07.03.2014 Views

Direct Torque Control with Space Vector Modulation (DTC-SVM) of ...

Direct Torque Control with Space Vector Modulation (DTC-SVM) of ...

Direct Torque Control with Space Vector Modulation (DTC-SVM) of ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

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

POLITECHNIKA<br />

WARSZAWSKA<br />

WARSAW UNIVERSITY OF TECHNOLOGY<br />

Faculty <strong>of</strong> Electrical Engineering<br />

ROZPRAWA DOKTORSKA<br />

Ph.D. Thesis<br />

Dariusz Świerczyński, M. Sc.<br />

<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong><br />

<strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>) <strong>of</strong> Inverter-Fed<br />

Permanent Magnet Synchronous Motor Drive<br />

WARSZAWA<br />

2005


WARSAW UNIVERSITY OF TECHNOLOGY<br />

Faculty <strong>of</strong> Electrical Engineering<br />

Institute <strong>of</strong> <strong>Control</strong> and Industrial Electronics<br />

Ph.D. Thesis<br />

M. Sc. Dariusz Świerczyński<br />

<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong><br />

<strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>) <strong>of</strong> Inverter-Fed<br />

Permanent Magnet Synchronous Motor Drive<br />

Thesis supervisor<br />

Pr<strong>of</strong>. Dr Sc. Marian P. Kaźmierkowski<br />

Warsaw, Poland - 2005


Contents<br />

Table <strong>of</strong> Contents<br />

Chapter 1 1<br />

INTRODUCTION<br />

Chapter 2 8<br />

MODELING AND CONTROL MODES OF PM SYNCHRONOUS DRIVES<br />

2.1 Mathematical model <strong>of</strong> PM synchronous motor 8<br />

2.1.1 Voltage and flux-current equations 9<br />

2.1.2 Instantaneous power and electromagnetic torque 17<br />

2.1.3 Mechanical motion equation 22<br />

2.2 Static characteristic under different control modes 25<br />

2.3 Summary 33<br />

Chapter 3 34<br />

VOLTAGE SOURCE PWM INVERTER FOR PMSM SUPPLY<br />

3.1 Introduction 34<br />

3.2 Voltage source inverter (VSI) 35<br />

3.3 <strong>Space</strong> vector based pulse width modulation (PWM) methods 46<br />

3.4 Summary 52<br />

Chapter 4 53<br />

CONTROL METHODS OF PM SYNCHRONOUS MOTOR<br />

4.1 Introduction 53<br />

4.2 Field oriented control (FOC) 54<br />

4.3 <strong>Direct</strong> torque control (<strong>DTC</strong>) 57<br />

4.4 Summary 64<br />

Chapter 5 65<br />

DIRECT TORQUE CONTROL WITH SPACE VECTOR MODULATION<br />

(<strong>DTC</strong>-<strong>SVM</strong>)<br />

5.1 Introduction 65<br />

5.2 Cascade structure <strong>of</strong> <strong>DTC</strong>–<strong>SVM</strong> scheme 66<br />

5.2.1 Digital flux control loop 68<br />

5.2.2 Digital torque control loop 82<br />

5.3 Parallel structure <strong>of</strong> <strong>DTC</strong>–<strong>SVM</strong> scheme 91<br />

5.3.1 Digital flux control loop 92<br />

5.3.2 Digital torque control loop 102<br />

5.4 Speed control loop for <strong>DTC</strong>–<strong>SVM</strong> structure control 113<br />

5.5 Summary 122<br />

Chapter 6 121<br />

DIRECT TORQUE CONTROL WITH SPACE VECTOR MODULATION (<strong>DTC</strong>-<br />

<strong>SVM</strong>) OF PMSM DRIVE WITHOUT MOTION SENSOR<br />

6.1 Introduction 121<br />

6.2 Initial rotor position estimation method 123<br />

6.3 Stator flux estimation methods 127<br />

6.3.1 Overview 127<br />

6.3.2 Current model based flux estimator 127<br />

6.3.3 Voltage model based flux estimator <strong>with</strong> ideal integrator 128


Contents<br />

6.3.4 Voltage model based flux estimator <strong>with</strong> low pas filter 129<br />

6.3.5 Improved voltage model based flux estimator 130<br />

6.4 Electromagnetic torque estimation 132<br />

6.5 Rotor speed estimation methods 132<br />

6.5.1 Overview 132<br />

6.5.2 Back electromotive force (BEMF) technique 133<br />

6.5.3 Stator flux based technique 133<br />

6.6 Summary 136<br />

Chapter 7 137<br />

DSP IMPLEMENTATION OF <strong>DTC</strong>-<strong>SVM</strong> CONTROL<br />

7.1 Description <strong>of</strong> the laboratory test-stand 137<br />

7.2 Steady state behaviour 140<br />

7.3 Dynamic behaviour 143<br />

7.3.1 Flux and torque control loop 143<br />

7.3.2 Speed control loop 151<br />

Chapter 8 161<br />

SUMMARY AND CLOSING CONCLUSIONS<br />

Appendices 163<br />

Picture <strong>of</strong> rotor and stator <strong>of</strong> PMSM machine<br />

Basic transformation<br />

Model <strong>of</strong> PM synchronous motor- SABER<br />

Parameters <strong>of</strong> PMSM machine<br />

Parameters <strong>of</strong> voltage source inverter<br />

PI speed controller<br />

PWM technique - overmodulation<br />

List <strong>of</strong> Symbols 170<br />

References 172


Introduction<br />

Chapter 1 INTRODUCTION<br />

Recently, an increased interest in application <strong>of</strong> permanent magnet synchronous motors<br />

(PMSM) in speed controlled drives has been observed. This is stimulated mainly by:<br />

• development <strong>of</strong> modern high switching frequency semiconductor power devices (as<br />

for example IGBT modules <strong>of</strong> 5-th generation),<br />

• new rare earth magnetic materials as samarium-cobalt (Sm-Co) or neodymium-ironboron<br />

(Nd-Fe-B),<br />

• specialized digital signal processor (DSP) for AC drive applications <strong>with</strong> integrated<br />

PWM function, A/D converters as well as processing <strong>of</strong> encoder signals (e.g<br />

ADMC401, TMS320FL24XX, TMS320FL28XX).<br />

Synchronous motors <strong>with</strong> an electrically excited rotor winding have a conventional threephase<br />

stator winding (called armature) and an electrically excited field winding on the rotor,<br />

which carries a DC current. The armature winding is similar to the stator <strong>of</strong> induction motor.<br />

The electrically excited field winding can be replaced by permanent magnet (PM) [1]. The use<br />

<strong>of</strong> permanent magnets has many advantages including the elimination <strong>of</strong> brushes, slip rings,<br />

and rotor copper losses in the field winding. It leads to higher efficiency. Additionally since<br />

the copper and iron losses are concentrated in the stator, cooling <strong>of</strong> machines through the<br />

stator is more effective. The lack <strong>of</strong> field winding and higher efficiency results in reduction <strong>of</strong><br />

the machine frame size and higher power/weight ratio.<br />

Figure. 1.1. General classification <strong>of</strong> AC synchronous motors.<br />

1


Introduction<br />

Generally, the permanent magnet AC machines can be classified into two types (Fig.1.1):<br />

trapezoidal type called “brushless DC machine” (BLDCM) and sinusoidal type called<br />

permanent magnet synchronous machine (PMSM). The BLDC machines operate <strong>with</strong><br />

trapezoidal back electromagnetic force (EMF) and require rectangular stator phase current.<br />

The PMSM’s generate sinusoidal EMF and operate <strong>with</strong> sinusoidal stator phase current.<br />

The PMSM can be further divided into two main groups in respect how the magnet bars have<br />

mounted in the rotor [6,7]. In the first group magnets are mounted in the rotor (Fig. 1.2 c-d)<br />

and this type is called interior permanent magnet synchronous motors (IPMSM). The second<br />

group is represented by surface permanent magnet synchronous motors (SPMSM). In the<br />

SPMSM magnet bars are mounted on the rotor surface (Fig. 1.2 a-b).<br />

q<br />

q<br />

SPMSM<br />

S N<br />

S N d<br />

a )<br />

b)<br />

S<br />

N<br />

S<br />

N<br />

d<br />

q<br />

q<br />

IPMSM<br />

S<br />

N<br />

c )<br />

d)<br />

S<br />

N<br />

d<br />

d<br />

Fig. 1.2. The cross section <strong>of</strong> the PMSM rotor shaft and the magnet bars placements:<br />

a),b),c) axial field direction, d) radial field direction.<br />

The magnets can be placed in many ways on the rotor (Fig. 1.2). In radial field fashion the<br />

magnet bars are along the radius <strong>of</strong> the machine and this arrangement provides the highest air<br />

gap flux density, but it has the drawback <strong>of</strong> lower structural integrity and mechanical<br />

robustness. Machines <strong>with</strong> this arrangement <strong>of</strong> magnets are not preferred for high-speed<br />

applications (higher than 3000 rpm). In axial field manner the magnets are placed parallel to<br />

the rotor shaft. This arrangement <strong>of</strong> magnets is much more robust mechanically as compared<br />

2


Introduction<br />

to surface-mounted machine. It makes possible to use IPMSM for higher-speed applications<br />

(contrary to SPMSM’s).<br />

Regardless <strong>of</strong> the fashion <strong>of</strong> mounting the PM, the basic principle <strong>of</strong> motor control is the same<br />

and the differences are only in particularities. An important consequence <strong>of</strong> the method <strong>of</strong><br />

mounting the rotor magnets is the difference in direct and quadrature axes inductance values.<br />

The direct axis reluctance is greater than the quadrature axis reluctance, because the effective<br />

air gap <strong>of</strong> the direct axis is multiple times that <strong>of</strong> the actual air gap seen by the quadrature<br />

axis. As consequence <strong>of</strong> such an unequal reluctance, the quadrature inductance is higher than<br />

direct inductance L q<br />

> L d<br />

. It produces reluctance torque in addition to the mutual torque.<br />

Reluctance torque is produced due to the magnet saliency in the quadrature and the direct axis<br />

magnetic paths. Mutual torque is produced due to the interaction <strong>of</strong> the magnet field and the<br />

stator current. In case where the magnets bars are mounted on the rotor surface the quadrature<br />

inductance is equal direct inductance L q<br />

= L d<br />

, because <strong>of</strong> the same flux paths in d and q axis.<br />

As result the reluctance torque disappears.<br />

Among the main advantage <strong>of</strong> PM machines are [12]:<br />

• high air gap flux density,<br />

• higher power/weight ratio,<br />

• large torque/inertia ratio,<br />

• small torque ripples,<br />

• high speed operation,<br />

• high torque capability (quick acceleration and deceleration),<br />

• high efficiency and high cosφ (low expense for the power supply),<br />

• compact design.<br />

Thanks to this advantages the PMSM’s are usually used in high performance servo drives, in<br />

special applications as computer peripheral equipment, robotics, ect. However, recently the<br />

PMSM are also used as adjustable–speed drives in variety <strong>of</strong> application such as fans, pumps,<br />

compressors, blowers. Another area is automotive application as an alternative drive in hybrid<br />

mode <strong>with</strong> classical engine. The power <strong>of</strong> <strong>of</strong>fered synchronous motors is in the range several<br />

kW to MW.<br />

3


Introduction<br />

The main requirements for high performance PWM inverter-fed PMSM drive can be<br />

formulated as follows:<br />

• operation <strong>with</strong> and <strong>with</strong>out mechanical motion sensor,<br />

• fast flux and torque response,<br />

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

• constant switching frequency,<br />

• uni-polar voltage PWM,<br />

• low flux and torque ripples,<br />

• robustness to parameters variation,<br />

• four quadrant operation.<br />

To meet the above requirements, different control methods can be used [3,4,10].<br />

Variable<br />

Frequency<br />

<strong>Control</strong><br />

Scalar based<br />

controllers<br />

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

controllers<br />

V/Hz=const<br />

<strong>with</strong> stabilization<br />

loop<br />

Field<br />

Oriented<br />

(FOC)<br />

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

<strong>Control</strong><br />

(<strong>DTC</strong>)<br />

PM (rotor)<br />

Flux Oriented<br />

(RFOC)<br />

Stator Flux<br />

Oriented<br />

(SFOC)<br />

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

<strong>Control</strong> <strong>with</strong> <strong>Space</strong><br />

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

(<strong>DTC</strong>-<strong>SVM</strong>)<br />

Circular flux<br />

trajectory<br />

(Takahashi)<br />

Figure 1.3 Classification <strong>of</strong> PMSM control methods.<br />

The general classification <strong>of</strong> the variable frequency control for PMSM is presented in Fig. 1.3.<br />

The PMSM control methods can be divided into scalar and vector control. According to [3],<br />

in scalar control, which based on a relation valid for steady states, only the magnitude and<br />

frequency (angular speed) <strong>of</strong> voltage, currents, and flux linkage space vectors are controlled.<br />

Thus, the control system does not act on space vector position during transient. Therefore, this<br />

control is dedicated for application, where high dynamics is not demanded. Contrary, in<br />

4


Introduction<br />

vector control, which is based on relation valid for dynamics states, not just magnitude and<br />

frequency (angular speed), but also instantaneous position <strong>of</strong> voltage, current and flux space<br />

vectors are controlled. Thus, the control system adjust the position <strong>of</strong> the space vectors and<br />

guarantee their correct orientation for both steady states and transients.<br />

The scalar constant V/Hz control for PMSM <strong>with</strong>out damper winding (squire cage) is not<br />

simple as for induction motor. It requires additional stabilization control loop, which can be<br />

provide by feedback from: rotor velocity perturbation, active power or DC-link current<br />

perturbation [9].<br />

The most popular vector control method developed in 70s, known as field oriented control<br />

(FOC) [31] gives the permanent magnet synchronous motor high performance. In this method<br />

the motor equation are transformed in a coordinate system that rotates in synchronism <strong>with</strong><br />

permanent magnet flux. It allows separately and indirectly control flux and torque quantities<br />

by using current control loop <strong>with</strong> PI controllers like in well known DC machine control [3].<br />

In search <strong>of</strong> a simpler and more robust high performance control system in 80s new vector<br />

control called direct torque control (<strong>DTC</strong>) was developed [50]. It was innovative studies at<br />

this time and completely different approach which depart from the idea <strong>of</strong> coordinate<br />

transformation and the analogy <strong>with</strong> DC motor control. It allows direct control flux and torque<br />

quantities <strong>with</strong>out inner current control loops. Using bang-bang hysteresis controllers for flux<br />

and torque control loops made this control concept very fast and not complicated. However,<br />

the main disadvantage <strong>of</strong> <strong>DTC</strong> is fast sampling time required and variable switching<br />

frequency, because <strong>of</strong> hysteresis based control loops. In order to eliminate above<br />

disadvantages and kept basic control rules <strong>of</strong> classical <strong>DTC</strong>, at the beginning <strong>of</strong> 90’s a new<br />

developed control technique called direct torque control <strong>with</strong> space vector modulator (<strong>DTC</strong>-<br />

<strong>SVM</strong>) has been introduced [54,55]. However, from the formal consideration this method can<br />

also be viewed as stator flux oriented control (SFOC). This control employed instead <strong>of</strong><br />

hysteresis controller as for classical <strong>DTC</strong>, the PI controllers and space vector modulator<br />

(<strong>SVM</strong>). It allows to achieve fixed switching frequency, what considerably reduce switching<br />

losses as well as torque and current ripples. Also requirement <strong>of</strong> very fast sampling time is<br />

eliminated [113,115,117]. Therefore, this new method is subject <strong>of</strong> this thesis. In spite <strong>of</strong><br />

many control strategies there is no one which may be considered as standard solution.<br />

5


Introduction<br />

Therefore, the following thesis can be formulated:<br />

“In the view <strong>of</strong> commercial manufacturing process the most convenient control scheme<br />

for voltage source inverter-fed permanent magnet synchronous motor (PMSM) drives is<br />

direct torque control <strong>with</strong> space vector modulator (<strong>DTC</strong>-<strong>SVM</strong>)”.<br />

To prove the above thesis, the author used methodology based on an analyze and simulation<br />

as well as experimental verification on the laboratory setup <strong>with</strong> 3kW PMSM motor.<br />

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

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

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

• development <strong>of</strong> a simulation algorithm in SABER package for the investigation<br />

<strong>of</strong> PWM inverter-fed PMSM control,<br />

• elaboration and experimental verification <strong>of</strong> digital flux and torque controller<br />

design based on the Z-transform approach for series (cascade) and parallel<br />

structure <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> schemes,<br />

• implementation and verification <strong>of</strong> series (cascade) and parallel <strong>DTC</strong>-<strong>SVM</strong><br />

schemes on experimental laboratory setup <strong>with</strong> 3kW PM synchronous motor<br />

drive controlled by floating point DS1103 board.<br />

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

industry.<br />

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

model <strong>of</strong> PM synchronous motor and his basic control modes are presented. Chapter 3 is<br />

devoted to voltage source inverter, his nonlinear characteristics and different PWM<br />

techniques. Chapter 4 gives brief review <strong>of</strong> PM synchronous motor control method such as<br />

FOC and classical <strong>DTC</strong>. In Chapter 5 two kind <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control schemes are presented.<br />

Also, the analysis and synthesis <strong>of</strong> digital flux, torque and speed controllers based on Z<br />

transform approach are given. Chapter 6 is devoted to initial rotor detection methods, stator<br />

flux vector and rotor speed estimation algorithms. In Chapter 7 experimental results are<br />

6


Introduction<br />

presented and studied. Chapter 8 includes the finally conclusions. Description <strong>of</strong> the SABER<br />

based control algorithm, basic coordinate transformations and parameters <strong>of</strong> used PM<br />

synchronous machine as well as inverter are given in Appendices.<br />

7


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

Chapter 2<br />

MODELING AND CONTROL MODES OF PM SYNCHRONOUS DRIVES<br />

2.1 Mathematical model <strong>of</strong> PM synchronous motor<br />

Development <strong>of</strong> the machine model through the understanding <strong>of</strong> physics <strong>of</strong> the<br />

machine is the key requirement for any type <strong>of</strong> electrical machine control. Since in this<br />

project a Surface type Permanent Magnet Synchronous Motor (SPMSM) is used for the<br />

investigation [9,13,14,15,16]. The development <strong>of</strong> those models is under bellow<br />

assumptions as [3]:<br />

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

• only a fundamental harmonic <strong>of</strong> the magneto motive force (MMF) is taking in to<br />

account,<br />

• the spatially distributed stator and rotor winding are replaced by a concentrated<br />

coil,<br />

• an anisotropy effects, magnetic saturation, iron loses and eddy currents are not<br />

taking into considerations,<br />

• the coil resistances and reactances are taking to be constant,<br />

• in many cases, especially when is considered steady state, the currents and<br />

voltages are assumed to be sinusoidal,<br />

• thermal effect for permanent magnets is omitted.<br />

The synchronous motor model will be presented in space vector notation. <strong>Space</strong> vector<br />

form <strong>of</strong> the machine equations has many advantages such as compact notation, easy<br />

algebraic manipulation, and very simple graphical interpretation. Specially, this notation<br />

is very useful when analyzing the vector control based technique <strong>of</strong> the AC machines.<br />

The space vector representation <strong>of</strong> AC machine equations has been discussed in detail<br />

in number <strong>of</strong> text books ([3,4,12]).<br />

The instantaneous value <strong>of</strong> a three-phase system KA, KB,<br />

K<br />

C<br />

(such as currents, voltages<br />

and flux linkages) can be replaced by one resultant vector called the space vector,<br />

2<br />

K = ⎡ 1 ⋅ K + a⋅ K + a K<br />

3 ⎣<br />

2<br />

A B C<br />

⎤<br />

⎦<br />

(2.1)<br />

8


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

2π<br />

2π<br />

4π<br />

j 1 3<br />

where:1,<br />

3 2<br />

− j j 1 3<br />

a= e =− + j , a = e<br />

3<br />

= e<br />

3<br />

=− − j - complex vectors, 2/3 –<br />

2 2<br />

2 2<br />

normalization factor (guarantee that for balanced sinusoidal waveforms the magnitude<br />

<strong>of</strong> the space vector is equal to the amplitude <strong>of</strong> that phase waveforms).<br />

The elements <strong>of</strong> this space vector satisfy the condition:<br />

KA + KB + KC<br />

= 0<br />

(2.2)<br />

and it means that we have three-phase system <strong>with</strong>out neutral wire.<br />

2.1.1 Voltage and current equations<br />

For idealized motor (Fig. 2.1), the following equations <strong>of</strong> the instantaneous stator phase<br />

voltages can be written [3]:<br />

B<br />

b<br />

Z sB<br />

I sB<br />

a<br />

U sB<br />

S<br />

N<br />

N<br />

S<br />

U sA<br />

γ m<br />

Z sA<br />

A<br />

I sA<br />

I sC<br />

Z sC<br />

U sC<br />

N<br />

S<br />

C<br />

c<br />

Figure 2.1. Layout and symbols for three-phase PMSM electric motor windings.<br />

dΨ<br />

dt<br />

sA<br />

sA<br />

=<br />

sA sA<br />

+ (2.3a)<br />

U I R<br />

U I R<br />

dΨ<br />

dt<br />

sB<br />

sB<br />

=<br />

sB sB<br />

+ (2.3b)<br />

U I R<br />

dΨ<br />

dt<br />

sC<br />

sC<br />

=<br />

sC sC<br />

+ (2.3c)<br />

9


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

where<br />

U ,<br />

sA, U<br />

sB<br />

U<br />

sC<br />

are the instantaneous stator voltage values, I<br />

sA<br />

I<br />

sB<br />

, I<br />

sC<br />

, are<br />

instantaneous values <strong>of</strong> the current,<br />

R = R = R = R is the resistance <strong>of</strong> the stator<br />

s<br />

sA<br />

sB<br />

sC<br />

windings, and ΨsA,<br />

Ψ<br />

sB<br />

and Ψ<br />

sC<br />

are magnetic flux linkages stator windings A, B and<br />

C , respectively.<br />

Using the space vector theory to voltage equations we can written in vector form<br />

where:<br />

U<br />

d Ψ<br />

dt<br />

sABC<br />

sABC<br />

= Rs<br />

I<br />

sABC<br />

+ (2.4)<br />

2 2<br />

U<br />

sABC<br />

= (1 UsA + aUsB + a UsC<br />

) ,<br />

2 2<br />

2 2<br />

I<br />

sABC<br />

= (1 IsA + aIsB + a IsC<br />

) ,<br />

sABC (1<br />

sA<br />

a<br />

sB<br />

a<br />

sC )<br />

3<br />

3<br />

3<br />

stator voltage, current and flux space vectors, respectively.<br />

Ψ = Ψ + Ψ + Ψ are the<br />

The stator winding flux consist <strong>of</strong> rotor flux and stator flux linkages:<br />

where,<br />

Ψ<br />

sABC<br />

=Ψ<br />

ABC ( s) +Ψ<br />

ABC( r)<br />

(2.5)<br />

⎡ LsA MsAB MsAC⎤⎡IsA⎤<br />

⎢<br />

M L M<br />

⎥⎢<br />

I<br />

⎥<br />

Ψ<br />

ABC( s)<br />

= ⎢ sBA sB sBC ⎥⎢ sB ⎥<br />

⎢⎣ M<br />

sCA<br />

MsCB L ⎥⎢<br />

sC ⎦⎣I<br />

⎥<br />

sC ⎦<br />

(2.6)<br />

⎡<br />

⎤<br />

⎢ cosθ<br />

⎥<br />

r<br />

⎢<br />

⎥<br />

2π<br />

Ψ<br />

ABC ( r)<br />

=Ψ ⎢<br />

PM<br />

cos( θr<br />

− ) ⎥<br />

⎢ 3 ⎥<br />

⎢<br />

2π<br />

⎥<br />

cos( θr<br />

+ )<br />

⎢⎣<br />

3 ⎥⎦<br />

(2.7)<br />

and, θr<br />

is electrical rotor position. Mechanical rotor position is defined as:<br />

θ = p γ<br />

(2.8)<br />

r b m<br />

where: pb<br />

- number <strong>of</strong> pole pairs, γ<br />

m<br />

- mechanical position.<br />

In equation (2.6) LsA<br />

is the self-inductance <strong>of</strong> phase A winding, M<br />

sAB<br />

and M<br />

sAC<br />

are the<br />

mutual inductances between A and B phase, A and C phase, respectively. For self and<br />

mutual inductances <strong>of</strong> B and C phase the same notations used. In (2.7),<br />

Ψ<br />

PM<br />

is the<br />

10


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

amplitude <strong>of</strong> the flux linkages established by the permanent magnet on the rotor. The<br />

inductances are described below.<br />

Due to the rotor saliency in IPMSM the air gap is not uniform and, therefore, the self<br />

and mutual inductances <strong>of</strong> stator windings are a function <strong>of</strong> the rotor position.<br />

The derivation <strong>of</strong> these rotor position dependent inductances is available in details in<br />

[5]. The results are summarized here as follows:<br />

The stator winding self-inductances are<br />

L = L + L − L cos 2θ<br />

(2.9a)<br />

sA ls A B r<br />

2π<br />

4π<br />

LsB = Lls + LA −LB cos2( θr − ) = Lls + LA −LB cos(2 θr<br />

− )<br />

3 3<br />

(2.9b)<br />

2π<br />

4π<br />

LsC = Lls + LA − LB cos 2( θr + ) = Lls + LA − LB cos(2 θr<br />

+ )<br />

3 3<br />

(2.9c)<br />

where, Lls<br />

is stator-winding leakage inductance and LA,<br />

LB<br />

are given by<br />

L<br />

A<br />

⎛ms<br />

⎞<br />

= ⎜<br />

2 ⎟<br />

⎝ ⎠<br />

2<br />

πµ rlε<br />

0 1<br />

(2.10a)<br />

L<br />

B<br />

1 ⎛ms<br />

⎞<br />

=<br />

2 ⎜<br />

2 ⎟<br />

⎝ ⎠<br />

2<br />

πµ rlε<br />

0 2<br />

(2.10b)<br />

where,<br />

m<br />

s<br />

is number <strong>of</strong> turns <strong>of</strong> each phase winding, r is radius, which is from center<br />

