It and How Does It Compare with the Others?", IEEE<br />

Transactions on Industry Applications, Vol. 26, Issue: 3, May-June 1990, pp.489-497.<br />

[108] P. Vas, "Sensorless <strong>Vector</strong> and <strong>Direct</strong> <strong>Torque</strong> Control", Oxford University Press, 1998.<br />

[109] A.M. Walczyna, "Reduction of current distortions of VSI-fed induction machine controlled by<br />

DSC method-generalized approach", IEEE International Symposium on Industrial Electronics,<br />

Conference Proceedings, ISIE'93 - Budapest, 1-3 June 1993, pp.457-462.<br />

[110] A.M. Walczyna, R.J. Hill, "Novel PWM strategy for direct self-control of inverter-fed induction<br />

motors", IEEE International Symposium on Industrial Electronics, Conference Proceedings,<br />

ISIE'93-Budapest, 1-3 June 1993, pp.610-615.<br />

148


References<br />

[111] Y. Xue, X. Xu, T.G. Habetler, D.M. Divan, "A low cost stator flux oriented voltage source variable<br />

speed drive", Conference Record of the 1990 IEEE Industry Applications Society Annual Meeting,<br />

Vol.1, 7-12 Oct. 1990, pp.410-415.<br />

[112] Z. Yan, C. Jin, V. Utkin, "Sensorless Sliding-Mode Control of Induction Motors", IEEE<br />

Transactions on Industrial Electronics, Vol. 47, Issue: 6, Dec. 2000, pp.1286-1297.<br />

[113] D.S. Zinger, F. Profumo, T.A. Lipo, D.W. Novotny, "A direct field-oriented controller for<br />

induction motor drives using tapped stator windings", IEEE Transactions on Power Electronics,<br />

Vol. 5, Issue: 4, Oct. 1990, pp.446-453.<br />

Papers written during work on this thesis<br />

[114] M. Żelechowski, P. Grabowski, "Universal board - ICG240 for induction motor control drives",<br />

International XII Symposium on Micromachines and Servodrives, Kamień Śląski, Sep. 2000,<br />

pp.475-479. (in Polish)<br />

[115] M. Żelechowski, P. Grabowski, "SimTor <strong>–</strong> New <strong>Vector</strong> Controller for Energy-Efficient Inverter<br />

Fed Induction Motor Drives", International Scientific Conference "Energy Saving in Electrical<br />

Engineering", Proceedings 80th Anniversary of the Faculty of Electrical Engineering at the<br />

Warsaw University of Technology, Warsaw, May 2001, pp.370-372.<br />

[116] M. Żelechowski, M. P. Kaźmierkowski, P. Grabowski, "Practical implementation of direct torque<br />

control of induction motor drive with space vector modulation", XXXVIII International<br />

Symposium on Electrical Machines, Cedzyna-Kielce, June 2002, pp.237-243. (in Polish)<br />

[117] D. Świerczyński, M. Żelechowski, "<strong>Direct</strong> torque and flux control of synchronous and<br />

asynchronous motors", II Krajowa Konferencja MiS-2 Modelowanie i Symulacja, Kościelisko,<br />

June 2002, pp.187-194. (in Polish)<br />

[118] D. Świerczyński, M. Żelechowski, "Universal structure direct torque control for synchronous<br />

permanent magnet and asynchronous motors", International XIII Symposium on Micromachines<br />

and Servodrives, Krasiczyn, Sep. 2002, pp.333-340. (in Polish)<br />

[119] A. M. Trzynadlowski, Z. Wang, J. Nagashima, C. Stancu, M. Żelechowski, "Comparative<br />

Investigation of PWM Techniques for General Motors’ New Drive for Electric Vehicles", Industry<br />

Applications Conference, 37th IAS Annual Meeting, 2002, pp.2010-2015.<br />

[120] M. Żelechowski, P. Kaczyński, M. P. Kaźmierkowski, "Parameters estimation of PWM inverterfed<br />

induction motor", 39th International Symposium on Electrical Machines, Gdańsk-Jurata, June<br />

2003, pp.60.<br />

[121] D. Świerczyński, M. Żelechowski, "Universal structure of direct torque control for AC motor<br />

drives”, III Summer Seminar on Nordick Network for Multi Disciplinary Optimised Electric<br />