<strong>of</strong> machine to the inside circumference <strong>of</strong> the stator, and l is the axial length <strong>of</strong> the air<br />

gap <strong>of</strong> the machine, µ<br />

0<br />

is permeability <strong>of</strong> the air, ε<br />

1<br />

and ε<br />

2<br />

are defined as s:<br />

1 1 1<br />

ε<br />

1<br />

= ( + )<br />

(2.11a)<br />

2 g g<br />

min<br />

max<br />

1 1 1<br />

ε<br />

2<br />

= ( − )<br />

(2.11b)<br />

2 g g<br />

min<br />

max<br />

where,<br />

gmin<br />

is minimum air gap length and<br />

max<br />

g<br />

is maximum air gap length.<br />

The mutual inductances between stator phase are:<br />

1 π 1 2π<br />

MsAB = MsBA =− LA −LB cos 2( θr − ) =− LA −LB cos(2 θr<br />

− ) (2.12a)<br />

2 3 2 3<br />

11


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

1 π 1 2π<br />

MsAC = MsCA =− LA − LB cos 2( θr + ) =− LA − LB cos(2 θr<br />

+ ) (2.12b)<br />

2 3 2 3<br />

1 1<br />

MsBC = MsCB =− LA − LB cos 2( θr + π) =− LA − LB cos(2θr<br />

+ 2 π)<br />

2 2<br />

1<br />

=− LA −LBcos 2θ<br />

r<br />

2<br />

(2.12c)<br />

Using the space vector theory, the flux linkage<br />

Ψ<br />

sABC<br />

space vector can be written as:<br />

3 3<br />

∗ j2θr<br />

jθr<br />

Ψ<br />

sABC<br />

= ( Lls + LA)<br />

IsABC − LB IsABC<br />

e +Ψ<br />

PM<br />

e<br />

(2.13)<br />

2 2<br />

where,<br />

2 2<br />

(1<br />

sA sB sC )<br />

IsABC<br />

= I + aI + a I ,<br />

3<br />

space vector and conjugate stator current space vector.<br />

2 2 (1<br />

sA sB sC )<br />

I ∗ sABC<br />

= I + a I + aI are the stator current<br />

3<br />

Taking into account that:<br />

Ld = Lls + Lmd<br />

(2.14a)<br />

Lq = Lls + Lmq<br />

(2.14b)<br />

3 3<br />

where, Lmd = ( LA + LB<br />

) , Lmq = ( LA − LB<br />

) are d and q magnetizing inductances and<br />

2<br />

2<br />

are defined as [5].<br />

Finally, equations (2.13) comes as:<br />

where, L<br />

d<br />

, L<br />

q<br />

are d and q inductances.<br />

Ld + Lq Lq −Ld ∗ j2θr<br />

jθ<br />

r<br />

Ψ<br />

sABC<br />

= ( ) I<br />

sABC<br />

− ( ) IsABC<br />

e +Ψ<br />

PM<br />

e<br />

(2.15)<br />

2 2<br />

<strong>Space</strong> vector form <strong>of</strong> machine equations (2.4, 2.15) becomes more compact, but the<br />

rotor position dependent parameters still exist in that form <strong>of</strong> expressions for the stator<br />

flux linkage space vector. Therefore, the space vector model is still not simple to use for<br />

the analysis. A simplification can be made if the space vector model is referred to a<br />

suitably selected rotating frame.<br />

12


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

Figure 2.2 shows axes <strong>of</strong> reference for the three-stator phase ABC. , , It also shows a<br />

rotating set <strong>of</strong> x,<br />

y axes, where the angleθ K<br />

is position <strong>of</strong> x -axis in respect to the stator<br />

A phase axis. Variables along the AB , and C axes can be referred to the x − and<br />

y − axes by the expression:<br />

⎡K<br />

A ⎤<br />

⎡K<br />

x ⎤ 2 ⎡ cosθK cos( θK − 2 π / 3) cos( θK<br />

+ 2 π / 3) ⎤⎢<br />

K<br />

⎥<br />

⎢ B<br />

K<br />

⎥ =<br />

y 3<br />

⎢<br />

sinθK sin( θK 2 π / 3) sin( θK<br />

2 π / 3)<br />

⎥⎢ ⎥<br />

⎣ ⎦ ⎣− − − − + ⎦<br />

⎢⎣K<br />

⎥<br />

C ⎦<br />

(2.16)<br />

y<br />

K B<br />

K ABC<br />

K y<br />

K x<br />

Ω K<br />

x<br />

θ K<br />

K A<br />

K C<br />

Figure 2.2. Stator fixed three phase axes (A,B,C) and general rotating reference frame ( x,<br />

y ).<br />

Finally, the space vector in general rotating frame can be written as:<br />

j K<br />

K = K (cosΘ + jsin Θ ) = K e θ<br />

(2.17)<br />

ABCs K K<br />

K K<br />

In this case the voltage equation (2.4) using (2.17) can written as:<br />

jθK jθ d<br />

K jθK<br />

U<br />

sKe Rs IsKe (<br />

sKe<br />

)<br />

dt<br />

= + Ψ (2.18)<br />

Using chain rule, equation. (2.17) and divided by term<br />

j K<br />

e θ<br />

can be written as:<br />

d Ψ<br />

U<br />

sK<br />

R I j<br />

dt<br />

= + + Ω Ψ (2.19)<br />

sK<br />

s sK K sK<br />

where<br />

U<br />

sK<br />

rotating frame.<br />

, I sK , Ψ sK is the stator voltage, current and flux space vector in general<br />

Making similar arrangement like for the voltage equation the flux linkage vector in<br />

general reference frame can be expressed as:<br />

13


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

Ld + Lq Lq −Ld j2( θr θK) j( θr θK)<br />

Ψ<br />

sK<br />

= ( ) IsK − ( ) I ∗ sK<br />

e − +Ψ<br />

PM<br />

e<br />

− (2.20)<br />

2 2<br />

Stator fixed system ( α,<br />

β )<br />

Taking the angular speed <strong>of</strong> the reference frame to be Ω<br />

K<br />

= 0 and θ<br />

K<br />

= 0 , the set <strong>of</strong><br />

synchronous machine vector equations (2.19) and (2.20) my be written as:<br />

U<br />

d Ψ<br />

sαβ<br />

sαβ<br />

= Rs<br />

Isαβ<br />

+ (2.21)<br />

dt<br />

Ld + Lq Lq −Ld ∗ j2θr<br />

jθ<br />

r<br />

Ψ<br />

sαβ = ( ) I<br />

sαβ − ( ) Isαβ<br />

e +Ψ<br />

PM<br />

e<br />

(2.22)<br />

2 2<br />

Substituting to above equations the following expressions for complex vectors<br />

U U jU<br />

= s sα<br />

+ , αβ sβ<br />

s sα<br />

sβ<br />

I αβ = I + jI , Ψ αβ =Ψ<br />

sα<br />

+ jΨ sβ<br />

and splitting into real and<br />

s<br />

imaginary parts one can obtain the scalar form <strong>of</strong> the machine equations in stationary<br />

α,<br />

β reference frame:<br />

U<br />

U<br />

dΨ<br />

dt<br />

sα<br />

sα<br />

= RsIsα<br />

+ (2.23a)<br />

dΨ<br />

sβ<br />

sβ<br />

= RsIsβ<br />

+ (2.23b)<br />

dt<br />

Ld + Lq Lq −Ld Lq −Ld<br />

( cos2 θ ) Is<br />

( )sin2θ Is<br />

cosθ<br />

2 2 2<br />

Ψ = − − +Ψ (2.24a)<br />

sα r α r β PM r<br />

Lq − Ld Ld + Lq Lq −Ld<br />

Ψ<br />

sβ =− ( )sin 2 θrIsα + [( ) + ( )cos2 θr] Isβ<br />

+Ψ<br />

PM<br />

sinθ<br />

(2.24b)<br />

r<br />

2 2 2<br />

Note, that in the flux-current equations (2.24a and b) still we can observe that value <strong>of</strong><br />

inductances depends on rotor position θ<br />

r<br />

.<br />

Stator flux fixed system ( x,<br />

y )<br />

In order to take advantage <strong>of</strong> the set <strong>of</strong> equations (2.19) and (2.20) in rotating coordinate<br />

system, one assumes that the coordinate system rotates <strong>with</strong> the stator flux linkage<br />

angular speed Ω<br />

K<br />

=Ω<br />

Ψ<br />

and θK<br />

= θ Ψ<br />

. As a<br />

sx s<br />

Ψ =Ψ , δ = −( θ − θ )<br />

Ψ<br />

r<br />

Ψ<br />

d Ψ<br />

U R I j<br />

dt<br />

sxy<br />

sxy<br />

=<br />

s sxy<br />

+ + ΩΨsΨ (2.25)<br />

sxy<br />

Ld + Lq Lq −Ld ∗ − j2δΨ<br />

− jδΨ<br />

Ψ<br />

sxy<br />

= ( ) Isxy − ( ) Isxy<br />

e +Ψ<br />

PM<br />

e<br />

(2.26)<br />

2 2<br />

14


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

Substituting to above equations the following expressions for complex vectors<br />

U U jU<br />

sxy<br />

=<br />

sx<br />

+<br />

sy, Isxy<br />

Isx jIsy<br />

= + , Ψ =Ψ<br />

s<br />

and splitting into real and imaginary parts<br />

sxy<br />

one can obtain the scalar form <strong>of</strong> the machine equations in stationary x,<br />

y reference<br />

frame:<br />

U<br />

dΨ<br />

dt<br />

s<br />

sx<br />

= Rs Isx<br />

+ (2.27a)<br />

U = R I +ΩΨ<br />

Ψ (2.27b)<br />

sy s sy s<br />

s<br />

1 1<br />

Ψ<br />

s<br />

= [( Ld + Lq) −( Lq − Ld)cos 2 δΨ ] Isx<br />

+ ( Lq − Ld)sin 2δΨIsy<br />

+Ψ<br />

PM<br />

cosδΨ<br />

2 2<br />

(2.28a)<br />

1 1<br />

0 = ( Lq − Ld)sin2 δΨIsx<br />

+ [( Ld + Lq) + ( Lq −Ld)cos2 δΨ] Isy<br />

−Ψ<br />

PM<br />

sinδΨ<br />

2 2<br />

(2.28b)<br />

The current-flux equations can be expressed also in simplest form as:<br />

2 2<br />

s<br />

( Ld cos δΨ Lqsin δΨ) Isx<br />

( Lq Ld)sinδΨcosnδΨIsy<br />

PM<br />

cosδΨ<br />

Ψ = + + − +Ψ (2.29a)<br />

2 2<br />

q d<br />

δΨ δΨ sx d<br />

δΨ q<br />

δΨ sy PM<br />

δΨ<br />

0 ( L L )sin cos I ( L sin L cos ) I sin<br />

= − + + −Ψ (2.29b)<br />

Rotor flux fixed system ( dq) ,<br />

In order to take advantage <strong>of</strong> the set <strong>of</strong> equations (2.19) and (2.20) in rotating coordinate<br />

system, one assumes that the coordinate system rotates <strong>with</strong> the rotor flux angular speed<br />

Ω<br />

K<br />

= pbΩ m<br />

and θK = pbγm = θr<br />

d Ψ<br />

U<br />

sdq<br />

R I jp<br />

dt<br />

sdq<br />

s sdq<br />

b m<br />

= + + Ω Ψ (2.30)<br />

sdq<br />

Ld + Lq Lq −Ld<br />

∗<br />

Ψ<br />

sdq<br />

= ( ) Isdq − ( ) Isdq<br />

+Ψ<br />

PM<br />

(2.31)<br />

2 2<br />

Substituting the following expressions for complex vectors U = Usd + jUsq,<br />

I I jI<br />

sdq<br />

=<br />

sd<br />

+<br />

sq, sdq sd<br />

j<br />

sq<br />

Ψ =Ψ + Ψ to (2.30) and (2.31), and splitting for real and<br />

imaginary parts the scalar form <strong>of</strong> the machine equations in rotational fixed reference<br />

frame can be obtained:<br />

dΨ<br />

sd<br />

Usd = Rs Isd + − pbΩmΨ sq<br />

(2.32a)<br />

dt<br />

sdq<br />

15


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

dΨsq<br />

Usq = Rs Isq + + pbΩmΨ sd<br />

(2.32b)<br />

dt<br />

where,<br />

Ψ<br />

sd<br />

= LI<br />

d sd<br />

+Ψ (2.33a)<br />

PM<br />

Ψ<br />

sq<br />

= LqI<br />

(2.33b)<br />

sq<br />

It should be noted that when transforming the flux linkage vector<br />

Ψ s<br />

to the dq ,<br />

reference frame the rotor position θ r<br />

dependent terms disappear it can be seen from<br />

equation (2.31). This is the main advantage <strong>of</strong> rotor-oriented representation.<br />

Substituting the relationship <strong>of</strong> (2.33a-b) into (2.32a-b), and also considering<br />

dΨ PM = 0 , the most common scalar form <strong>of</strong> the machine voltage equations in the rotor<br />

dt<br />

reference frame can be obtained as:<br />

dIsd<br />

Usd = RsIsd + Ld − pbΩ mLqIsq<br />

(2.34a)<br />

dt<br />

dIsq<br />

Usq = Rs Isq + Lq + pbΩmΨ PM<br />

+ pbΩ mLd Isd<br />

(2.34b)<br />

dt<br />

Based on the above voltage-current equations it is possible to draw the equivalent<br />

electrical circuit separately for d and q axes (Fig. 2.3).<br />

R s<br />

p Ω<br />

L I<br />

b m q sq<br />

R s<br />

p Ω<br />

L I<br />

b m d sd<br />

U<br />

sd<br />

I sd<br />

L d<br />

U<br />

sq<br />

I sq<br />

p Ω Ψ<br />

b m PM<br />

L q<br />

Figure 2.3. Equivalent circuit model <strong>of</strong> PMSM in the rotor reference frame. (a) Rotor d-axis<br />

equivalent circuit, (b) Rotor q-axis equivalent circuit.<br />

16


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

2.1.2 Instantaneous power and electromagnetic torque<br />

The three-phase star-connection system <strong>with</strong>out neutral wire is shown in Fig. 2.4. This<br />

is classical configuration for AC motor windings connections.<br />

A<br />

I sA<br />

U sAC<br />

U sAB<br />

B<br />

U sA<br />

Z sC<br />

Z sA<br />

U sAB<br />

Z sB<br />

C<br />

U sBC<br />

I sC<br />

U sC U sB<br />

I sB<br />

Figure 2.4. Three-phase star connection system <strong>with</strong>out neutral wire.<br />

For this configuration the expression for instantaneous active power supplied to load<br />

can be expressed as:<br />

P= UsAIsA+ UsBIsB+ UsCIsC<br />

(2.35)<br />

Introducing space vector definition, after some arrangement and taking into account the<br />

relation: I + I + I = 0, the equation (2.35) can be written as:<br />

sA sB sC<br />

3 ∗<br />

P= Re[ U<br />

sABC<br />

IsABC<br />

]<br />

(2.36)<br />

2<br />

For dq , frame, the equation (2.35) for the active power can be expressed as:<br />

3<br />

P= ( UsdIsd + UsqIsq<br />

)<br />

(2.37)<br />

2<br />

Substituting voltage equation (2.4) into (2.36), and adopting Ω<br />

K<br />

= pbΩmone obtains<br />

3 ∗ d Ψ<br />

[Re(<br />

sABC ∗ ∗<br />

P= RsIsABC IsABC + IsABC − jpbΩmΨ sABCIsABC<br />

)] (2.38)<br />

2<br />

dt<br />

Note that<br />

sABC<br />

sABC<br />

2<br />

s<br />

I I ∗ = I and:<br />

3 2 d Ψ<br />

[ Re(<br />

sABC ∗<br />

∗<br />

P= Rs Is + IsABC ) + Re( −jpbΩmΨ sABC<br />

IsABC<br />

)] (2.39)<br />

2<br />

dt<br />

d ΨsABC<br />

Hence, neglecting the losses in stator resistance R<br />

s<br />

and assuming that = 0 , the<br />

dt<br />

electromagnetic power is expressed:<br />

17


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

3<br />

∗<br />

Pe = pbΩmIm( Ψ<br />

sABC<br />

IsABC<br />

)<br />

(2.40)<br />

2<br />

In dq , frame the active power can be written:<br />

3<br />

Pe = pbΩm( ΨsdIsq −Ψ<br />

sqIsd)<br />

(2.41)<br />

2<br />

For the presented system (Fig. 2.4) the expression for instantaneous reactive power<br />

supplied to the three-phase load system <strong>with</strong>out neutral wire can be calculated as:<br />

1<br />

Q= ( IsAUsBC + IsBUsCA + IsCUsAB<br />

)<br />

(2.42)<br />

3<br />

Introducing the space vector definition into equation (2.42), after some arrangement,<br />

and taking into account the relation: I + I + I = 0, one obtains:<br />

sA sB sC<br />

3 ∗<br />

Q= Im[ U<br />

sABC<br />

IsABC<br />

]<br />

(2.43)<br />

2<br />

In dq , frame the reactive power is expressed as:<br />

3<br />

Q= ( UsqIsd − UsdIsq<br />

)<br />

(2.44)<br />

2<br />

Substituting voltage equation (2.4) into (2.43), adopting Ω<br />

K<br />

= pbΩ m<br />

and made similar<br />

arrangements like for active power calculation, the final expression for reactive power<br />

is:<br />

3<br />

∗<br />

Q= pbΩmRe( Ψ<br />

sABC<br />

I<br />

sABC<br />

)<br />

(2.45)<br />

2<br />

In dq , frame the expression (2.45) for the reactive power becomes:<br />

3<br />

Q= pbΩm ( Ψ<br />

sd<br />

Isd +Ψ<br />

sqIsq<br />

)<br />

(2.46)<br />

2<br />

The important quantity <strong>of</strong> the drive is the power factor cosφ , which can be calculated<br />

as:<br />

Q<br />

cosφ = (2.47)<br />

S<br />

where S is module <strong>of</strong> apparent power vector S = P+ jQ:<br />

2 2<br />

S = P + Q<br />

(2.48)<br />

18


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

The instantaneous electromagnetic torque developed by an electric motor can be defined<br />

as:<br />

P<br />

e<br />

M<br />

e<br />

= (2.49)<br />

Ω<br />

m<br />

where, P<br />

e<br />

is the electromagnetic power and<br />

Ωm<br />

is the mechanical angular rotor speed.<br />

Finally, taking into account equation (2.49) the expression for electromagnetic torque<br />

can be obtained as:<br />

and in dqframe: ,<br />

3<br />

∗<br />

Me = pbIm( Ψ<br />

sABC<br />

IsABC<br />

) , (2.50)<br />

2<br />

3<br />

M<br />

e<br />

= pb ( Ψsd Isq −Ψ<br />

sqIsd<br />

)<br />

(2.51)<br />

2<br />

Substituting Ψ , Ψ from (2.33a-b), the torque expression <strong>of</strong> equations (2.47)<br />

becomes:<br />

sd<br />

sq<br />

3<br />

M<br />

e<br />

= pb ( ΨPM Isq −( Lq − Ld ) Isd Isq<br />

)<br />

(2.52)<br />

2<br />

It can be seen from (2.52), that developed torque consist <strong>of</strong> two parts, one produced by<br />

the permanent magnet flux called synchronous torque ( M<br />

reluctance torque ( M<br />

er<br />

es<br />

) and the second called<br />

), which is produced by the difference <strong>of</strong> the inductance in rotor<br />

d- and q-axes. Expressions for those two torque components are:<br />

3<br />

M<br />

es<br />

= pbΨ PM<br />

Isq<br />

(2.53a)<br />

2<br />

3<br />

M<br />

er<br />

=− pb ( Lq − Ld ) Isd Isq<br />

(2.53b)<br />

2<br />

It should be mentioned that for SPMSM ( L d<br />

= L ) the reluctance torque does not exist<br />

q<br />

due to the same inductance paths in rotor d- and q-axes.<br />

The torque expression (2.52) can also be written in polar form using the current vector<br />

amplitude<br />

vector (Fig. 2.5.).<br />

I<br />

s<br />

and the torque angle δ I<br />

, i.e. angle between rotor d-axis and current<br />

19


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

q−<br />

axis<br />

I sq<br />

I s<br />

Ω s<br />

δ I<br />

d<br />

− axis<br />

I sd<br />

Ψ PM<br />

Figure 2.5. Stator current vector in rotor reference frame.<br />

For two current components using trigonometrical rules we can write:<br />

I = I cosδ<br />

(2.54a)<br />

sd<br />

s<br />

I<br />

I = I sinδ<br />

(2.54b)<br />

sq<br />

s<br />

I<br />

Substituting I , I into equation (2.52), the torque expression can be obtain as:<br />

sd<br />

sq<br />

M 3 1<br />

2<br />

e = b[ PM sin I ( q d) sin2 I]<br />

2 p Ψ I s δ −<br />

2<br />

L − L I s δ<br />

(2.55)<br />

<br />

M<br />

es<br />

For given current amplitude the synchronous and reluctances torque varies according to<br />

the sine <strong>of</strong> torque angle δ<br />

I<br />

. The variation <strong>of</strong><br />

M<br />

er<br />

M<br />

es<br />

and M<br />

er<br />

and resultant torque M<br />

e<br />

<strong>with</strong><br />

torque angle are illustrated in Fig. 2.6. The IPMSM parameters used for this calculation<br />

are given in the Appendices.<br />

e [ ] M Nm<br />

M [ ] er<br />

Nm M<br />

es<br />

[ Nm ]<br />

[deg] δ I<br />

Figure 2.6. Variation <strong>of</strong> synchronous torque M es<br />

, reluctance torque M<br />

er<br />

and resultant<br />

torque M as a function <strong>of</strong> torque angle (for rated current amplitude).<br />

e<br />

20


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

Referring to Fig. 2.7, stator flux components in rotor reference frame can be written as:<br />

where:<br />

Ψ = Ψ cos = LI +Ψ (2.56a)<br />

sd s<br />

δ Ψ d sd PM<br />

Ψ = Ψ sin = LI<br />

(2.56b)<br />

sq s<br />

δ Ψ q sq<br />

Ψ<br />

s<br />

is stator flux linkage amplitude, Ψ<br />

PM<br />

is rotor permanent magnet and δ Ψ<br />

is<br />

torque angle (angle between stator flux linkage vector and rotor permanent magnets flux<br />

vector).<br />

q−<br />

axis<br />

I sq<br />

Ψ sd<br />

Ψ PM<br />

I s<br />

Ω s<br />

Ψ s<br />

Ψ<br />

sq<br />

= LqI<br />

sq<br />

LI<br />

d<br />

sd<br />

δ Ψ<br />

I sd<br />

d<br />

− axis<br />

Ψ<br />

sd<br />

= LI<br />

d sd<br />

+ΨPM<br />

Figure 2.7. Rotor permanent magnet flux vector and stator flux linkage vector in rotor reference<br />

frame.<br />

From (2.56a) and (2.56b) the I<br />

sd<br />

and Isq<br />

can be obtained as:<br />

I<br />

I<br />

sd<br />

sd<br />

Ψs<br />

cosδ Ψ<br />

−ΨPM<br />

= (2.57a)<br />

L<br />

d<br />

Ψs<br />

sinδ = Ψ<br />

(2.57b)<br />

L<br />

q<br />

Substituting current components (2.57a), (2.57b) into equation (2.51), one can obtain<br />

another useful torque expressions:<br />

M<br />

e<br />

3 Ψ sin<br />

2 s<br />

ΨPM<br />

δΨ Ψs<br />

( Lq<br />

−Ld)sin2δ<br />

Ψ<br />

= pb[ −<br />

]<br />

2 Ld 2LdL<br />

<br />

q<br />

<br />

Mes<br />

M<br />

er<br />

(2.58)<br />

where:<br />

Ψ<br />

s<br />

stator flux linkage amplitude, and Ψ<br />

PM<br />

rotor flux, δ Ψ<br />

is torque angle, M<br />

es<br />

-<br />

synchronous torque,<br />

M<br />

er<br />

- reluctance torque.<br />

21


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

For the PM synchronous motor the amplitude <strong>of</strong> stator flux<br />

Ψ<br />

s<br />

is established by<br />

permanent magnet. Operation <strong>with</strong> stator flux amplitude belong the nominal value <strong>of</strong><br />

rotor flux amplitude<br />

Ψ<br />

PM<br />

increases the amplitude <strong>of</strong> stator phase current. Please note<br />

that maximum amplitude <strong>of</strong> the stator current vector is calculated as:<br />

higher value may damage the PM (complete demagnetization).<br />

I<br />

s<br />

Ψ<br />

≤<br />

L<br />

PM<br />

d<br />

, and<br />

From the Fig. 2.8 it can be observed that rated torque is achieved for torque angle<br />

0< < 25 <br />

electrical degree.<br />

δ Ψ<br />

e[ ] M Nm<br />

M<br />

er<br />

[ Nm ]<br />

M<br />

es<br />

[ Nm ]<br />

δ Ψ<br />

[deg]<br />

Figure 2.8. Variation <strong>of</strong> synchronous torque M es<br />

, reluctance torque M<br />

er<br />

and resultant<br />

torque M as a function <strong>of</strong> torque angle (for constant stator flux equal value <strong>of</strong> PM).<br />

e<br />

2.1.3 Mechanical motion equation<br />

The equation <strong>of</strong> rotor motion dynamics describes the mechanical equilibrium <strong>of</strong> a drive<br />

system. Taking the moment <strong>of</strong> inertia to be constant ( J = const.<br />

) and neglecting friction<br />

and elastic torque we can write:<br />

where,<br />

M<br />

e<br />

= Ml + Md<br />

(2.59)<br />

Ml<br />

is the external torque on the motor shaft, and<br />

M<br />

dΩ<br />

dt<br />

where: J is total moment <strong>of</strong> inertia,<br />

M<br />

d<br />

the dynamic torque<br />

m<br />

d<br />

= J<br />

(2.60)<br />

Ω<br />

m<br />

angular speed <strong>of</strong> the rotor.<br />

22


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

In general, for a drive system,<br />

J = J + J<br />

(2.61)<br />

m<br />

where: J<br />

m<br />

- motor inertia, J<br />

l<br />

load moments <strong>of</strong> inertia.<br />

l<br />

From equation (2.55) and (2.56) one can write:<br />

and, <strong>with</strong> (2.50),<br />

dΩ m 1 = ( M<br />

e − M<br />

l )<br />

(2.62)<br />

dt J<br />

dΩ m 1 3<br />

∗<br />

= ( pb<br />

Im( Ψ<br />

s<br />

Is<br />

) − Ml<br />

)<br />

(2.63)<br />

dt J 2<br />

Finally, the full mathematical model <strong>of</strong> PM synchronous machine which is used in<br />

simulation studies [Appendices] is described in dq , reference frame as:<br />

dΨ<br />

sd<br />

Usd = Rs Isd + − pbΩmΨ sq<br />

(2.64a)<br />

dt<br />

dΨ<br />

sq<br />

Usq = Rs Isq + + pbΩmΨ sd<br />

(2.64b)<br />

dt<br />

Ψ<br />

sd<br />

= LsdIsd +Ψ<br />

PM<br />

(2.65a)<br />

Ψ<br />

sq<br />

= LsqIsq<br />

(2.65b)<br />

dΩ m 1 = ( M<br />

e − M<br />

l )<br />

(2.66)<br />

dt J<br />

3 ∗ 3<br />

M<br />

e<br />

= pbIm( Ψ<br />

s<br />

Is) = pb( ΨsdIsq −Ψ<br />

sqIsd)<br />

(2.67)<br />

2 2<br />

Based on above equations we can create the block scheme <strong>of</strong> the PMSM machine (Fig.<br />

2.9), where the input signals are the voltage components in dq , reference frame<br />

U , U and the output signal is the mechanical speed <strong>of</strong> the rotor Ω<br />

m<br />

. As the external<br />

sd<br />

sq<br />

load torque<br />

M<br />

l<br />

is disturbance.<br />

23


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

M l<br />

U sd<br />

+<br />

−<br />

pbΩmΨsd<br />

∫<br />

Ψ sd<br />

Ψ PM<br />

−<br />

1<br />

L d<br />

I sd<br />

Ψ<br />

sq<br />

I<br />

sd<br />

p b<br />

R s<br />

−<br />

3<br />

p<br />

2 b<br />

M e<br />

−<br />

1<br />

J<br />

∫<br />

Ω m<br />

R s<br />

U sq<br />

−<br />

−<br />

∫<br />

Ψ sq<br />

1<br />

L q<br />

I sq<br />

Ψ<br />

sd<br />

I<br />

sq<br />

pbΩmΨsq<br />

p b<br />

Figure 2.9. Block scheme <strong>of</strong> PM synchronous machine in rotating dq , frame.<br />

Based on equations (2.64-2.67) we can also draw the vector diagram <strong>of</strong> PM<br />

synchronous motor (Fig. 2.10). From this vector representation it can see the positions<br />

<strong>of</strong> the vectors (currents, voltages and fluxes). Especially, power angle φ (angle between<br />

voltage and current vectors) and torque angle defined in two manners: as an angle<br />

between current and rotor flux vectors -δ I<br />

, or as angle between stator flux and rotor<br />

flux vectors -δ Ψ<br />

.<br />

q−axis<br />

β<br />

U s<br />

RI<br />

s<br />

s<br />

I s<br />

Ψ sq<br />

I sq<br />

LI<br />

d sd<br />

ΩΨ<br />

s<br />

s<br />

φ<br />

Ψ s<br />

LqI<br />

sq<br />

I sd<br />

δ I<br />

δ Ψ<br />

θ r<br />

Ψ sd<br />

θ Ψs<br />

Ψ PM<br />

d −axis<br />

rotor<br />

α ( A)<br />

stator<br />

Figure 2.10. <strong>Vector</strong> diagram <strong>of</strong> PM synchronous motor in rotor reference frame dq. ,<br />

24


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

2.2 Static characteristic under different control modes<br />

In this section, basic steady state properties <strong>of</strong> the PMSM under different control mode<br />

strategies will be study [6,9]. The key control strategies for the PMSM can be listed as<br />

follows:<br />

• Constant torque angle control (CTAC).<br />

• Maximum torque per ampere control (MTPAC)<br />

• Unity power factor control (UPFC)<br />

• Constant stator flux control (CSFC)<br />

Constant torque angle (CTA) control<br />

This control strategy for PMSM keeps the torque angle δ<br />

I<br />

(angle between stator current<br />

vector and rotor permanent magnet flux) at constant value 90 .<br />

q−<br />

axis<br />

Isq<br />

= I<br />

s<br />

δ = 90<br />

I<br />

<br />

d<br />

− axis<br />

Figure 2.11. Current vector and permanent magnet flux vector for constant torque angle<br />

operation (CTAC)<br />

Hence, this control can be achieved by controlling the d-axis current components to<br />

zero leaving the current vector on the rotor q-axis (see Fig. 2.11). Therefore, this<br />

strategy is also referred to as I<br />

sd<br />

= 0 control. The amplitude <strong>of</strong> rotor flux vector is<br />

constant and also the torque angle is constant. So, the torque depends only on the value<br />

<strong>of</strong> stator current amplitude. Therefore, this control strategy is not recommended for<br />

IPMSM <strong>with</strong> high saliency ratio. However, for SPMSM, this strategy is commonly<br />

used.<br />

Ψ PM<br />

The torque equation in this mode <strong>of</strong> operation becomes:<br />

3 3<br />

M<br />

e<br />

= pbΨ PMIsq = pbΨ PM<br />

Is<br />

(2.68)<br />

2 2<br />

25


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

The steady state voltage components based on the equations (2.34a) and (2.34b) are:<br />

Usd =−pbΩ mLq Iqs =−pbΩ mLq Is<br />

(2.69a)<br />

U = R I + p Ω Ψ = R I + p Ω Ψ (2.69b)<br />

sq s qs b m PM s s b m PM<br />

The amplitude <strong>of</strong> stator voltage vector can be calculated as:<br />

s<br />

2 2<br />

sd sq<br />

U = U + U<br />

(2.70)<br />

The stator flux vector amplitude can be calculated from equations (2.65a-b) as:<br />

s<br />

2 2<br />

sd sq<br />

Ψ = Ψ +Ψ (2.71)<br />

The active and reactive power and also the power factor can be obtained from equations<br />

(2.41),(2.46), (2.47).<br />

Maximum torque per ampere (MTPA) control<br />

The main idea <strong>of</strong> this control is develop the torque using minimum value <strong>of</strong> stator<br />

current amplitude. In this case the I sd<br />

components is not equal zero, and may cancel the<br />

reluctance torque produced by high saliency ratio. Therefore, this control strategy is<br />

recommended for IPMSM.<br />

q−<br />

axis<br />

I s<br />

I sq<br />

δ >= 90<br />

I<br />

<br />

d<br />

− axis<br />

I sd<br />

Figure 2.12. Current vector I s and permanent magnet flux vector Ψ<br />

PM<br />

for maximum torque<br />

per ampere operation (MTPAC).<br />

In order to obtain the maximum torque per ampere we should solve the derivative <strong>of</strong><br />

torque equations (2.55) in respect to torque angle. Solving for torque angle α and taking<br />

into account that only negative sign should be considered for the solution, we can<br />

calculate torque angle as:<br />

Ψ PM<br />

−1 −1 1 1 2<br />

δ<br />

I<br />

= cos [ − + ( ) ]<br />

4( L −L ) I 2 4( L −L ) I<br />

d q s<br />

d q<br />

s<br />

(2.72)<br />

26


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

From Fig. 2.8, it can be seen that<br />

M is maximum when torque angle is 90 < < 180<br />

e<br />

The relevant torque equation in this mode <strong>of</strong> operation becomes from (2.55).<br />

<br />

δ <br />

I<br />

.<br />

The steady state voltage equations can be written using the current vector amplitude<br />

I<br />

s<br />

and the torque angle δ<br />

I<br />

as:<br />

U = R I cosδ<br />

+ p Ω L I sinδ<br />

(2.73a)<br />

sd s s I b m q s I<br />

U = R I sinδ<br />

− p Ω L I cosδ<br />

+ p Ω Ψ (2.73b)<br />

sq s s I b m d s I b m PM<br />

The amplitude <strong>of</strong> stator voltage vector can be calculated from equation (2.70) and<br />

amplitude <strong>of</strong> stator flux vector from (2.71). The active and reactive power and also the<br />

power factor can be obtained from equations (2.41),(2.46), (2.47).<br />

Unity power factor (UPF) control<br />

Under this control strategy there is no phase different between the current vector and the<br />

voltage vector. Hence, power factor angle φ (see Fig. 2.13) becomes zero. Since only<br />

active power is supplied to the machine under unity power factor operation, the VA<br />

rating requirement <strong>of</strong> the inverter can be reduced.<br />

q − axis<br />

U s<br />

φ = 0<br />

I s<br />

δ I<br />

d − axis<br />

Figure 2.13. Current vector and permanent magnet flux vector under unity power factor<br />

operation (UPFC).<br />

Ψ PM<br />

In this case when φ = 0 we have the relationship:<br />

U<br />

U<br />

sq<br />

sd<br />

Isq<br />

= = tanδ<br />

I<br />

(2.74)<br />

I<br />

sq<br />

Substituting the voltage equations (2.69a-b) into (2.71) and made some simplifying, we<br />

can obtain:<br />

I L −L −Ψ + L I = (2.75)<br />

2<br />

s<br />

(<br />

d q)cos δI PM<br />

cosδI q s<br />

0<br />

27


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

Solving for the torque angle δ<br />

I<br />

:<br />

2 2<br />

PM PM<br />

4 Is<br />

( Ld Lq)<br />

Lq<br />

Ψ − Ψ − −<br />

−1<br />

δ<br />

I<br />

= cos [ ]<br />

2( L − L ) I<br />

d q s<br />

(2.76)<br />

only positive sign should be take into consideration.<br />

After obtaining δ I<br />

the amplitude <strong>of</strong> stator voltage vector can be calculated from<br />

equation (2.70) and amplitude <strong>of</strong> stator flux vector from (2.71). The active and reactive<br />

power and also the power factor can be obtained from equations (2.41),(2.46), (2.47).<br />

Constant stator flux (CSF) control<br />

As it can be see from the torque expression (2.58) for a given stator flux amplitude<br />

the electromagnetic torque<br />

amplitude<br />

M<br />

e<br />

is a function <strong>of</strong> torque angle δ Ψ<br />

. The stator flux linkage<br />

Ψ<br />

s<br />

is kept constant <strong>of</strong> the permanent magnet flux amplitude Ψ<br />

PM<br />

.<br />

I s<br />

q − axis<br />

Ψs<br />

Ψ<br />

s<br />

δ I<br />

δ Ψ<br />

d − axis<br />

Figure 2.14. Flux vector and permanent magnet flux vector under constant stator flux operation<br />

(CSFC).<br />

The amplitude <strong>of</strong> the stator flux linkage vector is<br />

Ψ PM<br />

2 2 2 2<br />

Ψ<br />

s<br />

= Ψ<br />

sd<br />

+Ψ<br />

sq<br />

= ( LI<br />

q sq<br />

) + ( LI<br />

d sd<br />

+Ψ<br />

PM<br />

)<br />

(2.77)<br />

Equating<br />

Ψ =Ψ (2.78)<br />

s<br />

PM<br />

can be obtain the relationship for rotor frame currents as:<br />

2 2<br />

q sq d sd d PM sd<br />

( LI ) + ( LI ) + 2LΨ I = 0<br />

(2.79)<br />

This condition is true if I<br />

sd<br />

< 0 , because expression<br />

always positive values.<br />

2 2<br />

q sq<br />

LI<br />

d sd<br />

( LI ) + ( ) and L , Ψ are<br />

d<br />

PM<br />

28


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

2 2 2 2 2 2<br />

d q s<br />

δI d PM s<br />

δI q s<br />

( L − L ) I cos + 2L Ψ I cos + L I = 0 (2.80)<br />

Solving for the torque angle δ<br />

I<br />

2 2 2 2 2 2<br />

d PM d PM s<br />

(<br />

d q<br />

)<br />

q<br />

−L Ψ ± L Ψ − I L −L L<br />

−1<br />

δ<br />

I<br />

= cos [ ]<br />

( L L ) I<br />

2 2<br />

d<br />

−<br />

q s<br />

(2.81)<br />

For givenδ I<br />

, the amplitude <strong>of</strong> stator voltage vector we can be calculated from equation<br />

(2.70) and amplitude <strong>of</strong> stator flux vector from (2.71). The active and reactive power<br />

and also the power factor can be obtained from equations (2.41),(2.46),(2.47),<br />

respectively for defined speed.<br />

Comparison study<br />

In order to compare the control strategies and to cancel dependence <strong>of</strong> machine power,<br />

per unit values defined as shown in Table 2.1 below have been introduced [3,9].<br />

The value <strong>of</strong> current vector:<br />

I<br />

sN<br />

= Is<br />

Is<br />

I<br />

= 2I<br />

(2.82)<br />

b<br />

srms( rated )<br />

The value <strong>of</strong> voltage vector: U<br />

s<br />

U<br />

s<br />

U<br />

sN<br />

= = (2.83)<br />

Ub ΩΨ<br />

b PM<br />

where: Ω b<br />

= 2 π f and b<br />

fb<br />

is rated frequency<br />

<strong>of</strong> the PM motor.<br />

The value <strong>of</strong> flux vector is: Ψ<br />

s<br />

Ψ<br />

s<br />

Ψ<br />

sN<br />

= =<br />

(2.84)<br />

Ψb<br />

ΨPM<br />

The value <strong>of</strong> torque is:<br />

Me<br />

Me<br />

M<br />

eN<br />

= =<br />

(2.85)<br />

M 3<br />

b pbΨ<br />

PMIb<br />

2<br />

The value <strong>of</strong> apparent power vector<br />

S S<br />

S = N<br />

S<br />

= 3<br />

(2.86)<br />

b UI<br />

b b<br />

2<br />

The value <strong>of</strong> active power<br />

P<br />

PN<br />

= (2.87)<br />

Sb<br />

The value <strong>of</strong> reactive power<br />

Q<br />

QN<br />

= (2.88)<br />

S<br />

b<br />

Table 2.1. Per unit values definition.<br />

In order to compare the steady state performance characteristic <strong>of</strong> the above discussed<br />

control strategies, for each <strong>of</strong> the control strategy some important quantities <strong>of</strong> the<br />

machine have been plotted as a function <strong>of</strong> the torque. The PMSM parameters, which<br />

are used for the calculations are given in Appendices.<br />

29


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

The current requirement versus torque is illustrated in Fig. 2.15 for the different control<br />

strategies. It can be seen, that up to 1 pu torque, the requirement for current is lowest for<br />

CSF control. Highest than 1 pu torque the low current needs MTPA control requirement<br />

lowest current for a given torque.<br />

3<br />

I<br />

sN<br />

[ pu]<br />

2.5<br />

2<br />

1.5<br />

UPF<br />

CSF<br />

CTA<br />

MTPA<br />

1<br />

0.5<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2<br />

M [ ] eN<br />

pu<br />

Figure 2.15. Stator current amplitude under different control strategies versus electromagnetic<br />

torque.<br />

The voltage requirement versus torque for the different control strategies is illustrated in<br />

Fig. 2.16. It can be seen, that CSF requires the highest value <strong>of</strong> stator voltage.<br />

U<br />

sN<br />

[ pu]<br />

2.5<br />

2<br />

1.5<br />

1<br />

CSF<br />

CTA<br />

MTPA<br />

0.5<br />

UPF<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2<br />

M [ ] eN<br />

pu<br />

Figure 2.16. Stator voltage amplitude versus electromagnetic torque under different control<br />

strategies (at 1 pu rotor speed).<br />

3<br />

PN<br />

2<br />

[ pu]<br />

1<br />

UPF<br />

CSF<br />

CTA<br />

MTPA<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2<br />

M [ ] eN<br />

pu<br />

Figure 2.17. Active power versus electromagnetic torque under different control strategies.<br />

30


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

The active power requirement as a function <strong>of</strong> torque is illustrated in Fig. 2.17 for the<br />

different control strategies. It can be seen, that all control strategies require<br />

approximately the same value <strong>of</strong> active power for a given torque. CSF control needs<br />

less active power in the region up to 1.3 pu torque.<br />

The reactive power requirement as a function <strong>of</strong> torque is illustrated in Fig. 2.18. It can<br />

be seen, that CTA control requires the highest value <strong>of</strong> active power for a given torque<br />

and the CSF control lowest.<br />

4<br />

QN<br />

[ pu]<br />

3<br />

2<br />

CTA<br />

MTPA<br />

1<br />

UPF<br />

CSF<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2<br />

M [ ] eN<br />

pu<br />

Figure 2.18. Reactive power versus electromagnetic torque under different control strategies.<br />

1.1<br />

1<br />

0.9<br />

UPF<br />

CSF<br />

cosφ<br />

0.8<br />

MTPA<br />

0.7<br />

0.6<br />

CTA<br />

0.5<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2<br />

M [ ] eN<br />

pu<br />

Figure 2.19. Power factor as a function <strong>of</strong> electromagnetic torque under different control<br />

strategies.<br />

The power factor as a function <strong>of</strong> torque is illustrated in Fig. 2.19. It can be seen, that as<br />

it could be expected, UPF control requires constant power factor for a given torque.<br />

CSF control is very close to the unity power factor up to 1 pu torque.<br />

The above analysis can be summarized as shown in Table. 2.2.<br />

31


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

Table. 2.2. Summary <strong>of</strong> voltage, power, power factor requirements under control modes.<br />

Requirement<br />

<strong>Control</strong><br />

method<br />

Voltage<br />

Current<br />

Power<br />

factor<br />

CTA middle low<br />

low<br />

MTP<br />

A<br />

low<br />

low<br />

middle<br />

UPF low high 1<br />

CSF high lowes up to 1.1<br />

pu torque<br />

Close to 1 up to<br />

1 pu torque<br />

From this comparison study it can be concluded that CSF control appears to be superior<br />

in terms <strong>of</strong> steady state performance characteristics compared to other methods under<br />

discussion.<br />

32


Modeling and control modes <strong>of</strong> PM synchronous motor drives<br />

2.3 Summary<br />

‣ There are different forms to express the PMSM equations, but the rotor<br />

reference frame equations are the most widely used. The simplification in rotor<br />

dq , reference frame equations results from the disappearance <strong>of</strong> position<br />

dependent inductances.<br />

‣ The electromagnetic torque <strong>of</strong> the IPMSM is not only produced by the<br />

permanent magnet flux, but also by the reluctance difference in rotor d- and q-<br />

axes.<br />

‣ Electromagnetic torque as cross vector product <strong>of</strong> the stator flux linkage and<br />

current space vectors or rotor and stator flux linkages is independent <strong>of</strong><br />

coordinate system selected. Therefore, can be expressed in stationary ( α,<br />

β ) or<br />

rotated ( dq) , coordinates.<br />

‣ For further control strategies consideration it is convenient to express the<br />

electromagnetic torque <strong>of</strong> PMSM machine by:<br />

• vector product <strong>of</strong> stator current and rotor flux vectors. The rotor flux<br />

vector in PMSM machine is constant, because <strong>of</strong> the PM. Therefore, to<br />

increase and decrease the torque, the current amplitude and the torque<br />

angle δ<br />

I<br />

should be changed (see Fig. 2.20a),<br />

• vector product <strong>of</strong> stator flux vector and rotor flux vector. Generally, the<br />

value <strong>of</strong> the stator flux amplitude is kept constant at value <strong>of</strong> rotor flux<br />

produced by permanent magnets. So, in this case to change the torque we<br />

should adjust the torque angle δ Ψ<br />

(see Fig. 2.20b).<br />

q<br />

β<br />

a )<br />

q β b)<br />

I s<br />

Ψ PM<br />

d<br />

Ψ s<br />

Ψ<br />

PM<br />

d<br />

δ I<br />

γ m<br />

α<br />

δ Ψ<br />

γ m<br />

α<br />

Fig. 2.20 <strong>Torque</strong> production: a) current control, b) flux control<br />

‣ Taking into account discussion regarding static characteristic under different<br />

control strategies it can be said that –depart from special requirements- the most<br />

suited for general application PMSM drives is constant stator flux (CSF)<br />

operation.<br />

33


Voltage source PWM inverter for PMSM supply<br />

Chapter 3 VOLTAGE SOURCE PWM INVERTER FOR PMSM SUPPLY<br />

3.1 Introduction<br />

The block scheme <strong>of</strong> an adjustable speed drive (ASD) commonly used in industrial<br />

applications to supply three-phase AC motor is presented in Fig. 3.1.<br />

Three-phase<br />

grid<br />

Rectifier<br />

DC link<br />

filter<br />

Inverter<br />

AC motor<br />

Choke<br />

Figure 3.1 Basic scheme <strong>of</strong> adjustable speed AC motor system.<br />

An ASD is supplied from three or single phase grid. It consists <strong>of</strong> a diode rectifier, DC<br />

link filter and an inverter. The rectifier converts supply AC voltage into DC voltage.<br />

The DC voltage is filtered by a capacitor in the DC link. The inverter converts the DC<br />

to an variable voltage, variable frequency AC for motor speed or (torque/current)<br />

control.<br />

The rectifier section <strong>of</strong> an ASD, called the front end, is responsible for generating<br />

current harmonics into the power supply system. Therefore, to reduce the total harmonic<br />

distortion (THD) <strong>of</strong> phase current it is necessary to add additional choke inductances.<br />

There are generally to way how insert choke inductances (see Fig. 3.2).<br />

a)<br />

b)<br />

L F<br />

D 1<br />

D3<br />

D5<br />

D 1<br />

D3<br />

D5<br />

L F<br />

L F<br />

C F<br />

U DC<br />

C F<br />

U DC<br />

L F<br />

D 2<br />

D4<br />

D6<br />

D 2<br />

D4<br />

D6<br />

L F<br />

Fig. 3.2 Three phase diode rectifier <strong>with</strong> smoothing choke: a) at the input b) at the DC link side.<br />

By adding a choke inductance at the input <strong>of</strong> rectifier gives the significant harmonic<br />

reduction. Some drive manufactures are starting to include this choke inductance in the<br />

DC link <strong>of</strong> the drive, providing the same harmonic current reduction benefit.<br />

Regarding to power electronics standards IEEE Std 519 it is recommended that<br />

production <strong>of</strong> harmonics should be less than 5%. It is a trend to replace the diode<br />

34


Voltage source PWM inverter for PMSM supply<br />

rectifier by fully controllable active rectifier [8] (see Fig. 3.3), which guaranties<br />

following futures:<br />

• power flow from AC/DC or DC/AC side (there is no need <strong>of</strong> break resistor ),<br />

• significant reduction <strong>of</strong> phase current THD,<br />

• unity power factor (phase voltage is in phase <strong>with</strong> current),<br />

• reduction <strong>of</strong> DC link capacitor,<br />

• controllable DC link voltage.<br />

Three-phase<br />

grid<br />

Active<br />

Rectifier<br />

DC link<br />

filter<br />

Inverter<br />

AC motor<br />

Choke<br />

Fig. 3.3 Modern AC/DC/AC converter topology <strong>of</strong> adjustable speed drives.<br />

In the next part <strong>of</strong> this thesis the author will be focus on voltage source inverter.<br />

3.2 Voltage source inverter (VSI)<br />

The made constant DC voltage by rectifier is delivered the to the input <strong>of</strong> inverter (Fig.<br />

3.4), which thanks to controlled transistor switches, converts this voltage to three-phase<br />

AC voltage signal <strong>with</strong> wide range variable voltage amplitude and frequency [3].<br />

Voltage Source Inverter<br />

C F<br />

T 1<br />

D 7<br />

T 3<br />

D 9<br />

T 5<br />

D 11<br />

U DC<br />

O<br />

C F<br />

T 2<br />

D 8<br />

T 4<br />

D 10<br />

T 6<br />

D 12<br />

A B C<br />

U<br />

sAN<br />

U<br />

sBN<br />

U<br />

sCN<br />

N<br />

Three-phase motor windings<br />

Figure 3.4 Basic scheme <strong>of</strong> voltage source inverter circuit.<br />

35


Voltage source PWM inverter for PMSM supply<br />

The one leg <strong>of</strong> inverter consists <strong>of</strong> two transistor switches. A simple transistor switch<br />

consist <strong>of</strong> feedback diode connected in anti-parallel <strong>with</strong> transistor. Feedback diode<br />

conducts current when the load current direction is opposite to the voltage direction.<br />

Assuming that the power devices are ideal: when they are conducting the voltage across<br />

them is zero and they present an open circuit in their blocking mode. Therefore, each<br />

inverter leg can be represented as an ideal switch. Its gives possibility to connect each<br />

<strong>of</strong> the three motor phase coils to a positive or negative voltage <strong>of</strong> the dc link ( U ).<br />