Drives, Zegrze, Poland, June 2003, pp.23-27.<br />

149


References<br />

[122] M. Żelechowski, D. Świerczyński, M. P. Kazmierkowski, J. Załęski, "Universal inverter drives<br />

controlled by new generation microprocessors", Elektroinfo Nr 6 (17) 2003, pp.26-28. (in Polish)<br />

[123] M. Żelechowski, M. Malinowski, P. Kaczyński, W. Kołomyjski, M. Twerd, J. Załęski, "DSP<br />

Based Sensorless <strong>Direct</strong> <strong>Torque</strong> Control <strong>–</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulated</strong> (<strong>DTC</strong>-<strong>SVM</strong>) for Inverter Fed<br />

Induction Motor Drives", Problems of Automated Electrodrives Theory and Practice, Crimea,<br />

Ukraine, Sep. 2003, pp.90-92.<br />

[124] M. Jasiński, D. Świerczyński, M. P. Kaźmierkowski, M. Żelechowski, "Sensorless <strong>Direct</strong> Power<br />

and <strong>Torque</strong> Control of <strong>Space</strong> <strong>Vector</strong> <strong>Modulated</strong> AC/DC/AC Converter - Fed Induction Motor",<br />

Control in Power Electronics & Electrical Drives, SENE 2003, Łódź, Nov. 2003, pp.179-185.<br />

[125] M. Żelechowski, P. Kaczyński, "Automatic measurement of induction motor parameters",<br />

Przegląd Elektrotechniczny, No. 1/2004, pp.6-10. (in Polish)<br />

[126] D. Świerczyński, M. Żelechowski, "Universal structure of direct torque control for AC motor<br />

drives", Przegląd Elektrotechniczny, No. 5/2004, pp.489-492.<br />

[127] M. Żelechowski, W. Kolomyjski, M. Twerd, "Industrial Application of Sensorless <strong>Direct</strong> <strong>Torque</strong><br />

Control <strong>–</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulated</strong> (<strong>DTC</strong>-<strong>SVM</strong>) for Inverter Fed Induction Motor Drives", IV<br />

Summer Seminar on Nordick Network for Multi Disciplinary Optimised Electric Drives, Tallinn,<br />

Estonia, June 2004, pp.77-79.<br />

[128] M. Cichowlas, M. Żelechowski, "PWM Rectifier with active filtering", IV Summer Seminar on<br />

Nordick Network for Multi Disciplinary Optimised Electric Drives, Tallinn, Estonia, June 2004,<br />

pp.101-107.<br />

[129] M.P. Kaźmierkowski, M. Żelechowski, D. Świerczynski, "<strong>DTC</strong>-<strong>SVM</strong> an efficient method for<br />

control both induction and PM synchronous motor”, In Proc. of the EPE- PEMC, Riga, Latvia,<br />

Sep. 2004.<br />

[130] M. Jasiński, M.P. Kaźmierkowski, M. Żelechowski, "Unified Scheme of <strong>Direct</strong> Power and<br />

<strong>Torque</strong> Control for <strong>Space</strong> <strong>Vector</strong> <strong>Modulated</strong> AC/DC/AC Converter- Fed Induction Motor", In<br />

Proc. of the EPE- PEMC, Riga, Latvia, Sep. 2004.<br />

[131] M.P. Kaźmierkowski, M. Żelechowski, D. Świerczynski, "Simple <strong>DTC</strong>-<strong>SVM</strong> Control Scheme for<br />

Induction and PM Synchronous Motor", XVI International Conference on Electrical Machines<br />

ICEM’2004, Krakow, Poland, Sep. 2004.<br />

[132] M. Jasiński, M. P. Kaźmierkowski, M. Żelechowski, "<strong>Direct</strong> Power and <strong>Torque</strong> Control Scheme<br />

for <strong>Space</strong> <strong>Vector</strong> <strong>Modulated</strong> AC/DC/AC Converter- Fed Induction Motor", XVI International<br />

Conference on Electrical Machines ICEM’2004, Krakow, Poland, Sep. 2004.<br />

[133] M. Malinowski, W. Kołomyjski, M. Żelechowski, P. Wójcik, "New <strong>Space</strong> <strong>Vector</strong> Modulator in<br />

Industrial Application", IX Sympozjum - Energoelektronika w Nauce i Dydaktyce ENID’2004,<br />

Poznań, Sep. 2004, pp. 115-122.<br />

150


List of Symbols<br />

a = e<br />

j2 π 3<br />

1<br />

= − + j<br />

2<br />

3<br />

2<br />

B - viscous constant<br />

f - frequency<br />

f<br />

s<br />

- sampling frequency<br />

f<br />

sw<br />

- switching frequency<br />

I - current, absolute value<br />

I<br />

A<br />

, I<br />

B<br />

, I<br />

C<br />

- instantaneous values of stator phase currents<br />

I<br />

r<br />

- rotor current space vector<br />

I<br />

s<br />

- stator current space vector<br />

I , I - stator voltage vector components in stationary α − β coordinate<br />

sα sβ<br />

rα rβ<br />

system<br />

I , I - rotor voltage vector components in stationary α − β coordinate system<br />

k - space vector, generally<br />

K<br />

p<br />

- controller gain<br />

K<br />

pM<br />

- torque controller gain<br />

K<br />

pΨ<br />

- flux controller gain<br />

L - inductance, absolute value<br />

L<br />

M<br />

- main, magnetizing inductance<br />

L<br />

s<br />

- stator winding self-inductance<br />

L<br />

r<br />

- rotor winding self-inductance<br />

M - mutual inductance, absolute value


List of symbols<br />

M - torque, absolute value<br />

M<br />

e<br />

- electromagnetic torque<br />

M<br />

L<br />

- load torque<br />

M , m - modulation index<br />

m<br />

s<br />

- number of phase windings<br />

p<br />

b<br />

- number of pole pairs<br />

S A , S B , S C - switching states for the voltage source inverter<br />

R - resistance, absolute value<br />

R<br />

r<br />

- rotor phase windings resistance<br />

R<br />

s<br />

- stator phase windings resistance<br />

T<br />

i<br />

- controller integrating time<br />

T<br />

iM<br />

- torque controller integrating time<br />

TiΨ<br />

- flux controller integrating time<br />

T<br />

D<br />

- dead time of inverter<br />

L<br />

r<br />

T<br />

r<br />

= - rotor time constant<br />

Rr<br />

T<br />

s<br />

- sampling time<br />

T<br />

sw<br />

- switching time<br />

U - voltage, absolute value<br />

U<br />

A<br />

,<br />

U<br />

B<br />

,<br />

U<br />

C<br />

- instantaneous values of the stator phase voltages<br />

U<br />

s<br />

- stator voltage space vector<br />

U<br />

r<br />

- rotor voltage space vector<br />

U - inverter output voltage space vectors, ν = 0,..., 7<br />

ν<br />

U<br />

c<br />

- reference voltage vector<br />

152


List of symbols<br />

U , U - stator voltage vector components in stationary α − β coordinate<br />

U<br />

sα sβ<br />

system<br />

sα c, U<br />

sβc<br />

- reference stator voltage vector components in stationary α − β<br />

coordinate system<br />

U<br />

sdc, U sqc<br />

- reference stator voltage vector components in rotating d − q<br />

coordinate system<br />

U<br />

dc<br />

- inverter dc link voltage<br />

U - peak value of the n-th harmonic, n = 1, 2, 3,…<br />

m<br />

( n)<br />

U<br />

Ac<br />

, U<br />

Bc<br />

, U<br />

Cc<br />

- reference stator phase voltages<br />

U<br />

t<br />

- triangular carrier signal<br />

U<br />

AB<br />

, U<br />

BC<br />

, U<br />

CA<br />

- line to line voltages<br />

Ψ - flux linkage, absolute value<br />

Ψ<br />

A<br />

,<br />

Ψ<br />

B<br />

,<br />

Ψ<br />

C<br />

- flux linkages of the stator phase windings<br />

Ψ<br />

s<br />

- space vector of the stator flux linkage<br />

Ψ<br />

r<br />

- space vector of the rotor flux linkage<br />

Ψ<br />

s<br />

- stator flux amplitude<br />

Ψ<br />

r<br />

- rotor flux amplitude<br />

Ψ , Ψ - stator flux vector components in stationary α − β coordinate system<br />