Thus the equivalent scheme for three-phase inverter and possible eight combinations <strong>of</strong><br />

the switches in the inverter are shown in Fig. 3.5.<br />

DC<br />

U<br />

7<br />

= 111<br />

U<br />

0<br />

= 000<br />

U DC<br />

1 1<br />

S<br />

A<br />

SB<br />

1<br />

S C<br />

U DC<br />

S<br />

A<br />

0 0<br />

SB<br />

0<br />

S C<br />

U<br />

1<br />

= 100<br />

A B C<br />

U<br />

2<br />

= 110<br />

A B C<br />

U<br />

3<br />

= 010<br />

U DC<br />

1<br />

SA<br />

0<br />

SB<br />

0<br />

S C<br />

U DC<br />

1 1<br />

S<br />

A<br />

SB<br />

0<br />

S C<br />

U DC<br />

0<br />

S<br />

A<br />

1<br />

SB<br />

0<br />

S C<br />

U<br />

4<br />

= 011<br />

A B C<br />

U<br />

5<br />

= 001<br />

A B C<br />

U<br />

6<br />

= 101<br />

A B C<br />

U DC<br />

0<br />

SA<br />

1<br />

SB<br />

1<br />

S C<br />

U DC<br />

0<br />

S<br />

A<br />

0<br />

SB<br />

1<br />

S C<br />

U DC<br />

1<br />

S<br />

A<br />

0<br />

SB<br />

1<br />

S C<br />

A B C<br />

A B C<br />

A B C<br />

Figure 3.5 Possible switches state in VSI.<br />

The six positions <strong>of</strong> switches ( U1 − U6) produce an output phase voltage equal ± 1/3 or<br />

± 2/3 <strong>of</strong> the DC voltage. The last two ( U0,<br />

U<br />

7) give zero output voltage. The output<br />

phase voltages produced by inverter are shown in Fig. 3.6a. and adequate line to line<br />

voltage calculated in bellow formula also are presented in Fig. 3.6b.<br />

UsAB = UsAN − UsBN<br />

(3.1a)<br />

UsBC = UsBN − UsCN<br />

(3.1b)<br />

UsCA = UsCN − UsAN<br />

(3.1c)<br />

36


Voltage source PWM inverter for PMSM supply<br />

a )<br />

b)<br />

U sAN<br />

U sAB<br />

2<br />

U<br />

3 DC<br />

U DC<br />

2<br />

− U<br />

3 DC<br />

2π<br />

Ωt<br />

−U DC<br />

2π<br />

Ωt<br />

2<br />

U<br />

3 DC<br />

U sBN<br />

U DC<br />

U sBC<br />

2<br />

− U<br />

3 DC<br />

2π<br />

Ωt<br />

−U DC<br />

2π<br />

Ωt<br />

2<br />

U<br />

3 DC<br />

U sCN<br />

U DC<br />

U sCA<br />

2<br />

− U<br />

3 DC<br />

U1<br />

U<br />

2<br />

U3<br />

U<br />

4<br />

U5<br />

U6<br />

2π<br />

Ωt<br />

−U DC<br />

a )<br />

b)<br />

U1<br />

U<br />

2<br />

U3<br />

U<br />

4<br />

U5<br />

U6<br />

Figure. 3.6 Three voltage waveforms generated by the inverter: a) phase voltages, b) line to line<br />

voltages.<br />

2π<br />

Ωt<br />

Form the Fourier analysis for phase voltage produced by inverter (Fig. 3.7) the<br />

maximum amplitude <strong>of</strong> fundamental phase voltage for a given DC link voltage is given<br />

by:<br />

U<br />

_ amp<br />

2<br />

= UDC<br />

(3.2)<br />

π<br />

U out<br />

2<br />

U<br />

3 DC<br />

1<br />

U<br />

3 DC<br />

1<br />

− U<br />

3 DC<br />

2<br />

− U<br />

3 DC<br />

2<br />

U DC<br />

π<br />

2π<br />

ωt<br />

Figure 3.7 Inverter phase voltage generated during six step operation (solid line), corresponding<br />

fundamental component <strong>of</strong> output voltage (dashed line) and harmonic spectrum <strong>of</strong> phase<br />

voltage.<br />

37


Voltage source PWM inverter for PMSM supply<br />

The three-phase output voltage <strong>of</strong> the inverter can be described by space vector<br />

definition as:<br />

U<br />

k<br />

⎧2<br />

⎪ U<br />

3<br />

⎪<br />

= ⎨<br />

⎪⎪⎪ 0<br />

⎩<br />

DC<br />

e<br />

π<br />

j( k−1) 3<br />

for k =1,2...,6.<br />

for k=0,7.<br />

, (3.3)<br />

where k denotes numbers vector.<br />

<strong>Vector</strong>s from 1-6 are called active vectors, whereas vectors 0,7 are called zero vectors<br />

or non active vectors. The voltage space vector<br />

U<br />

k<br />

in complex plane forms a regular<br />

hexagon and divides in into six equal sectors (one sector takes 60 electrical degree) Fig.<br />

3.8.<br />

Im<br />

U 4<br />

(011)<br />

U 3 (010)<br />

sector3<br />

sector4<br />

sector2<br />

sector5<br />

U (110) 2<br />

sector1<br />

U (000) U 1<br />

(100)<br />

0 U (111) 7 Re<br />

sector6<br />

2<br />

U<br />

3 DC<br />

U (001)<br />

U (101)<br />

5 6<br />

Figure 3.8 Representation <strong>of</strong> the inverter states in the complex space.<br />

In practice the real voltage source inverter has non-linear characteristic due to [19]:<br />

• the dead-time,<br />

• a voltage drop across the power switches,<br />

• pulsation <strong>of</strong> the DC link voltage.<br />

Dead time effect [17,20,27,30]<br />

Semiconductors power switches <strong>of</strong> voltage source inverter operate not ideally. They do<br />

not turn-on or turn-<strong>of</strong>f instantaneously. Therefore it is necessary to include a protection<br />

time to avoid a short circuit in the DC link, when two switching devices are in the same<br />

leg (see Fig. 3.9). This time T d<br />

is included in the control signals and it is called “dead<br />

38


Voltage source PWM inverter for PMSM supply<br />

time”. It guarantees safe operation <strong>of</strong> the inverter. The typical value is from 1µ s - 5µ s .<br />

When the lowest value is for small power IGBT and is growing in respect to increasing<br />

<strong>of</strong> IGBT power. More details about real IGBT module you can find in Appendices.<br />

The effect <strong>of</strong> dead time can be examined from one phase <strong>of</strong> PWM inverter. The basic<br />

configuration is shown in Fig. 3.9. Consist <strong>of</strong> upper and lower power devices T<br />

1<br />

and T<br />

2<br />

,<br />

and reverse recovery diodes D<br />

1<br />

and D<br />

2<br />

, connected between the positive and negative<br />

rails <strong>of</strong> power supply. The gating signals S<br />

A<br />

and S<br />

Ai<br />

come from control block. Output<br />

voltage terminal U<br />

0<br />

is connected to motor phase.<br />

S A<br />

T 1<br />

D 1<br />

U DC<br />

I sA<br />

S Ai<br />

T 2<br />

D 2<br />

LOAD<br />

U 0<br />

S<br />

Dead time<br />

Td<br />

Figure 3.9 Circuit diagram <strong>of</strong> one inverter leg.<br />

Fig. 3.10 shows the ideal control signals and real control signals <strong>with</strong> inserted dead time<br />

T<br />

d<br />

. As can be observed the time duration <strong>of</strong> real drive signal for upper transistor is<br />

shorted than ideal drive signal and for lower transistor is longer than ideal.<br />

Ideal drive signals<br />

Real drive signals<br />

S A<br />

S A<br />

T d<br />

S Ai<br />

S Ai<br />

T d<br />

Figure 3.10 Gate signals control <strong>of</strong> one inverter leg.<br />

39


Voltage source PWM inverter for PMSM supply<br />

As a consequence when the phase current I sA<br />

is positive, the output voltage is reduced,<br />

and when the current I<br />

sA<br />

is negative the output voltage is increased (see Fig. 3.11).<br />

U0<br />

I<br />

sA<br />

> 0<br />

ideal voltage<br />

real voltage<br />

U I < 0<br />

0<br />

sA<br />

U DC<br />

2<br />

U DC<br />

2<br />

−U DC<br />

2<br />

deacresing<br />

−U DC<br />

2<br />

increasing<br />

Figure 3.11 Dead time effect on the inverter output voltage: (fat line real voltage, doted line<br />

ideal voltage).<br />

Voltage drop across power devices<br />

In real voltage source inverter power switches do not conduct ideally. When they are<br />

conducting the voltage across them is not zero and equal the voltage drop on the<br />

conducted transistor V T<br />

. Also in blocking mode the power switches have voltage drop<br />

on the conducted diode V D<br />

. More details about real IGBT module you can find in<br />

Appendices.<br />

The voltage drop across the power devices is dependent on the direction <strong>of</strong> the phase<br />

current. It has influence on the output voltage, especially at low speed operation <strong>of</strong><br />

motor and high load current [17,20,27,30]. Fig. 3.12 shows the voltage drop influence<br />

on the output voltage. Also shows that the output voltage is asymmetric (<strong>with</strong> <strong>of</strong>fset)<br />

and the voltage drop decreases the output voltage when the phase current is positive and<br />

increases the output voltage when the phase current is negative.<br />

U 0<br />

I > 0<br />

sA<br />

U 0<br />

I < 0<br />

sA<br />

U DC<br />

2<br />

V T<br />

U DC<br />

2<br />

V D<br />

V D<br />

V T<br />

−U DC<br />

2<br />

−U DC<br />

2<br />

Figure 3.12 Output voltage in voltage source inverter due to voltage drop across the power<br />

devices a) for I<br />

sA<br />

> 0 , b) I<br />

sA<br />

< 0 (fat line real voltage, doted line ideal voltage).<br />

The influence <strong>of</strong> dead time effect and voltage drop across power devices on the output<br />

voltage from inverter is illustrated in block diagram (Fig. 3.13). The ideal reference<br />

*<br />

voltage components in stationary reference frame ( U α<br />

,<br />

_ ideal<br />

*<br />

U β _ ideal<br />

) are equal real<br />

40


Voltage source PWM inverter for PMSM supply<br />

*<br />

( U α<br />

,<br />

_ real<br />

*<br />

U β _ real<br />

) and delivered to pulse width modulation (PWM) modulator block<br />

<strong>with</strong> real non-linear inverter. As a result the output voltages ( U α _ out<br />

, U β _ out<br />

) are distorted<br />

(green signals) and as consequence the phase currents ( I α _ out<br />

, I β _ out<br />

) in the load (red<br />

signals) are also distorted.<br />

*<br />

U *<br />

α _ideal<br />

U α _ real<br />

*<br />

*<br />

U β<br />

U<br />

_ideal<br />

β _ real<br />

PWM<br />

Modulator<br />

S<br />

A<br />

S B<br />

S C<br />

Inverter<br />

U α _out<br />

U β _out<br />

I α _ out<br />

I β _out<br />

Load<br />

Figure 3.13 Block diagram illustrated the dead time effect and voltage drop across power<br />

devices in three phase motor supplied from non-ideal voltage source inverter.<br />

Pulsation <strong>of</strong> the DC link voltage [19]<br />

In practice it should be take into account that the real input dc-link voltage required for<br />

supply VSI is not ideal. It has ripples and fluctuation, because <strong>of</strong> not ideal filtering and<br />

disadvantages <strong>of</strong> diode rectifier. Therefore, the quality <strong>of</strong> dc-link voltage has impact on<br />

the output voltage from inverter. If dc-link voltage will change we can observed the<br />

changing at the output <strong>of</strong> inverter. In order to overcome this problem:<br />

• in PWM modulator we can not assume a constant dc-link voltage and we should<br />

measured this voltage in order to calculate the modulation index (see subchapter<br />

3.3),<br />

41


Voltage source PWM inverter for PMSM supply<br />

• instate <strong>of</strong> diode rectifier will be use the active rectifier, which provided<br />

controllable DC-link voltage.<br />

• or used bigger capacitor in the DC-link side in order to increase possibility <strong>of</strong><br />

filtering.<br />

Let us summarize, adding influence <strong>of</strong> non-linear VSI causes by:<br />

• serious distortion in the inverter output voltage,<br />

• distorted machine currents,<br />

• torque pulsation,<br />

Additionally, also causes motor instability due to the interaction between motor and the<br />

PWM inverter, or the choice <strong>of</strong> the PWM strategy [25].<br />

Based on simulated and experimental observation one can say that the dead time effect<br />

is:<br />

• more visible in low speed operation <strong>of</strong> the motor,<br />

• may become significantly in drives where high switching frequency is required<br />

for good dynamics performances.<br />

In some applications such as sensor-less vector control, the inverter output voltages are<br />

needed to calculate the rotor or stator flux vectors. Unfortunately, it is very difficult to<br />

measure the output voltage and requires additional hardware. The most desirable<br />

method to obtain the output voltage feedback signal is to use the reference voltages<br />

instead. However, the relation between the output and reference voltage is nonlinear due<br />

to the dead-time effect and voltage drop across power devices. Thus, unless the properly<br />

dead-time and voltage drop compensation will be applied, the reference voltage can not<br />

be used instead <strong>of</strong> the inverter output voltage. Several compensation method were<br />

proposed to overcome this problem. One <strong>of</strong> them will be present bellow.<br />

Compensation based on modification <strong>of</strong> reference voltage waveform [17]<br />

The compensation process <strong>of</strong> dead time effect and voltage drop across power devices on<br />

the inverter output voltage from is illustrated in Fig. 3.14.<br />

42


Voltage source PWM inverter for PMSM supply<br />

Compensation<br />

block<br />

IsA<br />

IsB<br />

Compensation <strong>of</strong><br />

inverter<br />

I sC<br />

U DC<br />

T comp<br />

U α _comp<br />

U β _comp<br />

*<br />

U α _ideal<br />

*<br />

U β _ideal<br />

*<br />

U α _ real<br />

*<br />

U α _ real<br />

*<br />

U β _ real<br />

PWM<br />

Modulator<br />

S<br />

A<br />

S B<br />

S C<br />

Inverter<br />

*<br />

U β _ real<br />

U α _out<br />

U β _out<br />

I α _out<br />

I β _out<br />

Load<br />

Figure 3.14 Block diagram illustrating the dead time and voltage drop across power devices<br />

compensation method in three phase motor supplied from non-ideal voltage source inverter.<br />

In order to compensate the inverter non-linearity to the ideal reference voltage<br />

*<br />

components in stationary reference frame ( U α<br />

,<br />

(<br />

_ comp<br />

_ ideal<br />

*<br />

U β _ ideal<br />

) a compensation signal<br />

U α<br />

, U β _ comp<br />

) is added. As a consequence the real reference voltage components<br />

*<br />

( U α<br />

,<br />

_ real<br />

*<br />

U β _ real<br />

) are pre-distorted. Further those signals are delivered to PWM<br />

modulator <strong>with</strong> non-ideal inverter. As a result the output voltages are not distorted in<br />

Fig. 3.14 and thus phase currents in the load (red signals in Fig. 3.14) are almost<br />

sinusoidal.<br />

To calculate an average compensation voltages ( U α _ comp<br />

, U β _ comp<br />

), the parameters <strong>of</strong><br />

IGBT modules as:<br />

• dead time T<br />

d<br />

,<br />

• turn on T<br />

ON<br />

and turn <strong>of</strong>f T<br />

OFF<br />

<strong>of</strong> IGBT transistors,<br />

• and also on a voltage drop on diode V<br />

D<br />

and transistor<br />

should be know.<br />

V<br />

T<br />

,<br />

43


Voltage source PWM inverter for PMSM supply<br />

The total compensation time to compensate the non-linearity <strong>of</strong> inverter can be<br />

calculated as:<br />

U<br />

T = T + T − T + T<br />

(3.4)<br />

dp<br />

comp d ON OFF s<br />

U<br />

DC<br />

Where Udp = VD −ton ( VD − VT )/ Ts<br />

and t on<br />

is conducting time <strong>of</strong> IGBT devices in one<br />

sampling time.<br />

The compensation voltage vector can be obtained as:<br />

Tcomp<br />

Ucomp = 2 UDC s ign( Is ) = 2Uth s ign( Is<br />

)<br />

(3.5)<br />

Ts<br />

where<br />

and<br />

2<br />

2<br />

sign( Is) = ( sign( IsA) + asign( IsB) + a sign( IsC))<br />

,<br />

3<br />

⎧ sign( IsA) = 1 if IsA<br />

> 0<br />

sign( IsA)<br />

= ⎨<br />

⎩sign( IsA) = 0 if IsA<br />

< 0<br />

(3.6)<br />

The sign function for remain phase currents are calculated similarly.<br />

Solving equations (3.5) for real and imagine part in stationary frame, one obtains:<br />

1<br />

U U sign I sign I sign I<br />

3<br />

α _ comp<br />

= 2<br />

th<br />

(2 (<br />

sA) −0.5 (<br />

sB<br />

) − 0.5 (<br />

sC<br />

)) (3.7a)<br />

1<br />

U U sign I sign I<br />

3<br />

β _ comp<br />

= 2<br />

th<br />

( (<br />

sB<br />

) − (<br />

sC<br />

))<br />

(3.7b)<br />

The waveform <strong>of</strong> compensation voltages in stationary frame are shown in Fig. 3.15.<br />

44


Voltage source PWM inverter for PMSM supply<br />

4<br />

U<br />

3 th<br />

8<br />

U<br />

3 th<br />

2<br />

U<br />

3 th<br />

Figure 3.15 Voltage compensation components in stationary reference frame.<br />

From the top reference α and β components.<br />

In order to illustrate the effectiveness <strong>of</strong> the proposed compensation a simulation study<br />

has been performed. Fig. 3.16a shows the phase current in α,<br />

β frame and their<br />

hodograph <strong>with</strong>out compensation and Fig. 3.16b <strong>with</strong> proposed compensation method.<br />

a)<br />

a)<br />

b)<br />

b)<br />

Figure 3.16 Nonlinearity effect <strong>of</strong> voltage source inverter on phase current <strong>of</strong> AC machine:<br />

a) <strong>with</strong>out compensation, b) <strong>with</strong> compensation.<br />

45


Voltage source PWM inverter for PMSM supply<br />

3.3 <strong>Space</strong> vector based pulse width modulation (PWM) methods<br />

In voltage source inverter the transistors are controlled in a on-<strong>of</strong>f fashion. In order to<br />

obtain a suitable duty cycle for each switches the technique pulse <strong>with</strong> modulation is<br />

used. The modulation methods [18,21,22,23,24,26,28,29,30] have the influence on:<br />

wide range <strong>of</strong> linear operation, low content <strong>of</strong> higher harmonics in voltage and current,<br />

low frequency harmonics, minimal number <strong>of</strong> switching to decrease switching losses in<br />

the power components.<br />

The most important factor in PWM mode is modulation index defined as the ratio <strong>of</strong> the<br />

reference voltage amplitude value to the maximum voltage amplitude value during sixstep<br />

operation (see Fig 3.17) and is given by:<br />

M<br />

U<br />

ref<br />

= (3.8)<br />

2<br />

U<br />

DC<br />

π<br />

where the<br />

U<br />

DC<br />

is the DC link voltage (for three phase six diodes rectifier is 560 V ).<br />

The modulation index varies between 0-1 and can be divided into two regions: the<br />

linear ( 0< M ≤ 0.907) and the nonlinear ( 0.907 < M ≤ 1) as is shown in Fig. 3.17.<br />

Im<br />

U3(010)<br />

U 2(110)<br />

End <strong>of</strong> overmodulation<br />

region (six-step mode)<br />

2<br />

π<br />

Umax<br />

= U DC<br />

⇒ M = 1<br />

sector = 2<br />

sector = 3<br />

2<br />

U<br />

3 dc<br />

U ref<br />

End <strong>of</strong> linear region<br />

U<br />

max<br />

= U DC<br />

⇒ M = 0.907<br />

3<br />

U 4<br />

(011)<br />

sector = 4<br />

U 0<br />

(000)<br />

U 7<br />

(111)<br />

sector = 5<br />

θ ref<br />

T1<br />

U<br />

1<br />

Ts<br />

sector = 1<br />

sector = 6<br />

Re<br />

U (100) 1<br />

nonlinear<br />

(overmodulation) region<br />

linear region<br />

U 5 (001)<br />

U 6 (101)<br />

Figure 3.17 <strong>Space</strong> vector diagram <strong>of</strong> the available switching vectors.<br />

46


Voltage source PWM inverter for PMSM supply<br />

Linear range <strong>of</strong> operation ( 0> M


Voltage source PWM inverter for PMSM supply<br />

Next from<br />

U<br />

ref<br />

, α<br />

ref<br />

it is necessary to calculate the time interval for particular vectors.<br />

U 2<br />

U ref<br />

120°- α<br />

UU<br />

0, 7<br />

t 0<br />

t 2<br />

α ref<br />

t 1<br />

Using the low <strong>of</strong> sine it is possible to write:<br />

U 1<br />

120°<br />

Figure 3.18 One sector in voltage plane.<br />

U<br />

ref U1 U2<br />

= =<br />

sin120° sin(60 °−α<br />

) sinα<br />

ref<br />

ref<br />

(3.12)<br />

From this relations calculated value <strong>of</strong> vectors<br />

sin(60 °−αref<br />

) 2<br />

U1<br />

= Uref = Uref sin(60 °−αref<br />

)<br />

sin120°<br />

3<br />

(3.13a)<br />

sin<br />

ref 2<br />

U2<br />

= U α = U<br />

sin120°<br />

3<br />

sinα<br />

ref ref ref<br />

(3.13b)<br />

and respectively the normalized times are given:<br />

3 U<br />

1<br />

ref<br />

1<br />

= = sin(60 °− αref<br />

)<br />

(3.14a)<br />

2<br />

U<br />

U<br />

DC<br />

DC<br />

t<br />

3<br />

U<br />

3 U<br />

2<br />

ref<br />

2<br />

= = sinα<br />

ref<br />

(3.14b)<br />

2<br />

U<br />

U<br />

DC<br />

DC<br />

t<br />

3<br />

U<br />

Putting Eq. (3.13a-b) in to Eq. (3.14a-b) the normalized value can be presented as:<br />

2 3<br />

t1<br />

= Msin(60 °− α ref<br />

)<br />

(3.15a)<br />

π<br />

2 3<br />

t2<br />

= Msinα<br />

ref<br />

(3.15b)<br />

π<br />

or in other form:<br />

48


Voltage source PWM inverter for PMSM supply<br />

2 3<br />

t2<br />

= Msinα<br />

ref<br />

(3.16a)<br />

π<br />

3 1<br />

t1 = M cosαref<br />

− t2<br />

(3.16b)<br />

π<br />

2<br />

After t 1<br />

and t 2<br />

calculation, the remaining normalized time is reserved for zero vectors<br />

U<br />

0<br />

, U7<br />

<strong>with</strong> condition t 1<br />

+ t 2<br />

≤ 1.Therefore, the normalized total time for zero vectors<br />

becomes:<br />

t = t + t = 1 − ( t + t )<br />

(3.17)<br />

07 0 7 1 2<br />

The equations (3.15a-b) for time interval <strong>of</strong> active vectors and equation (3.17) for total<br />

time interval <strong>of</strong> zero vectors are identical for all variants <strong>of</strong> space vector modulation<br />

(<strong>SVM</strong>) techniques.<br />

The absence <strong>of</strong> neutral wire in star connected load provides a degree <strong>of</strong> freedom in<br />

selecting the partitioning (zero sequence signals -ZSS) time <strong>of</strong> the two zero vectors. It is<br />

equivalent to the freedom <strong>of</strong> injected signals in to phase signals. Therefore, it gives<br />

different equations <strong>of</strong> t 0<br />

and t 7<br />

for each PWM method, but normalized duration time <strong>of</strong><br />

must fulfill condition in Eq. 3.17. As a consequence is only in different placement <strong>of</strong><br />

zero vectors U<br />

0<br />

, U<br />

7<br />

. Therefore, we can introduce the portioning factor <strong>of</strong> zero vectors,<br />

which is defined as:<br />

t t<br />

k = =<br />

t + t t<br />

7 7<br />

0 7 07<br />

(3.18)<br />

Please note that, the zero sequence signals does not change the inverter output line-toline<br />

voltage.<br />

From knowledge <strong>of</strong> the neutral voltage U<br />

N 0<br />

(see Fig. 3.4) and information what kind <strong>of</strong><br />

zero sequence signal (ZSS) will be injected in each phase <strong>of</strong> motor it is possible to<br />

calculate normalized duration time <strong>of</strong> zero vectors t 0<br />

and t 7<br />

. In bellow Table 3.1 are<br />

summarized different three-phase modulation techniques and remarks. Also graphical<br />

interpretation are shown in Fig. 3.19.<br />

49


Voltage source PWM inverter for PMSM supply<br />

Zero Sequence Signal Time interval <strong>of</strong> zero vectors Remark<br />

SPWM<br />

ZSS =0<br />

for even sectors<br />

4<br />

t0<br />

= (1 − ( M cos θ )) / 2<br />

π<br />

for odd sectors<br />

4<br />

t0 = t07<br />

−(1 −( M cos θ )) / 2<br />

π<br />

and t7 = t07 − t0<br />

U<br />

DC<br />

/2 π<br />

M = = = 0.785<br />

2U<br />

DC 4<br />

π<br />

complicated calculation<br />

for time interval <strong>of</strong><br />

vectors<br />

THIPWM<br />

ZSS =sinusoidal signal<br />

<strong>with</strong> triple harmonic<br />

for even sectors<br />

4 1<br />

t0<br />

= (1 −( M cos θ) − cos 3 θ)) / 2<br />

π 6<br />

for odd sectors<br />

4 1<br />

t0 = t07<br />

−(1 −( M cos θ) − cos 3 θ)) / 2<br />

π 6<br />

and t7 = t07 − t0<br />

U<br />

DC<br />

/ 3 π<br />

M = = = 0.907<br />

2U<br />

DC 2 3<br />

π<br />

Increase the linearity <strong>of</strong><br />

inverter grater than 15%<br />

<strong>of</strong> SPWM<br />

SVPWM<br />

ZSS=triangle signal<br />

<strong>with</strong> triple harmonic<br />

0.906<br />

for all sectors<br />

M =<br />

t = t t t /2<br />

time interval<br />

7 07 0 07 t = t /2<br />

0 07 Simple calculation for<br />

k=0.5<br />

Table 3.1. Variants <strong>of</strong> three-phase space vector modulation techniques.<br />

a) b)<br />

c)<br />

the portioning<br />

factor <strong>of</strong> zero vectors<br />

Fig. 3.19 PWM techniques <strong>with</strong> various zero signal sequence shape: a) SPWM, b) THIPWM,<br />

c) SVPWM. The upper part <strong>of</strong> figure: phase voltage U<br />

AN<br />

(green), pole voltage U<br />

AO<br />

(blue),<br />

voltage between neutral points U<br />

NO<br />

(black). The lower part <strong>of</strong> figure shows the portioning<br />

factor <strong>of</strong> zero vectors for all PWM techniques.<br />

50


Voltage source PWM inverter for PMSM supply<br />

Except SPWM technique all PWM method guarantee that the ZSS extends the range <strong>of</strong><br />

modulation index from 0.78 to 0.906, i.e. 15% greater than that obtained <strong>with</strong> standard<br />

version <strong>of</strong> SPWM.<br />

The duty time cycles in sector 1 for each phase can be written:<br />

d = t + t + kt<br />

(3.19a)<br />

a<br />

b<br />

1 2 07<br />

d = t + kt<br />

(3.19b)<br />

dc<br />

2 07<br />

= kt<br />

(3.19c)<br />

07<br />

for all sectors the duty time calculations for each phase can be calculated:<br />

sec tor1 sec tor 2 sector 3 sec tor 4 sector5 sec tor 6<br />

⎡<br />

⎤<br />

⎡d 1 1 1 0 0 0 0 0 0 1 1 1<br />

a ⎤ k k k k k k t<br />

⎢<br />

⎥ ⎡ 1 ⎤<br />

⎢<br />

d<br />

⎥<br />

= ⎢0 1 k 1 1 k 1 1 k 1 0 k 0 0 k 0 0 k ⎥ ⎢<br />

t<br />

⎥<br />

b<br />

2<br />

⎢ ⎥ ⎢<br />

⎥ ⎢ ⎥<br />

⎢⎣d ⎥ 0 0 k 0 0 k 0 1 k 1 1 k 1 1 k 1 0 k ⎢t<br />

⎥<br />

c ⎦ ⎣ 07 ⎦<br />

⎢<br />

⎣<br />

T<br />

⎥<br />

⎦<br />

(3.20)<br />

Depending on the location <strong>of</strong> the space vector, the basic vectors must be chosen in order<br />

to get the minimum number <strong>of</strong> changes in the switches <strong>of</strong> the converter. The switching<br />

sequence for each sector and suitable pulse pattern for first sector are shown in Fig.<br />

3.20.<br />

Sector<br />

Three-phase <strong>Modulation</strong><br />

T S<br />

/2<br />

T S<br />

t<br />

0 t1<br />

2 7<br />

t<br />

t<br />

1 U0, U1, U2, U7, U2, U1,<br />

U<br />

0<br />

2 U0, U3, U2, U7, U2, U3,<br />

U<br />

0<br />

3 U0, U3, U4, U7, U4, U3,<br />

U<br />

0<br />

4 U0, U5, U4, U7, U4, U5,<br />

U<br />

0<br />

5 U0, U5, U6, U7, U6, U5,<br />

U<br />

0<br />

6 U0, U1, U6, U7, U6, U1,<br />

U<br />

0<br />

0 1 1 1 1 1 1 0<br />

0 0 1 1 1 1 0 0<br />

0 0 0 1 1 0 0 0<br />

U0<br />

U1<br />

U2<br />

U7<br />

U<br />

2<br />

U1<br />

U0<br />

d A<br />

d B<br />

d C<br />

Fig. 3.20 Switching sequence for three-phase PWM techniques (on the left ) and pulse pattern<br />

<strong>of</strong> three-phase vector modulator in sector 1 (on the right).<br />

51


Voltage source PWM inverter for PMSM supply<br />

3.4 Summary<br />

The general conclusions from this chapters can be summarized as follows:<br />

‣ to supply the voltage source inverter can be used diode rectifier or active<br />

rectifier <strong>with</strong> IGBT transistors,<br />

‣ supplied PMSM machine from VSI do not required mechanical comutator. It is<br />

thanks to electronic commutation. This overcome the problem <strong>with</strong> brushes and<br />

periodical service,<br />

‣ The voltage source inverter is non-linear power amplifier in respect to:<br />

o dead time effect,<br />

o voltage drop across the power devices,<br />

o DC link voltage pulsation.<br />

‣ Without appropriate dead-time and voltage drop compensation, the sensorless<br />

operation in low speed range is not possible.<br />

‣ The quality <strong>of</strong> DC link voltage has influence on proper operation <strong>of</strong> AC drive,<br />

‣ Sinusoidal modulation technique (SPWM) guarantees the inverter output voltage<br />

U<br />

DC<br />

amplitude from 0 until V. This correspond to changes modulation index<br />

2<br />

from 0-0.785.<br />

‣ <strong>Modulation</strong> techniques <strong>with</strong> zero sequence signals not equal zero guarantees the<br />

U<br />

DC<br />

inverter output voltage amplitude from 0 until V (it is 15% grater than<br />

3<br />

SPWM). This correspond to changes modulation index from 0-0.907.<br />

‣ In this work the PWM modulator <strong>with</strong> zero sequence signal <strong>of</strong> triple harmonics<br />

(SVPWM) will be used. Mainly because <strong>of</strong> simple calculation <strong>of</strong> zero vector<br />

duration and placement.<br />

52


<strong>Control</strong> methods <strong>of</strong> PM Synchronous motor<br />

Chapter 4 CONTROL METHODS OF PM SYNCHRONOUS MOTOR<br />

4.1 Introduction<br />

The basic block scheme <strong>of</strong> adjustable speed drive <strong>with</strong> control block for PMSM is<br />

presented in Fig. 4.2. It consists <strong>of</strong> two parts: power (fat line) and control part<br />

employed microprocessor (thin line). The first one previously have been explained in<br />

chapter 3. The second one will be described bellow.<br />

Figure 4.2. The basic block scheme <strong>of</strong> PMSM drive supplied voltage source inverter.<br />

The main task <strong>of</strong> control block is follow demand reference speed by motor and provide<br />

proper operation in static (insight <strong>of</strong> the limits) and dynamic states <strong>with</strong>out any<br />

instability. This is ensured through suitable generated gate signals for the IGBT<br />

transistor inside <strong>of</strong> the inverter. Therefore, to make good decision how to control power<br />

transistors in the inverter, the following feedback signals are measured and used:<br />

• DC link voltage,<br />

• motor phase currents,<br />

• speed or position <strong>of</strong> the rotor.<br />

This significantly improve dynamic behavior <strong>of</strong> the system (good performance <strong>of</strong> the<br />

torque and speed response, very fast dynamics response <strong>with</strong> fully controllable torque in<br />

wide speed range).<br />

The scalar control for PMSM <strong>with</strong>out damper winding (squire cage) is not simple as for<br />

induction motor [39,40]. It requires additional stabilization loop, which can be provide<br />

by feedback loop from: rotor velocity perturbation, active power or DC-link current<br />

perturbation [9].<br />

The vector control method will be described bellow.<br />

53


<strong>Control</strong> methods <strong>of</strong> PM Synchronous motor<br />

4.2 Field oriented control (FOC)<br />

During many years a DC motor has been mostly used. Because <strong>of</strong> simple control<br />

method, which based on fact that flux and torque can be controlled separately using<br />

current control loop <strong>with</strong> PI controllers. However the weak point <strong>of</strong> this drive was DC<br />

motor, which could not worked in aggressive or volatile environment and required<br />

cyclical maintenance. This disadvantages has been eliminated, when instead <strong>of</strong> DC<br />

machine a three phase PMSM motor were used.<br />

In searching new control method for induction machine in 1971 was developed vector<br />

control method known as field oriented control (FOC) [31,38,49]. This method allows<br />

control the flux and the torque in the AC machine in similar way as for DC motor. It<br />

was achieved by transform current vector in stationary reference frame ( α,<br />

β ) into new<br />

coordinate system ( dq) , <strong>with</strong> respect to rotor (magnet) flux vector. So the flux<br />

produced by permanent magnet is frozen to the direct axis <strong>of</strong> the rotor (see Fig. 4.4).<br />

q−axis<br />

β −axis<br />

I s<br />

I sq<br />

δ I<br />

γ I<br />

I sd<br />

γ m<br />

Ω s<br />

Ψ PM<br />

d−axis<br />

α −axis<br />

Figure 4.4 <strong>Vector</strong> diagram illustrated the principle <strong>of</strong> FOC.<br />

Further, stator current vector can be split into two current components: flux current I sd<br />

and torque producing current I sq<br />

. In analogy to separate commutator motor, the flux<br />

current components corresponds to excitation current and torque-producing current<br />

corresponds to the armature current. Therefore, the goal <strong>of</strong> the control system is to<br />

reference the<br />

I<br />

sd_<br />

ref<br />

,<br />

sq_<br />

ref<br />

I stator current components in respect to requirement <strong>of</strong><br />

references torque and flux. The flux and torque producing stator current references are<br />

obtained on the output <strong>of</strong> the reference current generation block (see Fig. 4.5).<br />

54


<strong>Control</strong> methods <strong>of</strong> PM Synchronous motor<br />

Reference Current<br />

Calculation<br />

Reference Current<br />

Calculation<br />

reference<br />

torque<br />

M e _ ref<br />

FG1<br />

FG2<br />

I sd _ ref<br />

reference<br />

d-axis current<br />

I sq _ ref<br />

reference<br />

q-axis current<br />

reference<br />

torque<br />

M e _ ref<br />

FG3<br />

FG3<br />

δ I<br />

I s<br />

γ I<br />

Figure 4.5. Reference current generator block for FOC technique<br />

a) in cartesian form, b) in polar form.<br />

γ m<br />

Function generation FG1 gives the relationship between the torque and the direct axis<br />

stator current component<br />

I<br />

sd_<br />

ref<br />

, and function generator FG2 gives the relationship<br />

between the torque and the quadrature axis stator current<br />

I<br />

sq_<br />

ref<br />

. His graphical<br />

illustration in Fig. 4.6 are presented.<br />

0.5<br />

0<br />

CTA<br />

I [ . ]<br />

sdN<br />

pu<br />

0.5<br />

1<br />

CSF<br />

MTPA<br />

1.5<br />

2<br />

UPF<br />

2.5<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2<br />

M<br />

eN<br />

[ pu . ]<br />

2.5<br />

I [ . ]<br />

sqN<br />

pu<br />

2<br />

1.5<br />

1<br />

CSF<br />

CTA<br />

MTPA<br />

0.5<br />

UPF<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2<br />

M<br />

eN<br />

[ pu . ]<br />

Figure 4.6. Generated current components I<br />

sd<br />

and I sq<br />

dependent on required electromagnetic<br />

torque under difference control strategies. CTA-constant torque angle control, MTPA-maximum<br />

torque per ampere control, CSF-constant stator flux control, UPF – unity power factor control.<br />

55


<strong>Control</strong> methods <strong>of</strong> PM Synchronous motor<br />

For many years the CTA control ( I<br />

_<br />

= 0 ) method has been a popular technique for a<br />

sd<br />

ref<br />

long time because <strong>of</strong> simple control. This method was dedicated for surface permanent<br />

magnet synchronous motor (SPMSM), where the magnetic saliency does not exist<br />

[126]. So the maximum torque per ampere is obtained, when stator current vector is<br />

shifted in respect to rotor flux vector 90 degree. However, in IPMSM, maximum torque<br />

per ampere is obtained <strong>with</strong> torque angle more than 90 degree. This is because <strong>of</strong><br />

existence <strong>of</strong> reluctance torque component due to magnetic saliency (see subchapter<br />

2.2.2). Therefore, the I<br />

sd_<br />

ref<br />

should be negative value [44].<br />

The main question is how or in which manner produce the reference currents in d-q<br />

frame. Its leads to many realization <strong>of</strong> current control structure. Among them generally<br />

we can distinguish two structures <strong>of</strong> current control loop. One <strong>of</strong> them is hysteresis<br />

based control (Fig.4.7a) [3,52] and the second one is PI based current controllers<br />

(Fig.4.7b).<br />

Hysteresis based current control has following disadvantage such as [3]:<br />

• measurement <strong>of</strong> three phase currents are required,<br />

• three independent hysteresis current controllers are required,<br />

• variable switching frequency is achieved,<br />

• fast sampling time is required.<br />

All this listed above disadvantage can be eliminate, when the PI current control are<br />

used. This structure are mostly used in industrial application (Fig. 4.7b).<br />

a)<br />

U DC<br />

M e _ ref<br />

Fig. 4.5<br />

I d _ ref<br />

I q _ ref<br />

d,q/ABC<br />

I A _ ref<br />

I B _ ref<br />

I C _ ref<br />

−<br />

−<br />

S A<br />

S B<br />

S C<br />

Inverter<br />

current<br />

feedback<br />

−<br />

I A current<br />

I<br />

sensors<br />

B<br />

I C<br />

rotor<br />

position<br />

sensor<br />

PMSM<br />

56


<strong>Control</strong> methods <strong>of</strong> PM Synchronous motor<br />

b)<br />

U DC<br />

M e _ ref<br />

Fig. 4.5<br />

I sd _ ref<br />

I sq _ ref<br />

-<br />

e Isd<br />

-<br />

e Isq<br />

PI<br />

PI<br />

U sd<br />

U sq<br />

Reference<br />

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

Calculation<br />

U s _ ref<br />

ϕ Us _ ref<br />

<strong>Space</strong><br />

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

Modulator<br />

S A<br />

S B<br />

S C<br />

Inverter<br />

γ s<br />

I sd<br />

I sq<br />

dq/ABC<br />

current<br />

sensors<br />

I s<br />

γ m<br />

rotor<br />

position<br />

sensor<br />

PMSM<br />

Figure 4.7 <strong>Vector</strong> control structure for PM synchronous motor <strong>with</strong>: a) hysteresis current<br />

control, b) synchronous PI current control<br />

4.3 <strong>Direct</strong> torque control (<strong>DTC</strong>)<br />

The name direct torque control is deliver by the fact that, on the basis <strong>of</strong> the errors<br />

between the reference and the estimated values <strong>of</strong> torque and flux, it is possible to<br />

directly control the inverter states <strong>with</strong>out inner current control loop as for FOC<br />

[32,34,35,50] and [57-66].<br />

The basic idea <strong>of</strong> this control rely on stator voltage vector equation <strong>of</strong> AC motor.<br />

U<br />

dΨ<br />

s<br />

s<br />

= RsIs<br />

+ (4.1)<br />

dt<br />

Making the assumption that ohmic voltage drop on the stator resistor can be neglected<br />

the equation for stator flux vector takes the form:<br />

Ψ =∫ ( U ) dt<br />

(4.2)<br />

s<br />

s<br />

It can be said that the stator voltage vector has directly influence on control stator flux<br />

vector. Using a three phase voltage source inverter to supply the AC motor, there are six<br />

non-zero vectors and two zero voltage vectors. The active vectors change the amplitude<br />

and position <strong>of</strong> stator flux vector, while the zero vectors stop the stator flux vector as<br />

shown in Fig. 4.8.<br />

57


<strong>Control</strong> methods <strong>of</strong> PM Synchronous motor<br />

y<br />

moment <strong>with</strong> active<br />

forward vector<br />

stops <strong>with</strong> zero<br />

vector<br />

Ψ s<br />

moment <strong>with</strong> active<br />

backward vector<br />

γ s<br />

δ Ψ<br />

Ψ =Ψ<br />

r<br />

PM<br />

x<br />

rotates<br />

continuously<br />

α<br />

stator<br />

Figure. 4.7 Stator flux vector Ψ<br />

s<br />

movement relative to rotor flux vector Ψ<br />

r<br />

=ΨPM<br />

under the<br />

influence <strong>of</strong> active and zero inverter voltage vectors.<br />

Therefore, it is possible control the torque angle δ Ψ<br />

across control stator flux vector<br />

position in respect to rotor flux vector produced by permanent magnet Ψ<br />

r<br />

=Ψ<br />

PM<br />

Ψ<br />

s<br />

, what<br />

further allows to have impact on control the electromagnetic torque in accordance <strong>with</strong><br />

following formula:<br />

M<br />

e<br />

3 Ψ sin<br />

2 s<br />

ΨPM<br />

δΨ Ψs<br />

( Lq<br />

−Ld)sin2δ<br />

Ψ<br />

= pb[ − ]<br />

(4.3)<br />

2 L<br />

2L L<br />

d d q<br />

Generally the <strong>DTC</strong> technique operate at constant stator flux amplitude<br />

Ψ<br />

s<br />

, what<br />

correspond to CSF operation, because <strong>of</strong> simple reference stator flux amplitude equal<br />

nominal value <strong>of</strong> permanent magnet. For <strong>DTC</strong> technique can be also apply all control<br />

strategies discussed in Chapter 2.<br />

Using the block generator for reference stator flux amplitude and electromagnetic<br />

torque as is shown in Fig 4.8. it is possible to draw the relationship between required<br />

Reference<br />

Flux<br />

Calculation<br />

reference<br />

flux<br />

Ψ s _ ref<br />

M<br />

e_<br />

ref<br />

reference<br />

torque<br />

reference<br />

torque<br />

M e _ ref<br />

Figure 4.8. Reference flux and torque generator block for <strong>DTC</strong> technique.<br />

reference flux and torque. His graphical illustration in Fig. 4.9 is presented.<br />

58


<strong>Control</strong> methods <strong>of</strong> PM Synchronous motor<br />

Ψ<br />

sN<br />

[ pu]<br />

2<br />

1.5<br />

1<br />

CTA<br />

MTPA<br />

CSF<br />

0.5<br />

UPF<br />

0<br />

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2<br />

M<br />

eN<br />

[ pu]<br />

Figure 4.9. Generated stator flux amplitude Ψ s _ ref dependent on required electromagnetic<br />

torque under difference control strategies. CTA-constant torque angle control, MTPA-maximum<br />

torque per ampere control, CSF-constant stator flux control, UPF – unity power factor control.<br />

The basic structure <strong>of</strong> direct flux and torque control voltage-sourced PWM inverter-fed<br />

permanent magnet synchronous motor is shown in Fig.4.10.[3,36,37,41,45]<br />

Flux<br />

hysteresis<br />

U DC<br />

Ψ s _ ref<br />

M e _ ref<br />

−<br />

−<br />

H Ψ s<br />

H m<br />

d Ψ s<br />

d M e<br />

Switching<br />

Table<br />

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

sector γ<br />

histeresis s ( N )<br />

S A<br />

S B<br />

S C<br />

Inverter<br />

Ψ s<br />

M e<br />

Flux and<br />

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

Estimation<br />

I s<br />

γ m<br />

PMSM<br />

Figure. 4.10. Block diagram <strong>of</strong> switching table based direct torque control ST-<strong>DTC</strong>.<br />

The command stator flux amplitude<br />

Ψ<br />

s _ ref<br />

and electromagnetic torque<br />

e_<br />

ref<br />

M values<br />

are compared <strong>with</strong> the actual<br />

Ψ s<br />

and M e<br />

values, in hysteresis flux and torque<br />

controllers, respectively. The flux and the torque controllers are a two-level<br />

comparators.<br />

The digitized outputs signals <strong>of</strong> the flux controllers are defined as:<br />

d = 1 (increase flux) for Ψ > Ψ<br />

_<br />

+ (4.4a)<br />

Ψ s<br />

s s ref<br />

H Ψ<br />

d = 0 (decrease flux) for Ψ < Ψ<br />

_<br />

− (4.4b)<br />

Ψ s<br />

s s ref<br />

H Ψ<br />

59


<strong>Control</strong> methods <strong>of</strong> PM Synchronous motor<br />

and those <strong>of</strong> the torque controller as<br />

where<br />

d = 1(increase torque) for M > M<br />

_<br />

+ H<br />

(4.5a)<br />

M e<br />

e e ref m<br />

d = 0 (decrease torque) for M < M<br />

_<br />

− H<br />

(4.5b)<br />

M e<br />

e e ref m<br />

H<br />

m<br />

and H Ψ<br />

are hysteresis bands for torque and flux, respectively.<br />

The digitized variables d Ψ<br />

, d and the stator flux position sector γ M<br />

s( N )<br />

information<br />

s<br />

e<br />

create a digital word, which select appropriate voltage vector from the switching table.<br />

Next, from the selection table the proper voltage vectors are selected and the gate pulses<br />

SA, SB,<br />

S<br />

C<br />

to control the power switches in the inverter are generated.<br />

The circular stator flux vector trajectory can be divided into six symmetrical sectors<br />

(according to the non zero voltage vectors), which are defined as (see Fig. 4.11):<br />

sector 1 − 30°≤ γ s<br />

< 30°<br />

V 3 (010)<br />

V 2 (110)<br />

sector 2 30°< γ s<br />

≤ 90°<br />

sector 3 90°< γ s<br />

≤ 150°<br />

sector 4 150°< γ s<br />

≤ 210°<br />

sector 5 210°< γ s<br />

≤ 270°<br />

sector 6 270°< γ s<br />


<strong>Control</strong> methods <strong>of</strong> PM Synchronous motor<br />

β<br />

Ψ<br />

s<br />

↓<br />

M<br />

e<br />

↑<br />

Ψ<br />

s<br />

↑<br />

M<br />

e<br />

↑<br />

Ψ<br />

s<br />

↓<br />

Ψs<br />

↑<br />

Ψ s<br />

γ<br />

s γ<br />

s<br />

α<br />

Ψ<br />

s<br />

↓<br />

M<br />

e<br />

↓<br />

Ψ<br />

s<br />

↑<br />

M<br />

e<br />

↓<br />

Figure. 4.12 Voltage vector effects in sector 1 on stator flux and torque.<br />

The presented rule for first sector can be extended for other sectors, what further help<br />

construct the switching Tables 4.1 and 4.2 as below [56]<br />

Flux <strong>Torque</strong> sector 1 sector 2 sector 3 sector 4 sector 5 sector 6<br />

d ψs =1<br />

d me. =1 V<br />

2<br />

V<br />

3<br />

V<br />

4<br />

V<br />

5<br />

V<br />

6<br />

V<br />

1<br />

d me =0 V<br />

6<br />

V<br />

1<br />

V<br />

2<br />

V<br />

3<br />

V<br />

4<br />

V<br />

5<br />

d ψs =0<br />

d me. =1 V<br />

3<br />

V<br />

4<br />

V<br />

5<br />

V<br />

6<br />

V<br />

1<br />

V<br />

2<br />

d me =0 V<br />

5<br />

V<br />

6<br />

V<br />

1<br />

V<br />

2<br />

V<br />

3<br />

V<br />

4<br />

Table 4.1. Switching table for <strong>DTC</strong> <strong>with</strong> active vectors.<br />

Flux <strong>Torque</strong> sector 1 sector 2 sector 3 sector 4 sector 5 sector 6<br />

d ψs =1<br />

d me. =1 V<br />

2<br />

V<br />

3<br />

V<br />

4<br />

V<br />

5<br />

V<br />

6<br />

V<br />

1<br />

d me =0 V<br />

7<br />

V<br />

0<br />

V<br />

7<br />

V<br />

0<br />

V<br />

7<br />

V<br />

0<br />

d ψs =0<br />

d me. =1 V<br />

3<br />

V<br />

4<br />

V<br />

5<br />

V<br />

6<br />

V<br />

1<br />

V<br />

2<br />

d me =0 V<br />

0<br />

V<br />

7<br />

V<br />

0<br />

V<br />

7<br />

V<br />

0<br />

V<br />

7<br />

Table 4.2. Switching table for <strong>DTC</strong> <strong>with</strong> zero and active vectors.<br />

Tables 4.1 and 4.2 represent the eight and six voltage-vectors switching tables.<br />

61


<strong>Control</strong> methods <strong>of</strong> PM Synchronous motor<br />

The <strong>DTC</strong> has lesser parameter dependence and fast torque when compare <strong>with</strong> the<br />

torque control via PWM current control.<br />

Among the well-know advantages <strong>of</strong> the <strong>DTC</strong> scheme are the following:<br />

• Simple control,<br />

• Excellent torque dynamics,<br />

• Absence <strong>of</strong> coordinate transformations,<br />

• Absence <strong>of</strong> separate voltage modulation block,<br />

• Absence <strong>of</strong> voltage decoupling circuits,<br />

• There are no current control loops, hence, the current is not regulated directly,<br />

• Stator flux vector and torque estimation is required.<br />

Among the well-know disadvantages <strong>of</strong> the <strong>DTC</strong> scheme are the following:<br />

• variable switching frequency (difficulties <strong>of</strong> LC input EMI filter design),<br />

• high sampling time is required (fast microprocessor and A/D converter ),<br />

• inverter switching frequency depending on: flux and torque hysteresis bands,<br />

machine parameters, sampling frequency,<br />

• violence <strong>of</strong> polarity consistency rules (huge voltage stress for IGBT transistor),<br />

• current and torque distortion caused by sector changes,<br />

• start and low speed operation problems,<br />

• high sampling frequency needed for digital implementation <strong>of</strong> hysteresis<br />

comparators,<br />

• high noisy level,<br />

• high current and torque ripple.<br />

Many modifications <strong>of</strong> the basic switching table based direct torque control (ST-<strong>DTC</strong>)<br />

scheme at improving starting, overload condition, very low speed operation, torque<br />

ripple reduction, variable switching frequency functioning, and noise level attenuation<br />

have been proposed during last decade.<br />

In the last five years, many researcher have been carried out to try solve the above<br />

mentioned problems <strong>of</strong> ST-<strong>DTC</strong> scheme. Therefore, the following solutions have been<br />

developed in order to eliminated mentioned before problems:<br />

62


<strong>Control</strong> methods <strong>of</strong> PM Synchronous motor<br />

• Use <strong>of</strong> improved switching table,<br />

• Use <strong>of</strong> comparators <strong>with</strong> and <strong>with</strong>out hysteresis, at two or three levels,<br />