sα sβ<br />

Ψ , Ψ - rotor flux vector components in stationary α − β coordinate system<br />

rβ rβ<br />

γ<br />

m<br />

- motor shaft position angle<br />

γ<br />

sr<br />

- rotor flux vector angle<br />

γ<br />

ss<br />

- stator flux vector angle<br />

Ω - angular speed, absolute value<br />

153


List of symbols<br />

Ω<br />

K<br />

- angular speed of the coordinate system<br />

Ω<br />

m<br />

- angular speed of the motor shaft<br />

Ω<br />

m<br />

dγ<br />

=<br />

dt<br />

m<br />

Ω<br />

sr<br />

- angular speed of the rotor flux vector<br />

Ω<br />

ss<br />

- angular speed of the stator flux vector<br />

Ω<br />

sr<br />

Ω<br />

ss<br />

dγ<br />

=<br />

dt<br />

sr<br />

dγ<br />

=<br />

dt<br />

ss<br />

Ω<br />

sl<br />

- slip frequency<br />

L 2<br />

M<br />

σ = 1− - total leakage factor<br />

L L<br />

s<br />

r<br />

Superscript<br />

^ - estimated value<br />

Subscripts<br />

..c - reference value<br />

Rectangular coordinate systems<br />

α − β - stator oriented, stationary coordinate system<br />

' '<br />

d − q - rotor oriented, rotated coordinate system<br />

x − y - stator flux oriented, rotated coordinate system<br />

d − q - rotor flux oriented, rotated coordinate system<br />

Abbreviations<br />

IM <strong>–</strong> Induction Motor<br />

MMF <strong>–</strong> Magnetomotive Force<br />

PWM <strong>–</strong> Pulse Width Modulation<br />

154


List of symbols<br />

ZSS <strong>–</strong> Zero Sequence Signals<br />

SPWM <strong>–</strong> Sinusoidal (triangulation) Pulse Width Modulation<br />

SVPWM <strong>–</strong> <strong>Space</strong> <strong>Vector</strong> Pulse Width Modulation<br />

THIPWM <strong>–</strong> Third Harmonic Pulse Width Modulation<br />

DPWM <strong>–</strong> Discontinues Pulse Width Modulation<br />

<strong>SVM</strong> <strong>–</strong> <strong>Space</strong> <strong>Vector</strong> Modulation<br />

OM <strong>–</strong> Overmodulation<br />

RPWM <strong>–</strong> Random Pulse Width Modulation<br />

RLL <strong>–</strong> Random Lead-Lag Modulation<br />

RCD <strong>–</strong> Random Center Pulse Displacement<br />

RZD <strong>–</strong> Random Distribution of the Zero Voltage <strong>Vector</strong><br />

155


Appendices<br />

A.1.<br />

Derivation of Fourier Series Formula for Phase Voltage<br />

If function f is a periodic, piecewise continuous and an odd, then its trigonometric<br />

Fourier series is given by [56]:<br />

( t) b sin( nωt)<br />

= ∑ ∞ f ω<br />

n<br />

(A.1.1)<br />

n=1<br />

where, for n = 1, 2, 3, …<br />

π<br />

2<br />

b n<br />

= f<br />

(A.1.2)<br />

∫<br />

π 0<br />

( ωt) sin( nωt) d( ωt)<br />

Function which describes phase inverter voltage is shown in the Fig. A.1.1<br />

U A<br />

2<br />

U dc<br />

3<br />

1<br />

U dc<br />

3<br />

1<br />

−<br />

3<br />

2<br />

−<br />

3<br />

0<br />

U dc<br />

U dc<br />

π<br />

3<br />

2π<br />

3<br />

π<br />

4π<br />

3<br />

5π<br />

3<br />

2π<br />

ωt<br />

Fig. A.1.1. Phase voltage of the inverter<br />

Taking into consideration this function coefficient b n can be written as follows:<br />

b<br />

n<br />

=<br />

2<br />

π<br />

∫<br />

π 0<br />

U<br />

A<br />

() t sin( nωt) d( ωt)<br />

π<br />

2π<br />

⎛<br />

= ⎜ 3<br />

3<br />

π<br />

2 1<br />

2<br />

⎜<br />

π<br />

∫<br />

∫<br />

1<br />

U<br />

dc<br />

sin<br />

⎜ 3<br />

0<br />

π 3<br />

2π<br />

3<br />

⎝<br />

3<br />

3<br />

2 1<br />

= U<br />

3<br />

⎛<br />

⎜<br />

− cos<br />

⎝<br />

( nωt) d( ωt) + ∫ U<br />

dc<br />

sin( nωt) d( ωt) + U<br />

dc<br />

sin( nωt) d( ωt)<br />

π<br />

2π<br />

π<br />

3<br />

( nωt) 3 − 2cos( nωt) − ( ) π<br />

⎟ π cos nωt<br />

2<br />

0<br />

3<br />

⎠<br />

π n dc<br />

3<br />

2 1 ⎛<br />

⎛ π ⎞ ⎛ 2 ⎞⎞<br />

= U ⎜1−<br />

cos( nπ<br />

) + cos⎜n<br />

⎟ − cos⎜n<br />

π ⎟⎟ (A.1.3)<br />

3π n dc ⎝<br />

⎝ 3 ⎠ ⎝ 3 ⎠ ⎠<br />

⎞<br />

⎞<br />

⎟<br />

⎟<br />

⎟<br />


Appendices<br />

for even n:<br />

π 2<br />

1−<br />

cos ⎜ ⎟ − ⎜n<br />

⎝ 3 ⎠ ⎝ 3<br />

⎛ ⎞ ⎛ ⎞<br />

( n π ) + cos n cos π ⎟ ⎠<br />

⎛ π ⎞ ⎛ π ⎞<br />

= 1 −1+<br />

cos⎜n ⎟ − cos⎜nπ<br />

− n ⎟ = 0<br />

(A.1.4)<br />

⎝ 3 ⎠ ⎝ 3 ⎠<br />

and for uneven n:<br />

π 2<br />

π<br />

π<br />

1−<br />

cos ⎜ ⎟ − ⎜ ⎟ ⎜ ⎟ − ⎜<br />

⎝ 3 ⎠ ⎝ 3 ⎠ ⎝ 3 ⎠ ⎝<br />

3<br />

⎛ ⎛ π ⎞⎞<br />

= 2⎜1+<br />

cos⎜n ⎟⎟<br />

(A.1.5)<br />

⎝ ⎝ 3 ⎠⎠<br />

⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛<br />

⎞<br />

( nπ<br />

) + cos n cos n π = 1+<br />

1+<br />

cos n cos π + ( n −1) π − n ⎟ ⎠<br />

From above formulas the Fourier series for U A is given by:<br />

U<br />

A<br />

4<br />

= U<br />

3π<br />

∑ ∞ dc<br />

n=<br />

1<br />

1 ⎛ ⎛ π ⎞⎞<br />

⎜1+<br />

cos⎜n<br />

⎟⎟sin<br />

n ⎝ ⎝ 3 ⎠⎠<br />

( nωt)<br />

2<br />

= U<br />

π<br />

∑ ∞ dc<br />

n=<br />

1<br />

1<br />

sin<br />

n<br />

( nωt)<br />

(A.1.6)<br />

where:<br />

n=1+6k, k=0, ±1, ±2,…<br />

157


Appendices<br />

A.2.<br />

SABER Simulation Model<br />

The control structures of IM were implemented in SABER v.2.4 Synopsys Inc.<br />

package. SABER provides analysis behavior of the complete analog and mixed-signal<br />

systems including electrical subsystem. SABER model scheme is presented in Fig.<br />

A.2.1.<br />

Fig. A.2.1. SABER model<br />

The SABER package include the electrical and mechanical elements library. The<br />

scheme of inverter (Fig. A.2.2) is based on the transistors and diodes models from<br />

library.<br />

The user of SABER package can create own model using mathematical equation. In<br />

this way is build model of induction motor. The equations (2.14-2.16) described<br />

induction motor in<br />

α − β coordinates system are written in properly form in<br />

“motor.sin” SABER file. The content of this file is shown in Fig. A.2.3<br />

158


Appendices<br />

Fig. A.2.2. Model of inverter<br />

The control algorithm of induction motor has been written in MAST SABER<br />

programming language. The code in MAST language is connected to “Control Block”,<br />

which is shown in Fig. A.2.1. The MAST programming language is very similar to C<br />

language. Therefore, implementation in laboratory setup of simulated structure is easier.<br />