• Use <strong>of</strong> multi-level inverter,<br />

• Introduction <strong>of</strong> fuzzy or neuro-fuzzy techniques,<br />

• Use <strong>of</strong> sophisticated flux estimators to improve the low speed behavior,<br />

• Implementation <strong>of</strong> <strong>DTC</strong> schemes <strong>with</strong> constant switching frequency operation<br />

In multi-level inverter there will be more voltage vectors available to control the flux<br />

and torque. Therefore, a smoother torque can be expected. However, more power<br />

switches are needed to achieved a lower ripple, which will increase the system cost and<br />

complexity.<br />

All this contributions allow the <strong>DTC</strong> performance to be improved, but at the same time<br />

they lead to more complex schemes. As expected, conventional <strong>DTC</strong> is growing in<br />

field-oriented control area and the so-called improved <strong>DTC</strong> <strong>with</strong> space vector<br />

modulation (<strong>SVM</strong>). Let us call it <strong>DTC</strong>-<strong>SVM</strong>. This control concept will be deeply<br />

discussed in the next Chapter.<br />

63


<strong>Control</strong> methods <strong>of</strong> PM Synchronous motor<br />

4.4 Summary<br />

‣ In FOC drive flux linkage and electromagnetic torque are controlled indirectly<br />

and independently by PI controllers <strong>with</strong> space vector modulator (<strong>SVM</strong>). In this<br />

control concept current control loop is required,<br />

‣ In <strong>DTC</strong> drive, flux linkage and electromagnetic torque are controlled directly<br />

and independently by hysteresis controllers and selection <strong>of</strong> optimum inverter<br />

switching modes. In this control concept flux and torque control loop is required,<br />

‣ In <strong>DTC</strong> all switch changes <strong>of</strong> the inverter are based on the electromagnetic state<br />

<strong>of</strong> the motor.<br />

‣ The <strong>DTC</strong> technique is different from the traditional methods <strong>of</strong> controlling<br />

torque, where the current controllers in the rotor reference frame are used. It is<br />

completely different control concept (approach) from FOC. The new control<br />

technology was characterized by simplicity, good performance and robustness,<br />

because <strong>of</strong> bang-bang hysteresis control. Using <strong>DTC</strong> it is possible to obtain a<br />

good dynamic control <strong>of</strong> the torque <strong>with</strong>out current controllers and any<br />

mechanical transducers on the machine shaft. Moreover, in this control structure<br />

the PWM modulator is not required. Its is occupied by variable switching<br />

frequency.<br />

‣ The flux weakening control becomes easier because stator flux linkage can be<br />

controlled directly in the <strong>DTC</strong> system <strong>of</strong> PMSM.<br />

64


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

Chapter 5<br />

DIRECT TORQUE CONTROL WITH SPACE VECTOR MODULATION<br />

(<strong>DTC</strong>-<strong>SVM</strong>)<br />

5.1 Introduction<br />

The <strong>DTC</strong>-<strong>SVM</strong> greatly improves torque and flux performance by:<br />

• Achieved fixed switching frequency,<br />

• Reduced torque and flux ripples.<br />

The main idea <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> is based on analyze <strong>of</strong> the torque equation<br />

M<br />

Assuming that the Ld = Lq = Ls<br />

e<br />

3 Ψ sin<br />

2 s<br />

ΨPM<br />

δΨ Ψs<br />

( Lq<br />

−Ld)sin2δ<br />

Ψ<br />

= pb[ − ]<br />

(5.1)<br />

2 L<br />

2L L<br />

d d q<br />

M<br />

e<br />

3 Ψs<br />

ΨPM<br />

sinδ = pb[ Ψ<br />

]<br />

(5.2)<br />

2 L<br />

From equation (5.2) we can see that for constant stator flux amplitude<br />

by permanent magnet<br />

Ψ<br />

PM<br />

s<br />

Ψ<br />

s<br />

and flux produced<br />

, the electromagnetic torque can be changed by control <strong>of</strong> the<br />

torque angle δ Ψ<br />

. This is the angle between the stator and rotor flux linkage, when the stator<br />

resistance is neglected. The torque angle, in turn, can be changed by changing position <strong>of</strong> the<br />

stator flux vector θ Ψ s<br />

in respect to PM vector using the actual voltage vector supplied by<br />

PWM inverter.<br />

In the steady state, δ Ψ<br />

is constant and corresponds to a load torque, whereas stator and rotor<br />

flux rotate at synchronous speed. In transient operation, δ Ψ<br />

varies and the stator and rotor<br />

flux rotate at different speeds (Fig. 5.1).<br />

65


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

q − axis<br />

β<br />

δ Ψ<br />

Ψ s _ ref<br />

∆δ Ψ<br />

θ Ψ<br />

s<br />

Ψ s<br />

Ψ PM<br />

d − axis<br />

θ r<br />

Figure 5.1. <strong>Space</strong> vector diagram illustrating torque control conditions.<br />

α<br />

5.2 Cascade structure <strong>of</strong> <strong>DTC</strong>–<strong>SVM</strong> scheme<br />

The structure <strong>of</strong> proposed control scheme is shown in the Fig. 5.2. [11,33,42,48,51,53,54]<br />

U DC<br />

Ψ s _ ref<br />

U<br />

s α _ ref<br />

S A<br />

S B<br />

M e _ ref<br />

e M<br />

∆δ Ψ<br />

U<br />

s β _ ref<br />

S C<br />

θ Ψs Ψs<br />

I s<br />

M e<br />

I s<br />

γ m<br />

Figure 5.2. Cascade structure <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> scheme.<br />

The error between reference and measured torque can be expressed as:<br />

Ψ Ψ sin( δ +∆δ ) Ψ Ψ sinδ<br />

eM = M − M = p<br />

− (5.3)<br />

3 s_<br />

ref PM Ψ Ψ<br />

_<br />

[<br />

s PM Ψ<br />

e ref e b<br />

]<br />

2<br />

Ls<br />

Ls<br />

From equation (5.3) we can see that the relation between torque error and increment <strong>of</strong> load<br />

angel<br />

∆δ Ψ<br />

is nonlinear. Therefore, we used PI controller which generates the load angel<br />

66


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

increment required to minimize the instantaneous error between reference<br />

M and actual<br />

e_<br />

ref<br />

M<br />

e<br />

torque.<br />

In control scheme <strong>of</strong> Fig. 5.2 the torque error signal e M<br />

is delivered to the PI controller,<br />

which determines the increment <strong>of</strong> torque angle<br />

∆ δ Ψ<br />

. Based on this signal and reference<br />

amplitude <strong>of</strong> stator flux<br />

Ψ , the reference voltage vector in stator coordinates α,<br />

β is<br />

s _ ref<br />

calculated. The calculation block <strong>of</strong> reference voltage vector also uses information about the<br />

actual stator flux vector (amplitude<br />

Ψ<br />

s<br />

and position θ Ψ s<br />

) as well as measured current vector<br />

I<br />

s<br />

. The reference stator voltage vector is delivered to space vector modulator (<strong>SVM</strong>), which<br />

generates the switching signals<br />

S , S , S for power transistors <strong>of</strong> inverter.<br />

A<br />

B<br />

C<br />

The calculation block <strong>of</strong> reference voltage vector is shown in Fig. 5.3.<br />

Ψ s _ ref<br />

RsI s α<br />

Ψ<br />

s α _ ref<br />

∆Ψ sα<br />

U α<br />

_<br />

s<br />

_ ref<br />

∆δ Ψ<br />

−<br />

Ψ<br />

s β _ ref<br />

∆Ψ sβ<br />

−<br />

U<br />

s β ref<br />

θ Ψs<br />

Ψ sα<br />

Ψ sβ<br />

Rs<br />

Is<br />

β<br />

Figure 5.3. Calculation block <strong>of</strong> reference voltage vector.<br />

Based on<br />

∆ signal, reference <strong>of</strong> stator flux amplitude Ψ<br />

s _ ref<br />

and measured stator flux<br />

δ Ψ<br />

vector position<br />

θ Ψ s<br />

(Fig. 5.3.), the reference flux components<br />

Ψ Ψ in stator<br />

s α _ ref<br />

,<br />

sβ<br />

_ ref<br />

coordinate system are calculated as:<br />

Ψ = Ψ cos( θ +∆δ<br />

)<br />

sα<br />

_ ref s_<br />

ref Ψs<br />

Ψ = Ψ sin( θ +∆δ<br />

)<br />

sβ<br />

_ ref s _ ref Ψs<br />

Ψ<br />

Ψ<br />

(5.4)<br />

Pleas note that for constant flux operation region the reference value <strong>of</strong> stator flux amplitude<br />

Ψ is equal flux amplitude <strong>of</strong> permanent magnet Ψ<br />

PM<br />

.<br />

s _ ref<br />

67


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

The references <strong>of</strong> stator flux components (see Fig. 5.3) are compared <strong>with</strong> estimated value:<br />

Ψ<br />

Ψ<br />

= Ψ<br />

cos θ ,<br />

sα<br />

s Ψs<br />

= Ψ<br />

sin θ ,<br />

sβ<br />

s Ψs<br />

The command voltage can be calculated from flux errors in<br />

(5.5)<br />

α, β coordinate system as<br />

follows:<br />

U<br />

U<br />

∆Ψ<br />

= + R I<br />

sα<br />

sα<br />

_ ref s sα<br />

Ts<br />

∆Ψ<br />

= + R I<br />

sβ<br />

sβ<br />

_ ref s sβ<br />

Ts<br />

(5.6)<br />

Where: T<br />

s<br />

is sampling time,<br />

∆Ψ = Ψ − Ψ , ∆Ψ =Ψ<br />

_<br />

−Ψ .<br />

sα sα_<br />

ref sα<br />

sβ sβ ref sβ<br />

The presented bellow design methodology for flux and torque control loops based on the<br />

approach presented in literature [11,43].<br />

5.2.1 Digital flux control loop<br />

The flux control loop is based on the voltage equations <strong>of</strong> PMSM machine in stator<br />

coordinates.<br />

U<br />

U<br />

dΨ<br />

dt<br />

sα<br />

sα<br />

= RsIsα<br />

+ (5.7a)<br />

dΨ<br />

sβ<br />

sβ<br />

= RsIsβ<br />

+ (5.7b)<br />

dt<br />

Using Laplace transformation the above equations can be written as:<br />

U = sΨ + R I<br />

(5.8a)<br />

sα sα s sα<br />

U = sΨ + R I<br />

(5.8b)<br />

sβ sβ s sβ<br />

It corresponds to flux model <strong>of</strong> PMSM machine in α,<br />

β system presented in Fig. 5.4.<br />

68


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

R I α<br />

U<br />

s α<br />

U<br />

s β<br />

s<br />

−<br />

s<br />

1<br />

s<br />

1<br />

s<br />

−<br />

R I β<br />

s<br />

s<br />

Ψ sα<br />

Ψ sβ<br />

Figure 5.4. Flux model <strong>of</strong> PMSM in stator coordinates.<br />

In order to control the flux components in α,<br />

β frame the bellow control structure can be<br />

applied.<br />

Ψ<br />

s α _ ref<br />

Ψ<br />

s β _ ref<br />

Flux control<br />

part<br />

−<br />

∆Ψ sα<br />

−<br />

∆Ψ sβ<br />

P block<br />

P=<br />

1/ Ts<br />

P block<br />

P=<br />

1/ Ts<br />

∆Ψ<br />

s α<br />

T s<br />

∆Ψ<br />

s β<br />

T s<br />

RI<br />

s s α<br />

U<br />

s α _ ref<br />

U<br />

s β _ ref<br />

RI<br />

s s β<br />

RI<br />

s s α<br />

−<br />

1<br />

s<br />

1<br />

s<br />

−<br />

RI β<br />

s<br />

s<br />

Flux PMSM<br />

model<br />

Ψ<br />

sα<br />

Ψ sβ<br />

Ψ sα<br />

Ψ sβ<br />

Figure 5.5. Flux control loop <strong>with</strong> two P controller in α,<br />

β reference frame.<br />

Pleas note that regarding to Fig. 5.1 the following rules are keeping:<br />

Ψ<br />

sα _ ref<br />

= Ψ<br />

s _ ref<br />

cos( θΨs<br />

+∆ δ )<br />

(5.9a)<br />

Ψ<br />

sβ _ ref<br />

= Ψ<br />

s _ ref<br />

sin( θΨs<br />

+∆ δ )<br />

(5.9b)<br />

and<br />

sα<br />

s_<br />

refcosθ Ψ s<br />

Ψ = Ψ (5.10a)<br />

sβ<br />

s_<br />

refsinθ Ψ s<br />

Ψ = Ψ (5.10b)<br />

In order to find the formula for tuning the P controllers in the flux loop, the following<br />

assumptions should be made:<br />

69


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

• increment <strong>of</strong> torque angle ∆δ Ψ<br />

coming from torque control loop (see Fig.5.2.) is equal<br />

zero. It means that the torque is not produced,<br />

• stator flux vector position θ Ψ s<br />

and rotor flux vector position θ r<br />

are equal zero. It<br />

corresponds to situation, where the those two flux vectors lie along the α axis.<br />

In this special case the reference stator flux amplitude Ψ<br />

_<br />

=Ψ<br />

α _<br />

can be controlled<br />

s ref s ref<br />

trough the reference stator voltage component U<br />

α _<br />

= U<br />

_<br />

, when the voltage drop on the<br />

s ref s ref<br />

stator resistances in α,<br />

β axes are neglected (see Fig. 5.5). Therefore, the simplified flux<br />

control loop can be shown in Fig. 5.6.<br />

Ψ =Ψ α<br />

s _ ref s _ ref<br />

s sα<br />

Ψ =Ψ<br />

−<br />

∆Ψ sα<br />

controller<br />

P<br />

U<br />

sα _ ref<br />

= Us _ ref<br />

U β<br />

=<br />

s _ ref<br />

0<br />

PMSM<br />

Ψ sα<br />

Ψ sβ<br />

Continuous s-domain<br />

Figure 5.6. Simplified flux control loop in α,<br />

β coordinates.<br />

Simplified flux control loop in s domain is shown in Fig. 5.7, where CΨ ( s)<br />

is a transfer<br />

function <strong>of</strong> the P controller given by:<br />

CΨ() s = KpΨ<br />

(5.11)<br />

The transfer function between stator flux amplitude Ψ<br />

s<br />

=Ψ sα and stator voltage amplitude<br />

U<br />

s<br />

can be expressed as:<br />

G<br />

Ψ<br />

Ψ<br />

s 1<br />

() s = = (5.12)<br />

U s<br />

s<br />

70


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

<strong>Control</strong> Plant<br />

P controller<br />

Ψ s_ref<br />

C<br />

Ψ<br />

() s<br />

U s<br />

G<br />

Ψ<br />

() s<br />

Ψ s<br />

Figure 5.7. Block diagram <strong>of</strong> flux controller in s domain.<br />

Hence the transfer function <strong>of</strong> the closed stator flux amplitude control loop is obtained as:<br />

G<br />

Ψ _ closed<br />

Ψ<br />

s<br />

() s<br />

_ ref CΨ() s GΨ()<br />

s<br />

() s = =<br />

Ψ () s 1 + C () s G () s<br />

s<br />

Ψ<br />

Ψ<br />

(5.12)<br />

Substituting transfer function for CΨ ( s)<br />

and GΨ ( s)<br />

one becomes:<br />

⎛1⎞ ⎛1⎞<br />

KpΨ⎜ ⎟ KpΨ⎜ ⎟<br />

s s K<br />

p<br />

GΨ _ closed<br />

() s =<br />

⎝ ⎠<br />

=<br />

⎝ ⎠<br />

=<br />

⎛1<br />

⎞ s+ KpΨ<br />

s+<br />

K<br />

1+ K<br />

pΨ<br />

⎜ ⎟<br />

⎝s<br />

⎠ s<br />

Ψ<br />

pΨ<br />

(5.13)<br />

Discrete design<br />

The transfer function for P controller in discrete system is expressed as:<br />

CΨ( z) = KpΨ<br />

(5.14)<br />

Ψ<br />

s_ ref( z)<br />

C<br />

Ψ<br />

( z)<br />

U<br />

= sα U<br />

s<br />

Dz ( )<br />

z −1<br />

GΨ<br />

( z)<br />

}<br />

ZOH<br />

1<br />

s<br />

Ψ<br />

s<br />

( z)<br />

Figure 5.8. Block diagram <strong>of</strong> flux controller in discrete domain.<br />

Where, CΨ ( z)<br />

is discrete transfer function for P controller, Dz ( )<br />

delay for voltage generation from PWM .<br />

1<br />

z −<br />

is for one sampling time<br />

71


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

The GΨ ( z)<br />

is discrete transfer function for voltage-flux relationship <strong>with</strong> zero hold order<br />

(ZOH) can be calculated as:<br />

−1<br />

GΨ<br />

() s z−1 1<br />

GΨ<br />

( z) = (1 − z ) Z[ ] = Z[ ]<br />

(5.15)<br />

2<br />

s z s<br />

Using table <strong>of</strong> Z transformation [2]. Finally, it gives<br />

G<br />

Ψ<br />

( z −1)<br />

zTs<br />

AΨ<br />

d<br />

2<br />

z ( z −1) ( z−1)<br />

( z)<br />

= =<br />

(5.16)<br />

Where A = Ψ d<br />

T and s<br />

T<br />

s<br />

is sampling time <strong>of</strong> the discrete system.<br />

Hence, the closed loop transfer function between Ψ<br />

s<br />

( z)<br />

and Ψ ( z)<br />

is obtained as:<br />

_ ref<br />

s<br />

G<br />

Ψ _ closed<br />

K<br />

pΨ<br />

Ψ<br />

s<br />

( z) _ ref CΨ( z) GΨ( z) D( z)<br />

( z)<br />

= =<br />

Ψ ( z) 1 + C ( z) G ( z) D( z)<br />

A<br />

Ψd<br />

( −1)<br />

z z<br />

K A<br />

= =<br />

K A z z K A<br />

1+<br />

z z<br />

pΨ<br />

Ψd<br />

2<br />

pΨ Ψd − +<br />

pΨ Ψd<br />

( −1)<br />

s<br />

Ψ<br />

Ψ<br />

(5.16)<br />

The flux step response depended on poles placement <strong>of</strong> closed flux control loop. The pole<br />

placement can be selected by setting the<br />

K<br />

p Ψ<br />

.<br />

Assuming, that CΨd = KpΨAΨd<br />

the GΨ _ closed<br />

( z)<br />

expressed by equations (5.16) will take the<br />

following form:<br />

G<br />

CΨd<br />

( z) =<br />

z − z+<br />

C<br />

Ψ _ closed<br />

2<br />

Ψd<br />

(5.17)<br />

The nomogram <strong>of</strong> Fig. 5.9 shows the relationship between overshoot M<br />

p[%]<br />

, rise time t r<br />

and<br />

settling time t s<br />

in respect to<br />

C Ψ d<br />

.<br />

Please not that t r<br />

is time calculate from 10% to 90% <strong>of</strong> output signals and t s<br />

is the time it<br />

takes the system transient to decay +-1%.<br />

72


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

Figure 5.9. The relationship between overshoot, rise time and settling time versus to<br />

flux amplitude control loop.<br />

C Ψ d<br />

for stator<br />

Now from few values <strong>of</strong> C Ψ d<br />

=[0.4046, 0.2688, 0.1720] we can choose C Ψ d<br />

=0.2688, which<br />

guaranties overshoot about 0% and settling time about 10 times <strong>of</strong> sampling time.<br />

1<br />

2<br />

3<br />

1<br />

Figure 5.10. Step flux response for different to<br />

C Ψ d<br />

=0.2688, black line (3)<br />

d<br />

C Ψ d<br />

: red line(1) C Ψ d<br />

=0.4046, blue line (2)<br />

C Ψ<br />

=0.1720.<br />

It corresponds to the transfer function <strong>of</strong> closed stator flux control loop as:<br />

G<br />

C<br />

0.2688<br />

( z)<br />

= =<br />

z − z+ C z −z+ 0.2688<br />

Ψd<br />

Ψ _ closed<br />

2 2<br />

Ψd<br />

(5.18)<br />

Using digitalized motor parameters A = Ψ d<br />

T and chosen s<br />

C Ψ d<br />

value we can calculate the<br />

parameters <strong>of</strong> P digital flux controller as:<br />

73


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

K<br />

pΨ<br />

C<br />

A<br />

Ψd<br />

Ψd<br />

= = (5.19)<br />

Ψd<br />

C<br />

T<br />

s<br />

For example let us assume that sampling time in digital flux control loop is equal T = 200µ<br />

s.<br />

The gain <strong>of</strong> P controller is:<br />

s<br />

K<br />

pΨ<br />

0.2688 0.2688<br />

= = = 1344<br />

(5.20)<br />

A 0.0002<br />

Ψd<br />

In digital control when the sampling time changes the parameters <strong>of</strong> digitalized plant control<br />

A Ψ d<br />

will also change. Therefore, to keep closed loop transfer function as close as possible to<br />

G<br />

C<br />

0.2688<br />

( z)<br />

= =<br />

, the gain <strong>of</strong> P flux controller should also be<br />

z − z+ C z −z+ 0.2688<br />

Ψd<br />

Ψ _ closed<br />

2 2<br />

Ψd<br />

changed (see Table 5.1.).<br />

Keeping constant transfer function GΨ _ closed<br />

( z)<br />

the flux step response for different sampling<br />

times<br />

T<br />

s<br />

= 50µ s , 100µ s , 200µ s , 400µ s , which correspond to switching frequency f s<br />

=<br />

20kHz, 10kHz, 5kHz, 2.5kHz are presented in Fig. 5.11.<br />

Figure 5.11. Flux tracking performance for different sampling times T<br />

s<br />

= 50µ s (blue line -1), 100µ<br />

s<br />

(green line -2), 200µ s (red line -3), 400µ s (light blue line -4).<br />

74


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

From Fig. 5.11 it can observed that overshoot is around 0% and the settling time took 10<br />

times <strong>of</strong> microprocessor sampling time. So, it is possible control the flux amplitude in 10<br />

samples<br />

The settings <strong>of</strong> P flux controller for different sampling time T<br />

s<br />

= 50 µ s,100 µ s,<br />

200 µ s,400µ<br />

s<br />

are summarized in Table 5.1.<br />

75


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

76


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

The behavior <strong>of</strong> the flux control loop was tested using SABER simulation package. The<br />

model created in SABER takes into account the whole control system, which include<br />

real models <strong>of</strong> inverter and permanent magnet synchronous motor.<br />

The flux step response is shown in Fig. 5.12., when parameters <strong>of</strong> P flux controller<br />

designed for sampling time T s<br />

= 200µ s were used for control plant for different<br />

sampling times T s<br />

= 50µ s , 100µ s , 200µ s , 400µ s , which correspond to switching<br />

frequency f<br />

s<br />

= 20kHz, 10kHz, 5kHz, 2.5kHz.<br />

Figure 5.12. Flux tracking performance for different sampling time T<br />

s<br />

= 50µ s , 100µ s ,<br />

200µ s , 400µ s ( f<br />

s<br />

= 20kHz, 10kHz, 5kHz,2.5kHz) using gain <strong>of</strong> P controller designed<br />

for T<br />

s<br />

= 200µ s ( f<br />

s<br />

= 5kHz).<br />

After modification <strong>of</strong> P flux controller gain according to Table 5.1 it is possible to<br />

achieve better results as shown in Fig. 5.13, what confirms proper flux tracking<br />

performance in steady and dynamics state.<br />

77


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

Figure 5.13. Flux tracking performance for different sampling time T s<br />

= 50µ s , 100µ s ,<br />

200µ s , 400µ s ( f<br />

s<br />

= 20kHz, 10kHz, 5kHz,2.5kHz) using designed gain <strong>of</strong> P controller<br />

calculated individually (see table 5.1.)<br />

The simulation results in SABER package for Ts=200us is presented in Fig. 5.14.<br />

Figure 5.14. Simulated (SABER) flux tracking performance for step change from 70% - 100%<br />

<strong>of</strong> nominal flux.<br />

As we can observed from Fig. 5.14 that overshoot is around M<br />

p<br />

= 0% and settling time<br />

took about 10 sampling time <strong>of</strong> microprocessor, what proved the design procedure <strong>of</strong> P<br />

digital flux controller gain.<br />

78


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

5.2.2 Digital torque control loop<br />

The considered torque control loop is shown in Fig. 5.15.<br />

Ψ s _ ref<br />

M e _ ref<br />

−<br />

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

controller<br />

PI<br />

∆δ Ψ<br />

Reference<br />

flux generator<br />

In stator frame<br />

Ψ<br />

s α _ ref<br />

Ψ<br />

s β _ ref<br />

−<br />

∆Ψ sα<br />

−<br />

∆Ψ sβ<br />

Flux<br />

controllers<br />

In stator frame<br />

U<br />

s α _ ref<br />

U<br />

s β _ ref<br />

M e<br />

θ Ψs<br />

Ψ sα<br />

Ψ sβ<br />

Figure 5.15. <strong>Torque</strong> control loop <strong>with</strong> PI controller.<br />

Based on the equation (5.9a-b and 5.10a-b) the stator flux errors in α,<br />

β coordinates can<br />

be calculated as:<br />

∆Ψ<br />

α<br />

= Ψ cos( θ + ∆δ ) − Ψ cos θ = Ψ [cos( θ + ∆δ ) − cos θ ]<br />

s s_ ref Ψs Ψ s_ ref Ψs s_<br />

ref Ψs Ψ Ψs<br />

(5.21a)<br />

∆Ψ<br />

β<br />

= Ψ sin( θ +∆δ ) − Ψ sin θ = Ψ [sin( θ +∆δ ) − sin θ ]<br />

s s_ ref Ψs Ψ s_ ref Ψs s_<br />

ref Ψs Ψ Ψs<br />

(5.21b)<br />

Assuming that for small changes <strong>of</strong> ∆δ Ψ<br />

the cos ∆δ Ψ<br />

≅ 1and sin ∆δ Ψ<br />

≅∆ δ<br />

Ψ<br />

, the equations<br />

(5.21a) and (5.21b) are given by:<br />

∆Ψ<br />

α _<br />

= − Ψ<br />

_<br />

∆ δ sinθ<br />

(5.22a)<br />

s ref s ref Ψ Ψs<br />

∆Ψ<br />

β _<br />

= Ψ<br />

_<br />

∆ δ cosθ<br />

(5.22b)<br />

s ref s ref Ψ Ψs<br />

In order to design the PI torque controller the following assumption are made:<br />

• stator flux vector position θ Ψ s<br />

and rotor flux vector position θ<br />

r<br />

are equal zero. It<br />

correspond to situation, when those two flux vectors lie along theα axis,<br />

• the reference stator flux amplitude is equal value <strong>of</strong> permanent magnet flux<br />

Ψ<br />

s _ ref<br />

=Ψ<br />

PM<br />

,<br />

• stator resistance is neglected.<br />

Therefore, the error stator fluxes in α,<br />

β coordinates are calculated as:<br />

∆Ψ = , (5.23a)<br />

sα<br />

_ ref<br />

0<br />

79


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

∆Ψ = Ψ ∆ . (5.23b)<br />

sβ<br />

_ ref PM<br />

δ Ψ<br />

Stator voltage components using equations (2.23a-b) can be expressed as:<br />

U α<br />

= (5.24a)<br />

s _ ref<br />

0<br />

U<br />

= ∆Ψ K<br />

(5.24b)<br />

sβ _ ref sβ _ ref pΨ<br />

And further because <strong>of</strong> U<br />

β _<br />

= U<br />

_<br />

s ref s ref<br />

U<br />

=Ψ ∆ δ K<br />

(5.25)<br />

s PM pΨ<br />

So, the transfer function between stator voltage amplitude<br />

angle ∆ δ Ψ<br />

can be written as:<br />

U<br />

s<br />

G<br />

δ<br />

() s = = K<br />

ΨΨ<br />

∆δ<br />

M p PM<br />

Ψ<br />

U<br />

s<br />

and increment <strong>of</strong> torque<br />

(5.26)<br />

Where<br />

K<br />

p Ψ<br />

is the gain <strong>of</strong> stator flux P controller.<br />

For example for sampling time T = 200µ<br />

s, calculated<br />

s<br />

0.2688<br />

K = pΨ 1344<br />

T<br />

= (see<br />

s<br />

Table5.1.) and nominal value <strong>of</strong> Ψ<br />

PM<br />

= 0.264Wb<br />

the calculated GMδ () s = 354,82V / rad .<br />

The obtained transfer function between electromagnetic torque<br />

amplitude<br />

U<br />

s<br />

is (see equation 5.72):<br />

M<br />

e<br />

and stator voltage<br />

G<br />

M<br />

M () s A s<br />

() s = =<br />

U s s + B s+ C<br />

(5.27)<br />

e<br />

M<br />

2<br />

s<br />

()<br />

M M<br />

Where<br />

A<br />

M<br />

2<br />

3pbΨ<br />

PM<br />

RsΨ<br />

3 Ψ<br />

PM<br />

s<br />

ΨPM pb<br />

= and BM<br />

= CM<br />

=<br />

2L<br />

Ψ L<br />

2JL<br />

s<br />

Using the motor parameters (see Appendices), one obtains:<br />

A = 198 and B = 115.3 C = 9065<br />

M<br />

M<br />

M<br />

s<br />

s<br />

s<br />

Continuous s-domain<br />

The torque control loop is shown in Fig. 5.16, where C ( s ) is a transfer function <strong>of</strong> the<br />

PI controller given by [105]:<br />

M<br />

80


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

C<br />

M<br />

⎛ K ⎞<br />

iM<br />

KpM<br />

s+<br />

⎜<br />

K ⎟<br />

pM<br />

() s =<br />

⎝ ⎠<br />

(5.28)<br />

s<br />

where<br />

K<br />

iM<br />

=<br />

K<br />

T<br />

pM<br />

iM<br />

M e _ ref<br />

() ∆δ U = sβ Us<br />

CM s Ψ<br />

G ()<br />

GM<br />

() s<br />

Mδ<br />

s<br />

M e<br />

Figure 5.16. Block diagram <strong>of</strong> torque control loop represented in s-domain.<br />

Hence, the transfer function between M<br />

_<br />

() s and M ( s ) is obtained as:<br />

e<br />

ref<br />

e<br />

G<br />

M _ closed<br />

() s<br />

M () s C () s G () s G () s<br />

e_<br />

ref M M Mδ<br />

= = (5.29)<br />

M<br />

e() s 1 + CM() z GM() s GMδ<br />

() s<br />

Substituting transfer function for CM<br />

( s)<br />

and GM<br />

( s ) equation (5.29) becomes:<br />

KiM<br />

K Ψ p PMAMKpM( s +<br />

Ψ<br />

)<br />

K<br />

pM<br />

GM<br />

_ closed()<br />

s = =<br />

2<br />

s + ( B + K A ) s+ C + K A<br />

M pM M M iM M<br />

(5.30)<br />

Discrete design<br />

The transfer function (equations 5.28) for PI controller in discrete system using<br />

z −1<br />

backward difference method for discretization process ( s = ) [2] is expressed as:<br />

Tz<br />

s<br />

K<br />

pM<br />

( z − )<br />

KpM<br />

+ KiM<br />

CM( z) = ( KpM + KiM)<br />

( z −1)<br />

(5.31)<br />

where:<br />

K<br />

iM<br />

K<br />

pM<br />

= Ts<br />

,<br />

s<br />

TiM<br />

T - sampling time, C ( z ) is the discrete transfer function <strong>of</strong><br />

M<br />

torque PI controller, Dz ( )<br />

1<br />

z −<br />

is one sampling time delay for voltage generation from<br />

PWM, and<br />

81


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

D( z)<br />

} G ( ) M<br />

z<br />

M<br />

e_ ref( z)<br />

C ( ) M<br />

z<br />

∆δ Ψ<br />

G<br />

Mδ<br />

( z)<br />

U s<br />

z −1<br />

AM<br />

s<br />

ZOH<br />

2<br />

s + B s+<br />

C<br />

M<br />

M<br />

M ( ) e<br />

z<br />

Figure 5.17. Block diagram <strong>of</strong> torque control loop in discrete domain.<br />

G ( z)<br />

= K Ψ (5.32)<br />

Mδ<br />

pΨ<br />

PM<br />

is discrete transfer function for relation between stator voltage amplitude<br />

increment <strong>of</strong> torque angle ∆ δ Ψ<br />

(see Fig. 5.17)<br />

U s<br />

and<br />

The discrete transfer function G ( z ) for voltage-torque relationship <strong>with</strong> zero order<br />

hold (ZOH) can be calculated as:<br />

M<br />

G () s z −1<br />

A<br />

G z = − z Z = Z<br />

−1<br />

M<br />

M<br />

M<br />

( ) (1 ) [ ] [ ]<br />

2<br />

s z s + BMs+<br />

CM<br />

(5.33)<br />

Finally, the discrete transfer function <strong>of</strong> controlled plant G ( z ) can be written as:<br />

M<br />

⎛ A ⎞<br />

Md<br />

GM<br />

( z) = ( z−1) ⎜ 2<br />

⎟<br />

⎝ z − BMd<br />

z+<br />

CMd<br />

⎠<br />

(5.34)<br />

BM<br />

2<br />

A<br />

− T<br />

M<br />

s<br />

B<br />

2<br />

M<br />

Where: AM _ d<br />

= e sin( T )<br />

2<br />

s<br />

CM<br />

− ,<br />

B<br />

4<br />

M<br />

CM<br />

−<br />

4<br />

BM<br />

2<br />

− Ts<br />

B<br />

2<br />

M<br />

BM _ d<br />

= 2e cos( Ts CM<br />

− ), CM<br />

_<br />

4<br />

d<br />

BM<br />

Ts<br />

= e − , and<br />

s<br />

T is sampling time.<br />

Hence, the transfer function <strong>of</strong> closed torque control loop is obtained in the following<br />

form:<br />

G<br />

M _ closed<br />

( z)<br />

M ( z) C ( z) G ( z) D( z) G ( z)<br />

M ( z) 1 + C ( zG ) ( zDzG ) ( ) ( z)<br />

e_<br />

ref M M Mδ<br />

= = =<br />

e M M Mδ<br />

82


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

K<br />

K Ψ A ( K + K )( z− )( z−1)<br />

pM<br />

pΨ<br />

PM M _ d pM iM<br />

KpM<br />

+ KiM<br />

3 2<br />

M _ d M _ d pM iM M _ d M _ d pM<br />

= =<br />

[ z − B z + [ A ( K + K ) + C ] z− A K ]( z−1)<br />

K<br />

K Ψ A ( K + K )( z−<br />

)<br />

pM<br />

pΨ<br />

PM M _ d pM iM<br />

KpM<br />

+ KiM<br />

3 2<br />

−<br />

M _ d<br />

+ [<br />

M _ d( pM<br />

+<br />

iM) +<br />

M _ d]<br />

−<br />

M _ d pM<br />

z B z A K K C z A K<br />

(5.35)<br />

Selecting<br />

K Ψ<br />

, K Ψ<br />

will influence poles placement <strong>of</strong> closed torque control loop and as<br />

p<br />

i<br />

a consequence also torque step responses can be selected.<br />

The transfer function <strong>of</strong> closed torque control loop is more complicated than flux<br />

control loop (see design <strong>of</strong> P-flux controller – section 5.2.1). One possibility is use to<br />

the SISO tools from Matlab package to tune parameters <strong>of</strong> PI torque controller [106].<br />

a) b)<br />

Figure 5.18. a) <strong>Torque</strong> step response for sampling time T = 200µ<br />

s, b) <strong>with</strong> denoted rise time,<br />

overshoot and settling time.<br />

As can be observed in (Fig. 5.18) torque response is characterized by overshoot about<br />

40%, rise time 4 samples and settling time 17 samples.<br />

s<br />

To eliminate high overshoot it is recommended to insert at the input prefilter (see<br />

Fig.5.19 ) <strong>with</strong> transfer function:<br />

z−<br />

b z−<br />

0.6878<br />

PM<br />

( z)<br />

= K = K<br />

K<br />

pM z − 0.855<br />

( z − )<br />

K + K<br />

1<br />

where K =<br />

=0.466 is gain <strong>of</strong> the prefilter.<br />

z − 0.6878<br />

lim<br />

z→1<br />

z − 0.855<br />

pM<br />

iM<br />

(5.36)<br />

83


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

Dz ( )<br />

G ( ) M<br />

z<br />

M<br />

e_ ref( z)<br />

P ( ) M<br />

z<br />

C ( ) M<br />

z<br />

∆δ Ψ<br />

G<br />

Mδ<br />

( z)<br />

U s<br />

z −1<br />

AM<br />

s<br />

ZOH<br />

2<br />

s + B s+<br />

C<br />

M<br />

M<br />

M ( ) e<br />

z<br />

Figure 5.19. Block diagram <strong>of</strong> torque control loop <strong>with</strong> prefilter (discrete domain).<br />

Finally, the step response <strong>of</strong> closed torque control loop <strong>with</strong> prefilter at the input is<br />

presented bellow:<br />

a) b)<br />

Figure 5.20. a) <strong>Torque</strong> step response for sampling time T = 200µ<br />

s, b) <strong>with</strong> denoted rise time,<br />

overshoot and settling time.<br />

s<br />

As it can be observed, the response is now characterized by overshoot about 2%, rise<br />

time 5 samples and the settling time 15 samples.<br />

In digital control system when the sampling time is changed the parameters <strong>of</strong><br />

digitalized control plant AMd, BMd,<br />

C<br />

Md<br />

will also change. Therefore, the parameters <strong>of</strong> PI<br />

torque controller will change also (see Table 2).<br />

Simulation results for digital torque control loop in SABER package for 5KHz <strong>with</strong> and<br />

<strong>with</strong>out prefilter are shown in Fig. 5.21. Also, the torque step response for different<br />

level <strong>of</strong> reference torque are presented in Fig. 5.22.<br />

The settings for PI torque controller for different sampling time T s<br />

= 50µ s , 100µ s ,<br />

200µ s , 400µ s are summarized in Table 5.2.<br />

84


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

a)<br />

b)<br />

Figure 5.21. <strong>Torque</strong> step response: a) <strong>with</strong>out prefilter, b) <strong>with</strong> prefilter.<br />

Figure 5.22. <strong>Torque</strong> step response <strong>with</strong> prefilter (from 0 to 25%, 50%, 75% and 100% <strong>of</strong><br />

nominal torque).<br />

85


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

86


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

The torque step response is shown in Fig. 5.23, when parameters <strong>of</strong> PI torque controller<br />

designed for sampling time T s<br />

= 200µ s were used for control plant for different sampling<br />

times<br />

T<br />

s<br />

= 50µ s , 100µ s , 200µ s , 400µ s , which correspond to switching frequency f s<br />

=<br />

20kHz, 10kHz, 5kHz, 2.5kHz.<br />

Figure 5.23. <strong>Torque</strong> tracking performance for different sampling time T s<br />

= 50µ s , 100µ s , 200µ s ,<br />

400µ s ( f<br />

s<br />

= 20kHz, 10kHz, 5kHz,2.5kHz) using parameters <strong>of</strong> PI controller designed for<br />

T<br />

s<br />

= 200µ s ( f<br />

s<br />

= 5kHz). (Please note that for 2.5kHz the system was unstable).<br />

After modification <strong>of</strong> PI torque controller parameters according to Table 2 it is possible to<br />

achieve better results as shown in Fig. 5.23, what confirms and proper torque tracking<br />

performance in steady and dynamics state.<br />

Figure 5.24. <strong>Torque</strong> tracking performance for different sampling time T<br />

s<br />

= 50µ s , 100µ s , 200µ s ,<br />

400µ s ( f<br />

s<br />

= 20kHz, 10kHz, 5kHz,2.5kHz) using designed parameters <strong>of</strong> PI controller calculated<br />

individually (see table 5.2)<br />

87


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

5.3 Parallel structure <strong>of</strong> <strong>DTC</strong>–<strong>SVM</strong> scheme<br />

Block scheme <strong>of</strong> the control structure is shown in Fig. 5.25. Two PI controllers are used for<br />

regulation torque and flux magnitude loops [11,55].<br />

U DC<br />

Ψ s _ ref<br />

e Ψ<br />

ss<br />

U sx _ ref<br />

s<br />

_ ref<br />

S A<br />

S B<br />

M e _ ref<br />

e M<br />

U sy _ ref<br />

U α<br />

U<br />

s β _ ref<br />

S C<br />

θ Ψs<br />

Ψ<br />

I s<br />

M e<br />

γ<br />

m<br />

Figure 5.25. Parallel structure <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> scheme.<br />

In this control scheme the reference stator flux magnitude<br />

Ψ and reference<br />

s _ ref<br />

electromagnetic torque<br />

M are compared <strong>with</strong> estimated values, respectively. The flux and<br />

e_<br />

ref<br />

torque errors eΨ , e are delivered to PI controllers, which generate command value the stator<br />

s<br />

M<br />

voltage components in stator flux coordinates<br />

U , U<br />

_<br />

sx_<br />

ref<br />

sy<br />

ref<br />

. This voltage signals are<br />

transformed to stationary coordinates using the stator flux position angle θ Ψ s<br />

. The reference<br />

stator voltage vector ( U<br />

s α _ ref<br />

, U<br />

s β _<br />

ref<br />

) is delivered to space vector modulator (<strong>SVM</strong>), which<br />

generates the switching signals<br />

S , S , S to control power transistors <strong>of</strong> the inverter.<br />

A<br />

B<br />

C<br />

The presented control strategy is based on simplified stator voltage equations described in<br />

stator flux oriented x-y coordinates (equations 2.27a-b):<br />

U<br />

d<br />

Ψ<br />

s<br />

sx<br />

= Rs Isx<br />

+ (5.37)<br />

dt<br />

U = R I +Ω Ψ = R I + E = k M + E<br />

(5.38)<br />

sy s sy Ψs s s sy sy s e sy<br />

88


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

where:<br />

k<br />

s<br />

=<br />

2<br />

3<br />

R<br />

p<br />

= Ψ and Esy = ΩΨs Ψ<br />

s<br />

.<br />

s<br />

b<br />

Ψ , 2<br />

M<br />

e<br />

pb s<br />

Isy<br />

s<br />

3<br />

The above equations show that the<br />

U<br />

sx<br />

component has influence only on the change <strong>of</strong> stator<br />

flux magnitude<br />

Ψ<br />

s<br />

, and the component<br />

Ω<br />

U<br />

sy<br />

– if the term<br />

Ψs s<br />

Ψ is decoupled – can be<br />

used for torque adjustment. Therefore, the flux and torque quantities can be controlled as<br />

shown in Fig. 5.26.<br />

Note, that this <strong>DTC</strong>-<strong>SVM</strong> scheme formally corresponds to the stator flux oriented voltage<br />

source inverter-fed drive induction motor. The block diagram <strong>of</strong> the <strong>DTC</strong>-<strong>SVM</strong> scheme <strong>with</strong><br />

two PI controllers is shown in Fig. 5.26 The dashed line represents the PMSM part [124,125].<br />

Ψ s _ ref<br />

Ψ s<br />

_<br />

R I<br />

s<br />

sx<br />

U sx<br />

R I<br />

s<br />

sx<br />

∫<br />

Ψ s<br />

R I<br />

s<br />

sy<br />

Ω Ψs<br />

⊗<br />

Ω<br />

Ψ<br />

Ψs s<br />

M e _ ref<br />

_<br />

E sy<br />

U sy<br />

1 Isy<br />

R s<br />

3<br />

2<br />

p<br />

b<br />

Ψ<br />

s _ ref<br />

M e<br />

M e<br />

5.3.1 Digital flux control loop<br />

Figure 5.26. Block diagram <strong>of</strong> the scheme presented in Fig. 5.27<br />

Putting the stator x-axis current expression from equation 2.28a under the assumption Ld<br />

= L<br />

q<br />

into equation (5.37) one can obtaines<br />

U<br />

I<br />

sx<br />

Ψs<br />

−ΨPM<br />

cos<br />

= (5.39)<br />

L<br />

s<br />

δ Ψ<br />

Ψs −ΨPM cosδΨ<br />

d Ψs R d Ψ<br />

s s RsΨ<br />

PM<br />

= R ( ) + = Ψ + − cosδΨ<br />

(5.40)<br />

L dt L dt L<br />

sx s s<br />

s s s<br />

Using Laplace transformation to equation 5.40 can be written as:<br />

U<br />

sx<br />

Rs RsΨ<br />

PM<br />

=Ψ<br />

s<br />

( s+ ) − cosδ Ψ<br />

. (5.41)<br />

L L<br />

s<br />

s<br />

89


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

Assuming small changes <strong>of</strong> δ Ψ<br />

, the cosδ Ψ<br />

≅ 1, and equations (5.41) reduces to:<br />

Rs<br />

Usx<br />

=Ψ<br />

s<br />

( s+ ) + WΨ<br />

(5.42)<br />

L<br />

s<br />

RsΨ<br />

PM<br />

Where WΨ<br />

=<br />

L<br />

s<br />

The transfer function between stator flux amplitude<br />

Ψ<br />

s<br />

and x-axis <strong>of</strong> stator voltage is:<br />

G<br />

Ψ<br />

Ψ<br />

s Ls<br />

1 1<br />

() s = = = =<br />

U R<br />

sx<br />

+ WΨ<br />

sLs + Rs<br />

s<br />

s +<br />

s+<br />

A<br />

L<br />

s<br />

Ψ<br />

(5.43)<br />

Where W<br />

Ψ<br />

Rs<br />

Rs<br />

≅ ΨPM<br />

cosδΨ<br />

≅ Ψ<br />

L<br />

L<br />

s<br />

Rs<br />

For motor parameters (see Appendices ): A = Ψ<br />

115.333<br />

L<br />

= .<br />

s<br />

PM<br />

Continuous s-domain<br />

The flux control loop is shown in Fig. 5.27, where CΨ ( s)<br />

is a transfer function <strong>of</strong> the PI<br />

controller given by [105]:<br />

C<br />

Ψ<br />

s<br />

⎛ K ⎞<br />

iΨ<br />

KpΨ<br />

s+<br />

⎜<br />

K ⎟<br />

pΨ<br />

() s =<br />

⎝ ⎠<br />

(5.45)<br />

s<br />

where<br />

K<br />

iΨ<br />

=<br />

K<br />

T<br />

pΨ<br />

iΨ<br />

Ψ s_ref<br />

C<br />

Ψ<br />

() s<br />

U sx<br />

W Ψ<br />

()<br />

Rs<br />

Ψ<br />

L<br />

s<br />

PM<br />

cos<br />

δ Ψ<br />

+<br />

G<br />

Ψ<br />

s<br />

Ψ s<br />

Figure 5.27. Block diagram <strong>of</strong> flux control loop in s-domain.<br />

Hence, the transfer function <strong>of</strong> the closed stator flux amplitude control loop is obtained as:<br />

90


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

G<br />

Ψ _ closed<br />

Ψ<br />

s<br />

() s<br />

_ ref CΨ() s GΨ()<br />

s<br />

() s = =<br />

Ψ () s 1 + C () z G () s<br />

s<br />

Ψ<br />

Ψ<br />

(5.46)<br />

Substituting transfer function for CΨ ( s)<br />

and GΨ ( s)<br />

one becomes<br />

G<br />

_ closed<br />

() s<br />

⎛ K ⎞<br />

iΨ<br />

KpΨ<br />

s+<br />

⎜ K ⎟<br />

⎝ pΨ<br />

⎠ ⎛ 1 ⎞<br />

⎛ K ⎞<br />

iΨ<br />

⎜ ⎟ KpΨ<br />

s+<br />

s s A<br />

⎜ K ⎟<br />

⎝ + ⎠ pΨ<br />

=<br />

⎝ ⎠<br />

⎛ K ⎞<br />

s + ( A + Kp<br />

) s+<br />

K<br />

iΨ<br />

KpΨ<br />

s+<br />

⎜ K ⎟<br />

pΨ<br />

⎛ 1 ⎞<br />

1+ ⎝ ⎠<br />

⎜ ⎟<br />

s ⎝s+<br />

AΨ<br />

⎠<br />

Ψ<br />

Ψ<br />

=<br />

2<br />

Ψ Ψ iΨ<br />

(5.47)<br />

Discrete design<br />

z −1<br />

Using backward difference method for discretization process ( s = ) [2] the transfer<br />

Tz<br />

function <strong>of</strong> equation (5.45) for flux PI controller in discrete system is expressed as:<br />

s<br />

K<br />

pΨ<br />

( KpΨ<br />

+ KiΨ)( z−<br />

)<br />

Tz<br />

K<br />

s<br />

pΨ<br />

+ KiΨ<br />

CΨ( z) = KpΨ(1 + ) =<br />

T ( z−1) ( z−1)<br />

iΨ<br />

(5.48)<br />

Where:<br />

K<br />

s<br />

K<br />

pΨ<br />

iΨ<br />

= Ts; s<br />

Ti<br />

Ψ<br />

Ψ<br />

_<br />

( z)<br />

ref<br />

T - sampling time.<br />

C<br />

Ψ<br />

( z)<br />

W( z)<br />

U sx<br />

Dz ( )<br />

z −1<br />

GΨ<br />

( z)<br />

}<br />

ZOH<br />

1<br />

s+<br />

A Ψ<br />

Ψ<br />

s<br />

( z)<br />

Figure 5.28. Block diagram <strong>of</strong> flux control loop in discrete domain.<br />

Where: CΨ ( z)<br />

discrete transfer function <strong>of</strong> PI controller, Dz ( )<br />

1<br />

z − - one sampling time delay<br />

for voltage generation from PWM, and W( z)<br />

- disturbance voltage due to cross coupling<br />

between x-y axis (see Fig. 5.28).<br />

The GΨ ( z)<br />

is discrete transfer function <strong>of</strong> voltage-flux relationship <strong>with</strong> zero order hold<br />

(ZOH) block can be calculated as:<br />

91


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

−1 GΨ() s z −1<br />

AΨ<br />

GΨ<br />

( z) = (1 − z ) Z[ ] = Z[ ] =<br />

s z A s( s+<br />

A )<br />

( z − )<br />

1 1 AΨ<br />

Z[ ]<br />

z A s( s+<br />

A )<br />

Ψ<br />

Ψ<br />

Ψ<br />

Ψ<br />

(5.49)<br />

Using table <strong>of</strong> Z transformation [2] one can calculate:<br />

G<br />

Ψ<br />

( z − )<br />

− AΨTs<br />

1 1 z(1 − e ) AΨ<br />

d<br />

( z)<br />

=<br />

− A T<br />

z A<br />

Ψ s<br />

=<br />

z 1( z−e<br />

) z − B<br />

Ψ<br />

( − )<br />

Ψd<br />

(5.50)<br />

Where:<br />

A<br />

Ψd<br />

− AΨT (1 − e s<br />

)<br />

− A Ts<br />

= , BΨ d<br />

e Ψ<br />

A<br />

= and T<br />

s<br />

is sampling time.<br />

Ψ<br />

Hence, the transfer function <strong>of</strong> closed stator flux control loop can be expressed in the<br />

following form:<br />

G<br />

Ψ _ closed<br />

Ψ<br />

s<br />

( z) _ ref CΨ( z) GΨ( z) D( z)<br />

( z)<br />

= =<br />

Ψ ( z) 1 + C ( z) G ( z) D( z)<br />

s<br />

K<br />

pΨ<br />

( KpΨ + KiΨ) AΨd( z−<br />

)<br />

KpΨ<br />

+ KiΨ<br />

=<br />

K<br />

pΨ<br />

zz ( −1)( z− BΨd) + ( KpΨ + KiΨ) AΨd( z−<br />

)<br />

K + K<br />

Ψ<br />

Ψ<br />

pΨ<br />

iΨ<br />

(5.51)<br />

Now selecting<br />

K Ψ<br />

, K Ψ<br />

is possible to obtain poles placement, which define the dynamic <strong>of</strong><br />

p<br />

i<br />

closed torque control loop.<br />

Assuming, that<br />

B<br />

Ψd<br />

=<br />

K<br />

K<br />

pΨ<br />

pΨ<br />

+ K<br />

iΨ<br />

KpΨ − BΨdKpΨ KpΨ<br />

⇒ KiΨ<br />

= = (1 − BΨd)<br />

B B<br />

Ψd<br />

Ψd<br />

and the transfer function <strong>of</strong> closed stator flux control loop will take the following form:<br />