159


Appendices<br />

#motor.sin<br />

template motor t1 t2 t3 t0 = rs,rr,ls,lr,lm,ml,,j<br />

electrical t1, t2, t3, t0<br />

{<br />

t0)+=it1<br />

it1: it1=isa<br />

i(t2->t0)+=it2<br />

it2: it2=0.5*(-isa + sqrt(3)*isb)<br />

i(t3->t0)+=it3<br />

it3: it3=0.5*(-isa - sqrt(3)*isb)<br />

}<br />

}<br />

Fig. A.2.3. SABER file „motor.sin”<br />

160


Appendices<br />

A.3.<br />

Data and Parameters of Induction Motors<br />

Table A.3.1. Data of 3 kW induction motor<br />

Power<br />

Voltage<br />

Current<br />

Frequency<br />

Base speed<br />

Number of pole pairs<br />

Moment of inertia<br />

Nominal torque<br />

Nominal stator flux<br />

P N = 3 kW<br />

U N = 380 V<br />

I N = 6.9 A<br />

f N = 50 Hz<br />

Ω N<br />

= 1415 rpm<br />

p b = 2<br />

J = 0.007 kgm 2<br />

M N = 20 Nm<br />

Ψ sN<br />

= 0.98 Wb<br />

Table A.3.2. Parameters of 3 kW induction motor<br />

Stator winding resistance<br />

Rotor winding resistance<br />

Stator inductance<br />

Rotor inductance<br />

Mutual inductance<br />

R s<br />

= 1.85 Ω<br />

R r<br />

= 1.84 Ω<br />

L s<br />

= 170 mH<br />

L r<br />

= 170 mH<br />

L M<br />

= 160 mH<br />

Table A.3.3. Data of 15 kW induction motor<br />

Power<br />

Voltage<br />

Current<br />

Frequency<br />

Base speed<br />

Number of pole pairs<br />

Moment of inertia<br />

Nominal torque<br />

Nominal stator flux<br />

P N<br />

= 15 kW<br />

U N<br />

= 380 V<br />

I N<br />

= 28.9 A<br />

f N<br />

= 50 Hz<br />

Ω N<br />

= 1460 rpm<br />

p b<br />

= 2<br />

J = 0.875 kgm 2<br />

M N<br />

= 98 Nm<br />

Ψ sN<br />

= 0.98 Wb<br />

161


Appendices<br />

Table A.3.4. Parameters of 15 kW induction motor<br />

Stator winding resistance<br />

Rotor winding resistance<br />

Stator inductance<br />

Rotor inductance<br />

Mutual inductance<br />

R s<br />

= 0.28 Ω<br />

R r<br />

= 0.26 Ω<br />

L s<br />

= 63.5 mH<br />

L r<br />

= 63.5 mH<br />

L M<br />

= 58.1 mH<br />

Table A.3.5. Data of 90 kW induction motor<br />

Power<br />

Voltage<br />

Current<br />

Frequency<br />

Base speed<br />

Number of pole pairs<br />

Moment of inertia<br />

Nominal torque<br />

Nominal stator flux<br />

P N<br />

= 90 kW<br />

U N<br />

= 380 V<br />

I N<br />

= 158 A<br />

f N<br />

= 50 Hz<br />

Ω N<br />

= 1483 rpm<br />

p b<br />

= 2<br />

J = 1.50 kgm 2<br />

M N<br />

= 580 Nm<br />

Ψ sN<br />

= 0.98 Wb<br />

Table A.3.6. Parameters of 90 kW induction motor<br />

Stator winding resistance<br />

Rotor winding resistance<br />

Stator inductance<br />

Rotor inductance<br />

Mutual inductance<br />

R s<br />

= 0.020 Ω<br />

R r<br />

= 0.016 Ω<br />

L s<br />

= 16.36 mH<br />

L r<br />

= 16.74 mH<br />

L M<br />

= 16 mH<br />

162


Appendices<br />

A.4.<br />

Equipment<br />

Table A.4.1. List of equipment<br />

Instrument<br />

Digital oscilloscope<br />

Analyzer<br />

Voltage differential probe<br />

Current probe<br />

Simulation program<br />

Simulation program<br />

Type<br />

Tektronix TDS3034 300MHz<br />

NORMA D6000 Lem<br />

Tektronix P5200<br />

Tektronix TCP A300<br />

SABER 2002.4 Synopsys, Inc.<br />

Matlab 6.1 MathWorks, Inc.<br />

163


Appendices<br />

A.5.<br />

dSPACE DS1103 PPC Board<br />

Physically, DS1103 is built as a PC card that can be mounted into an ISA slot of a<br />

regular PC. The I/O capability is rather impressive providing 300 signals. In order to<br />

simplify the interface, 60 signals out of 300 are selected for further processing and then<br />

connected to the SCU for signal conditioning. The selection is carried out in the<br />

DEMUX card, which was fitted in a shielded box for EMC consideration.<br />

The DS1103 is a single board system based on the Motorola PowerPC 604e/333MHz<br />

processor (PPC), which forms the main processing unit.<br />

I/O Units<br />

A set of on-board peripherals frequently used in digital control systems has been<br />

added to the PPC. They include: analog-digital and digital-analog converters, digital I/O<br />

ports (Bit I/O), and a serial interface. The PPC can also control up to six incremental<br />

encoders, which allow the development of advanced controllers for robots.<br />

DSP Subsystem<br />

The DSP subsystem, based on the Texas Instruments TMS320F240 DSP fixed-point<br />

processor, is especially designed for the control of electric drives. Among other I/O<br />

capabilities, the DSP provides 3-phase PWM generation making the subsystem useful<br />

for drive applications.<br />

CAN Subsystem<br />

A further subsystem, based on Siemens 80C164 micro-controller (MC), is used for<br />

connection to a CAN bus.<br />

Master PPC Slave DSP Slave MC<br />

The PPC has access to both the DSP and the CAN subsystems. Spoken in terms of<br />

inter-processor communication, the PPC is the master, whereas the DSP and the CAN<br />

MC are slaves.<br />

Fig. A.5.14 gives an overview of the functional units of the DS1103 PPC.<br />

164


Appendices<br />

Fig. A.5.1. Block diagram of the dSPACE DS1103 board<br />

The DS1103 PPC Controller Board provides the following features summarized in<br />

alphabetical order:<br />

A/D Conversion<br />

• 4 parallel A/D-converters, multiplexed to 4 channels each, 16-bit resolution, 4 µs<br />

sampling time, ± 10V input voltage range,<br />

• 4 parallel A/D-converters with 1 channel each, 12-bit resolution, 800 ns sampling<br />

time ± 10V input voltage range,<br />

• Slave DSP ADC Unit providing.<br />

• 2 parallel A/D converters, multiplexed to 8 channels each, 10-bit resolution, 6 µs<br />

sampling time ± 10V input voltage range,<br />

Digital I/O<br />

165


Appendices<br />

• 32-bit input/output, configuration byte-wise,<br />

• Slave DSP Bit I/O-Unit providing,<br />

• 19-bit input/output, configuration bit-wise,<br />

CAN Support<br />

• Slave MC fulfilling CAN Specifications 2.0 A and 2.0 B, and ISO/DIS 11898.<br />

D/A Conversion<br />

• 2 D/A converters with 4 channels each, 14-bit resolution ±10 V voltage range<br />

Incremental Encoder Interface<br />

• 1 analog channel with 22/38-bit counter range,<br />

• 1 digital channel with 16/24/32-bit counter range,<br />

• 5 digital channels with 24-bit counter range.<br />

Interrupt Control - Interrupt Handling.<br />

Serial I/O<br />

• standard UART interface, alternatively RS-232 or RS-422 mode.<br />

Timer Services<br />

• 32-bit downcounter with interrupt function (Timer A),<br />

• 32-bit upcounter with pre-scaler and interrupt function,<br />

• 32-bit downcounter with interrupt function (PPC built-in Decrementer),<br />

• 32/64-bit timebase register (PPC built-in Timebase Counter).<br />

Timing I/O<br />

• 4 PWM outputs accessible for standard Slave DSP PWM Generation,<br />

• 3 x 2 PWM outputs accessible for Slave DSP PWM3 Generation and Slave DSP<br />

PWM-SV Generation,<br />

• 4 parallel channels accessible for Slave DSP Frequency Generation,<br />

• 4 parallel channels accessible for Slave DSP Frequency Measurement (F2D) and<br />

Slave DSP PWM Analysis (PWM2D).<br />

166


Appendices<br />

A.6.<br />

Processor TMS320FL2406<br />

Fig. A.6.1 gives overview of the TMS320FL2406 structure.<br />

C2xx<br />

DSP<br />

Core<br />

DARAM (B0)<br />

256 Words<br />

DARAM (B1)<br />

256 Words<br />

DARAM (B2)<br />

32 Words<br />

10 bit ADC<br />

PLL Clock<br />

SCI<br />

SPI<br />

CAN<br />

SARAM (2K Words)<br />

Flash<br />

(32K Words)<br />

Watchdog<br />

Digital I/O<br />

JTAG Port<br />

Event Manager A<br />

- Capture Inputs<br />

- Compare/PWM Outputs<br />

- GP Timers/ PWM<br />

Event Manager B<br />

- Capture Inputs<br />

- Compare/PWM Outputs<br />

- GP Timers/ PWM<br />

Fig. A.6.1. TMS320F2406 device overview<br />

The features of the TMS320FL2406 processor [101] can be summarized as:<br />

• High-Performance Static CMOS Technology:<br />

• 25-ns Instruction Cycle Time (40 MHz),<br />

• 40-MIPS Performance,<br />

• Low-Power 3.3-V Design.<br />

• Based on TMS320C2xx DSP CPU Core:<br />

• Code-Compatible With F243/F241/C242,<br />

• Instruction Set and Module Compatible With F240/C240.<br />

• On-Chip Memory:<br />

• 32K Words x 16 Bits of Flash EEPROM (4 Sectors),<br />

• Programmable "Code-Security" Feature for the On-Chip Flash,<br />

• 2.5K Words x 16 Bits of Data/Program RAM,<br />

167


Appendices<br />

• 544 Words of Dual-Access RAM,<br />

• 2K Words of Single-Access RAM.<br />

• Boot ROM:<br />

• SCI/SPI Bootloader,<br />

• Two Event-Manager (EV) Modules (EVA and EVB), Each Includes:<br />

• Two 16-Bit General-Purpose Timers,<br />

• Eight 16-Bit Pulse-Width Modulation (PWM) Channels Which Enable:<br />

• Three-Phase Inverter Control,<br />

• Center- or Edge-Alignment of PWM Channels,<br />

• Emergency PWM Channel Shutdown With External PDPINTx\<br />

Pin,<br />

• Programmable Deadband (Deadtime) Prevents Shoot-Through Faults,<br />

• Three Capture Units for Time-Stamping of External Events,<br />

• Input Qualifier for Select Pins,<br />

• On-Chip Position Encoder Interface Circuitry,<br />

• Synchronized A-to-D Conversion.<br />

• Watchdog (WD) Timer Module,<br />

• 10-Bit Analog-to-Digital Converter (ADC):<br />

• 16 Multiplexed Input Channels,<br />

• 375 ns or 500 ns MIN Conversion Time,<br />

• Selectable Twin 8-State Sequencers Triggered by Two Event Managers,<br />

• Controller Area Network (CAN) 2.0B Module,<br />

• Serial Communications Interface (SCI),<br />

• 16-Bit Serial Peripheral Interface (SPI),<br />

• Phase-Locked-Loop (PLL)-Based Clock Generation,<br />

168


Appendices<br />

• 40 Individually Programmable, Multiplexed General-Purpose Input/Output<br />

(GPIO) Pins,<br />

• Five External Interrupts (Power Drive Protection, Reset, Two Maskable<br />

Interrupts),<br />

• Power Management:<br />

• Three Power-Down Modes,<br />

• Ability to Power Down Each Peripheral Independently,<br />

• Real-Time JTAG-Compliant Scan-Based Emulation, IEEE Standard 1149.1<br />

(JTAG),<br />

• Development Tools Include:<br />

• Texas Instruments (TI) ANSI C Compiler, Assembler/Linker, and Code<br />

Composer Studio (CCS) Debugger,<br />

• Evaluation Modules,<br />

• Scan-Based Self-Emulation (XDS510),<br />

• Broad Third-Party Digital Motor Control Support,<br />

Package 100-Pin LQFP PZ.<br />

169

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!