G<br />

( KpΨ + KiΨ)<br />

AΨd<br />

( z)<br />

=<br />

z − z+ ( K + K ) A<br />

Ψ _ closed<br />

2<br />

pΨ iΨ Ψd<br />

(5.52)<br />

Putting into above equation<br />

K<br />

K<br />

pΨ<br />

pΨ<br />

+ KiΨ<br />

= one obtains:<br />

BΨd<br />

G<br />

K<br />

A<br />

CΨd<br />

K z − z+<br />

C<br />

pΨ<br />

Ψd<br />

BΨd<br />

Ψ _ closed<br />

( z)<br />

= =<br />

2<br />

2<br />

pΨ<br />

z − z+<br />

AΨ<br />

d<br />

BΨd<br />

Ψd<br />

(5.53)<br />

92


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

K<br />

pΨ<br />

where CΨd<br />

= AΨd<br />

BΨd<br />

The bellow diagrams shows the relationship between overshoot<br />

time t s<br />

as function<br />

C Ψ d<br />

value.<br />

M<br />

p<br />

, rise time t r<br />

and settling<br />

Please note that t r<br />

is time calculated from 10% to 90% <strong>of</strong> output signals and t s<br />

is the time in<br />

witch the system transient decay to +-1%.<br />

Figure 5.29. The relationship between overshoot, rise time and settling time versus<br />

amplitude control loop.<br />

From a few values <strong>of</strong><br />

C Ψ d<br />

=[0.4046, 0.2688, 0.1720] we can selected C Ψ d<br />

guaranties overshoot 0% and settling time about 10 samples.<br />

C Ψ d<br />

for stator flux<br />

=0.2688, which<br />

1<br />

2<br />

3<br />

Figure 5.30. Flux step response for different values <strong>of</strong><br />

C Ψ d<br />

=0.2688, black line (3) C Ψ d<br />

=0.1720.<br />

C Ψ d<br />

: red line (1) C Ψ d<br />

=0.4046, blue line (2)<br />

93


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

It corresponds to the transfer function <strong>of</strong> closed stator flux control loop:<br />

G<br />

C<br />

0.2688<br />

( z)<br />

= =<br />

z − z+ C z −z+ 0.2688<br />

Ψd<br />

Ψ _ closed<br />

2 2<br />

Ψd<br />

(5.54)<br />

Using digitalized motor parameters A<br />

can calculate the parameters <strong>of</strong> digital PI flux controller as:<br />

Ψd<br />

− AΨT (1 − e s<br />

)<br />

− A Ts<br />

= , BΨ d<br />

e Ψ<br />

A<br />

= and chosen C Ψ d<br />

value, we<br />

Ψ<br />

K<br />

pΨ<br />

C B<br />

A<br />

Ψd<br />

Ψd<br />

= (5.55a)<br />

Ψd<br />

T<br />

iΨ<br />

K<br />

T<br />

pΨ<br />

s<br />

= (5.55b)<br />

K<br />

iΨ<br />

K<br />

iΨ<br />

K<br />

pΨ<br />

= (1 − BΨd)<br />

(5.55c)<br />

B<br />

Ψd<br />

For example <strong>with</strong> sampling time T = 200µ<br />

s, parameters <strong>of</strong> PI controller are:<br />

s<br />

K<br />

pΨ<br />

0.2688BΨd<br />

0.2688*0.9772<br />

= = = 1328.64<br />

(5.56a)<br />

A 0.0001977<br />

Ψd<br />

T<br />

iΨ<br />

KpΨTs<br />

1328.64*200µ<br />

s<br />

= = = 8572µ<br />

s<br />

(5. 56b)<br />

K 30.999<br />

iΨ<br />

K<br />

iΨ<br />

K<br />

pΨ<br />

1328.64<br />

= (1 − BΨd) = *(1-0.9772) = 30.999 (5. 56c)<br />

B<br />

0.9772<br />

Ψd<br />

For different sampling time the closed transfer function GΨ _ closed<br />

( z)<br />

<strong>of</strong> digital flux control<br />

loop should be kept to:<br />

G<br />

C<br />

0.2688<br />

( z)<br />

= =<br />

z − z+ C z −z+ 0.2688<br />

Ψd<br />

Ψ _ closed<br />

2 2<br />

Ψd<br />

(5.57)<br />

In order to find the original function <strong>of</strong> Z transfer function GΨ _ closed<br />

( z)<br />

using the Z properties<br />

as (sum transformations) [2]:<br />

n<br />

z<br />

z<br />

∑ f kT Z F z Z G z<br />

k = 0<br />

z−1 z−1<br />

−1<br />

z a<br />

= Z [ ]<br />

2<br />

z−1( z − z+<br />

a)<br />

−1 −1<br />

(<br />

s) = [ ( )] = [<br />

Ψ _ closed( )]<br />

(5.58)<br />

94


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

As an example the calculated response for 4 samples are given:<br />

n<br />

∑<br />

k = 0<br />

f kT az az a a az<br />

−2 −3 −4<br />

(<br />

s<br />

) = + 2 − (1 + ) + 2 + .......<br />

which gives:<br />

f(0) = f( k) = 0<br />

f(1 T ) = f( k+ 1) = 0<br />

s<br />

f(2 T ) = f( k+ 2) = a=<br />

0.2688<br />

s<br />

f(3 T ) = f( k+ 3) = a( b+ 1) = 2a=<br />

0.5376<br />

s<br />

f T f k a b c a b a a a<br />

2<br />

(4<br />

s<br />

) = ( + 4) =− ( + ) + ( + 1) =− (1 + ) + 2 = 0.7342<br />

Keeping constant<br />

C Ψ d<br />

in equation (5.57) the flux step response for different sampling time<br />

T<br />

s<br />

= 50 µ s,<br />

100 µ s,<br />

200 µ s,<br />

400µ s , which correspond to switching frequency f s<br />

= 20kHz,<br />

10kHz, 5kHz, 2.5kHz are presented.<br />

1<br />

2<br />

3<br />

4<br />

Figure 5.31.Flux step response for different sampling time T<br />

s<br />

= 50 µ s,<br />

100 µ s,<br />

200 µ s,<br />

400µ<br />

s<br />

(switching frequency f<br />

s<br />

20kHz (1 -blue line), 10kHz (2 -green line), 5kHz (3 -red line),2.5kHz<br />

(4 -light blue line).<br />

We may observe from Fig. 5.31 that overshoot is 0% and the settling time is 10 samples.<br />

Selected parameters <strong>of</strong> PI flux controller for sampling time T s<br />

= 50 µ s,<br />

100 µ s,<br />

200 µ s,<br />

400µ s are summarized in Table 5.3<br />

95


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

96


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

The behavior <strong>of</strong> the flux control loop was tested using SABER simulation package. The<br />

model created in SABER takes into account the whole control system, which include real<br />

models <strong>of</strong> inverter and permanent magnet synchronous motor.<br />

The flux step response is shown in Fig. 5.32., when parameters <strong>of</strong> PI flux controller designed<br />

for sampling time T s<br />

= 200µ s were used for control plant for different sampling times<br />

T<br />

s<br />

= 50µ s , 100µ s , 200µ s , 400µ s , which correspond to switching frequency f<br />

s<br />

= 20kHz,<br />

10kHz, 5kHz, 2.5kHz.<br />

Figure 5.32. Flux tracking performance for different sampling time T<br />

s<br />

= 50µ s , 100µ s , 200µ s ,<br />

400µ s ( f<br />

s<br />

= 20kHz, 10kHz, 5kHz,2.5kHz) using parameters <strong>of</strong> PI controller designed for<br />

T<br />

s<br />

= 200µ s ( f<br />

s<br />

= 5kHz).<br />

After modification <strong>of</strong> PI flux controller parameters according to Table 5.31 it is possible to<br />

achieve better results as shown in Fig. 5.33, what confirms proper flux tracking performance<br />

in steady and dynamics state.<br />

97


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

Figure 5.33. Flux tracking performance for different sampling time T<br />

s<br />

= 50µ s , 100µ s , 200µ s ,<br />

400µ s ( f<br />

s<br />

= 20kHz, 10kHz, 5kHz,2.5kHz) using parameters <strong>of</strong> PI controller calculated individually<br />

(see table 5.3.)<br />

The simulation results in SABER package for Ts=200us is presented in Fig. 5.34.<br />

Figure 5.34. Simulated (SABER) flux tracking performance for step change from 70% - 100% <strong>of</strong><br />

nominal flux.<br />

As we can observed from Fig. 5.34 that overshoot is around M<br />

p<br />

= 0% and settling time took<br />

about 10 sampling time <strong>of</strong> microprocessor, what proved the design procedure <strong>of</strong> PI digital<br />

flux controller parameters.<br />

98


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

5.3.2 Digital torque control loop<br />

The PMSM equations (2.27a,b-2.28a,b) in stator flux coordinates under the assumption<br />

L<br />

d<br />

= L can be written as:<br />

q<br />

U = R I +ΩΨ<br />

Ψ (5.59)<br />

sy s sy s s<br />

0 LI sin<br />

= −Ψ (5.60)<br />

s sy PM<br />

δ Ψ<br />

3<br />

M<br />

e<br />

= pb Ψ I<br />

s sy<br />

(5.61)<br />

2<br />

dΩ m 1 = ( M<br />

e − M<br />

l )<br />

(5.62)<br />

dt J<br />

The load angle can be expressed (Fig. 5.1):<br />

δ = θ − p γ , (5.63)<br />

Ψ<br />

Ψs b m<br />

Where: δ Ψ<br />

is torque angle, θ Ψ s<br />

is stator flux vector position, and γ<br />

m<br />

is rotor position in stator<br />

α,<br />

β coordinates,<br />

p<br />

b<br />

is number <strong>of</strong> pole pars.<br />

After differentiation equation (5.63) can be written as:<br />

dδ<br />

dθ<br />

dγ<br />

dt dt dt<br />

Ψ Ψs<br />

m<br />

= − pb<br />

(5.64)<br />

δ Ψ<br />

d<br />

dt<br />

δ Ψ<br />

d<br />

=ΩΨ<br />

−p<br />

s bΩm<br />

⇒Ω Ψ s<br />

= + pbΩ m<br />

(5.65)<br />

dt<br />

Putting equations (5.64) and (5.65) into voltage equation (5.59) one obtains:<br />

dδ Usy = RsIsy +Ω ( )<br />

s s<br />

RsI Ψ<br />

Ψ<br />

Ψ =<br />

sy<br />

+ Ψ<br />

s<br />

+ pbΩ m<br />

(5.66)<br />

dt<br />

From equation 0= LI −Ψ sin <strong>with</strong> assumption that for small angle δ = sinδ<br />

, the<br />

s sy PM<br />

torque angle can be expressed as:<br />

δ Ψ<br />

Ψ<br />

Ψ<br />

L I<br />

s sy<br />

δ Ψ<br />

= (5.67)<br />

Ψ<br />

PM<br />

So, the voltage equation (5.59) becomes:<br />

L dI<br />

s sy<br />

Usy = Rs Isy +ΩΨ<br />

Ψ ( )<br />

s s<br />

= Rs Isy + Ψ<br />

s<br />

+ pbΩm<br />

Ψ dt<br />

PM<br />

(5.68)<br />

99


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

After differentiating <strong>of</strong> the above equation one obtains:<br />

2<br />

sy sy Ls<br />

sy<br />

= Rs + Ψ<br />

s<br />

+ pb<br />

Ψ<br />

PM<br />

dU dI d I dΩm<br />

( )<br />

dt dt dt dt<br />

Take into account that from equation (5.61) the y-axis current is equal<br />

I<br />

sy<br />

(5.69)<br />

2M<br />

e<br />

=<br />

3p<br />

Ψ ,<br />

b<br />

s<br />

dΩ m 1 = ( M<br />

e − M<br />

l ) and under assumption that the motor is no loaded equation (5.69) takes<br />

dt J<br />

form:<br />

dU dM L d M p<br />

( )<br />

dt p dt p dt J<br />

2<br />

sy 2 e s 2<br />

e b<br />

= Rs + Ψ<br />

s<br />

+ M<br />

e<br />

3<br />

b<br />

Ψs ΨPM 3<br />

b<br />

Ψs<br />

(5.70)<br />

Using Laplace transformation and after some arrangements the equation (5.70) can be written:<br />

2Ls<br />

2 2R<br />

Ψs<br />

p<br />

s<br />

b<br />

sU<br />

sy<br />

= M<br />

e<br />

( s + s + )<br />

3p Ψ 3p Ψ J<br />

b PM b s<br />

(5.71)<br />

Hence, the transfer function between electromagnetic torque<br />

be obtained as:<br />

G<br />

M<br />

e<br />

M<br />

2<br />

sy()<br />

M M<br />

M<br />

e<br />

and y-axis voltage<br />

U<br />

sy<br />

can<br />

M () s A s<br />

() s = =<br />

U s s + B s+ C<br />

(5.72)<br />

Where:<br />

A<br />

M<br />

2<br />

3pbΨ<br />

PM<br />

RsΨ<br />

3 Ψ<br />

PM<br />

s<br />

ΨPM pb<br />

= and BM<br />

= CM<br />

=<br />

2L<br />

Ψ L<br />

2JL<br />

s<br />

s<br />

s<br />

s<br />

Using the motor parameters (see Appendices) we may calculates:<br />

A = 198 , B = 115.3 and C = 9065<br />

M<br />

M<br />

M<br />

Continuous s-domain<br />

The torque control loop <strong>of</strong> the block scheme <strong>DTC</strong>-<strong>SVM</strong> from Fig. 5.25 is shown in Fig. 5.35,<br />

where CM<br />

( s ) is a transfer function <strong>of</strong> the PI controller given by equation 5.28:<br />

100


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

M e _ ref<br />

U sy<br />

C () M<br />

s GM<br />

() s<br />

M e<br />

Figure 5.35. Block diagram <strong>of</strong> torque control loop in s-domain.<br />

The transfer function <strong>of</strong> torque control loop is obtained as:<br />

G<br />

M _ closed<br />

() s<br />

M () s C () s G () s<br />

e_<br />

ref M M<br />

= = (5.73)<br />

M<br />

e() s 1 + CM() z GM()<br />

s<br />

Substituting in equation (5.73) transfer function for CM<br />

( s)<br />

-Eq.5.28 and GM<br />

( s)<br />

- Eq.5.72 we<br />

may calculate:<br />

KiM<br />

AMKpM( s+<br />

)<br />

K<br />

pM<br />

GM<br />

_ closed()<br />

s = =<br />

2<br />

s + ( B + K A ) s+ C + K A<br />

M pM M M iM M<br />

(5.74)<br />

Discrete design<br />

z −1<br />

Using backward difference method for discretization process ( s = ) the transfer function<br />

Tz<br />

for discrete PI controller is expressed as:<br />

s<br />

K<br />

pM<br />

( z − )<br />

Tz<br />

KpM<br />

+ K<br />

s<br />

iM<br />

CM( z) = KpM(1 + ) = ( KpM + KiM)<br />

T ( z−1) ( z−1)<br />

iM<br />

(5.75)<br />

Where:<br />

K<br />

iM<br />

K<br />

pM<br />

= Ts<br />

- integration gain;<br />

s<br />

TiM<br />

T - sampling time<br />

101


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

PI controller<br />

Time delay<br />

G ( ) M<br />

z<br />

Dz ( )} Plant<br />

M<br />

e_ ref( z)<br />

C ( ) M<br />

z<br />

U sy<br />

z −1<br />

AM<br />

s<br />

ZOH<br />

2<br />

s + B s+<br />

C<br />

M<br />

M<br />

M ( ) e<br />

z<br />

Figure 5.36. Block diagram <strong>of</strong> torque control loop in discrete domain.<br />

Where: CM<br />

( z)<br />

- discrete transfer function for PI controller, Dz ( )<br />

delay for voltage generation from PWM block (see Fig. 5.36).<br />

1<br />

z − - one sampling time<br />

The discrete transfer function G ( z ) for voltage-torque relationship <strong>with</strong> zero order hold<br />

(ZOH) can be calculated as:<br />

M<br />

G () s z −1<br />

A<br />

G z z Z Z<br />

−1<br />

M<br />

M<br />

M<br />

( ) = (1 − ) [ ] = [ ]<br />

2<br />

s z s + BMs+<br />

CM<br />

⎡<br />

⎤<br />

⎡ ⎤ ⎢ ⎥<br />

⎢ ⎥<br />

z−1 AM<br />

z−1<br />

⎢ A<br />

⎥<br />

M<br />

= Z ⎢<br />

⎥ = Z<br />

2<br />

⎢<br />

2<br />

2<br />

⎥ =<br />

z ⎢⎛ B 2<br />

M ⎞ B ⎥ z<br />

M ⎢ B ⎛<br />

M 2 B ⎞ ⎥<br />

⎢⎜s+ ⎟ + CM<br />

− ⎥<br />

M<br />

2 4 ⎢( s+ ) + CM<br />

−<br />

⎥<br />

⎣⎝ ⎠ ⎦<br />

⎢ 2 ⎜ 4 ⎟<br />

⎣ ⎝ ⎠ ⎥⎦<br />

.<br />

⎡<br />

⎤<br />

2<br />

⎢<br />

B ⎥<br />

M<br />

CM<br />

−<br />

z −1 A ⎢<br />

M<br />

Z<br />

4<br />

⎥<br />

= ⎢<br />

2<br />

2<br />

⎥<br />

z B 2<br />

M<br />

⎢<br />

C B ⎛<br />

M 2 B ⎞ ⎥<br />

M<br />

M<br />

−<br />

4<br />

⎢( s+ ) + CM<br />

−<br />

⎥<br />

⎢ 2 ⎜ 4 ⎟<br />

⎣ ⎝ ⎠ ⎥⎦<br />

(5.76)<br />

Assuming that<br />

have:<br />

B<br />

a = M<br />

and<br />

2<br />

2<br />

BM<br />

b= CM<br />

− , and using table <strong>of</strong> Z transformation [2] we<br />

4<br />

−aTs<br />

⎡ b ⎤ ze sin( bTs<br />

)<br />

Z ⎢ 2 2 2 aTs<br />

( s a) b<br />

⎥ =<br />

−<br />

⎣ + + ⎦ z − 2 e (cos( bTs<br />

)) z+<br />

e<br />

−2aTs<br />

(5.77)<br />

102


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

Finally, the discrete transfer function <strong>of</strong> controlled plant G ( z ) can be written as:<br />

M<br />

Where:<br />

C<br />

M _ d<br />

⎛ A ⎞<br />

Md<br />

GM<br />

( z) = ( z−1) ⎜ 2<br />

⎟<br />

⎝ z − BMd<br />

z+<br />

CMd<br />

⎠<br />

2<br />

A<br />

BM<br />

A = e sin( T C − ) ,<br />

4<br />

BM<br />

− T<br />

M<br />

s<br />

2<br />

M _ d<br />

2<br />

s M<br />

BM<br />

CM<br />

−<br />

4<br />

BM<br />

Ts<br />

= e − and T<br />

s<br />

is sampling time.<br />

(5.78)<br />

BM<br />

2<br />

− Ts<br />

B<br />

2<br />

M<br />

BM _ d<br />

= 2e cos( Ts CM<br />

− )<br />

4<br />

Hence, the transfer function <strong>of</strong> closed torque control loop is obtained as:<br />

G<br />

M _ closed<br />

( z)<br />

M ( z) C ( z) G ( z) D( z)<br />

M ( z) 1 + C ( z) G ( z) D( z)<br />

e_<br />

ref M M<br />

= = =<br />

e M M<br />

K<br />

A ( K + K )( z−<br />

)<br />

pM<br />

M _ d pM iM<br />

KpM<br />

+ KiM<br />

3 2<br />

−<br />

M _ d<br />

+ [<br />

M _ d( pM<br />

+<br />

iM) +<br />

M _ d]<br />

−<br />

M _ d pM<br />

=<br />

z B z A K K C z A K<br />

(5.79)<br />

Selecting<br />

K Ψ<br />

, K Ψ<br />

will influence poles placement <strong>of</strong> closed torque control loop and as a<br />

p<br />

i<br />

consequence also torque step responses can be selected.<br />

The transfer function <strong>of</strong> closed torque control loop is more complicated than flux control loop<br />

(see design <strong>of</strong> PI-flux controller – section 5.3.1). One possibility is use to the SISO tools from<br />

Matlab package to tune parameters <strong>of</strong> PI torque controller [106].<br />

a) b)<br />

Figure 5.37. a) <strong>Torque</strong> step response for sampling time T = 200µ<br />

s, b) <strong>with</strong> denoted rise time,<br />

overshoot and settling time.<br />

As can be observed the response is characterized by overshoot about 40%, rise time 4 samples<br />

and settling time 17 samples.<br />

s<br />

103


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

The torque control loop should be as fast as possible even <strong>with</strong> some overshoot. This improve<br />

response to disturbance (for example from flux control loop –see Fig. 5.38).<br />

M<br />

e_ ref( z)<br />

C ( ) M<br />

z<br />

U sy<br />

Dz ( )<br />

z −1<br />

G ( ) M<br />

z<br />

}<br />

AM<br />

s<br />

ZOH<br />

2<br />

s + B s+<br />

C<br />

M<br />

M<br />

M ( ) e<br />

z<br />

Figure 5.38. Block diagram <strong>of</strong> torque controller in discrete domain <strong>with</strong> disturbance.<br />

a )<br />

b)<br />

Figure 5.39. Disturbance rejection in torque control loop: a) short voltage impulse, b) voltage<br />

step.<br />

To improve reference tracking performance (<strong>with</strong>out any overshoot) it is recommended to<br />

insert a input prefilter (see Fig. 5.40 ) described by transfer function:<br />

z−b z−b z−0.6663<br />

PM<br />

( z)<br />

= K = K = K<br />

(5.80)<br />

K<br />

pM ( z−a) z−0.855<br />

( z − )<br />

K + K<br />

pM<br />

iM<br />

Where:<br />

K<br />

1<br />

=<br />

=0.43413 is gain <strong>of</strong> the prefilter.<br />

z − 0.6663<br />

lim<br />

z→1<br />

z − 0.855<br />

104


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

M<br />

e_ ref( z)<br />

P ( ) M<br />

z<br />

C ( ) M<br />

z<br />

U sy<br />

Dz ( )<br />

z −1<br />

G ( ) M<br />

z<br />

}<br />

AM<br />

s<br />

ZOH<br />

2<br />

s + B s+<br />

C<br />

M<br />

M<br />

M ( ) e<br />

z<br />

Figure 5.40. Block diagram <strong>of</strong> torque control loop <strong>with</strong> prefilter in discrete domain.<br />

Finally, the reference tracking performance <strong>of</strong> closed torque control loop <strong>with</strong> prefilter is<br />

presented in Fig. 5.41.<br />

Figure 5.41. a) Reference tracking performance <strong>of</strong> the torque control loop for sampling time<br />

Ts<br />

= 200µ<br />

s, b) <strong>with</strong> denoted rise time, overshoot and settling time.<br />

In digital control when the sampling time is changing the parameters <strong>of</strong> digitalized plant<br />

control AMd, BMd,<br />

C<br />

Md<br />

will be also changes. It is normally that the parameters <strong>of</strong> PI torque<br />

control will be changes also (see Table 5.4.).<br />

Simulation for PI flux calculated parameters <strong>with</strong> and <strong>with</strong>out prefilter are shown in Fig.<br />

5.42a-b. Also torque step response for different level <strong>of</strong> reference torque are presented in Fig.<br />

5.43.<br />

105


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

a)<br />

b)<br />

Figure 5.42. <strong>Torque</strong> step response: a) <strong>with</strong>out prefilter, b) <strong>with</strong> prefilter.<br />

Figure 5.43. <strong>Torque</strong> step response <strong>with</strong> prefilter (from 0 to 25%, 50%, 75% and 100% <strong>of</strong> nominal<br />

torque).<br />

106


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

107


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

The behavior <strong>of</strong> the torque control loop was tested using SABER simulation package.<br />

The model created in SABER takes into account the whole control system, which<br />

include real models <strong>of</strong> inverter and permanent magnet synchronous motor.<br />

The torque step response is shown in Fig. 5.44, when parameters <strong>of</strong> PI torque controller<br />

designed for sampling time T s<br />

= 200µ s were used for control plant for different<br />

sampling times T s<br />

= 50µ s , 100µ s , 200µ s , 400µ s , which correspond to switching<br />

frequency f<br />

s<br />

= 20kHz, 10kHz, 5kHz, 2.5kHz.<br />

Figure 5.44. <strong>Torque</strong> tracking performance for different sampling time T s<br />

= 50µ s , 100µ s ,<br />

200µ s , 400µ s ( f<br />

s<br />

= 20kHz, 10kHz, 5kHz,2.5kHz) using parameters <strong>of</strong> PI controller<br />

designed for T<br />

s<br />

= 200µ s ( f<br />

s<br />

= 5kHz). Pleas note that for 2.5kHz the system was unstable.<br />

After modification <strong>of</strong> PI flux controller parameters according to Table 5.4 it is possible<br />

to achieve better results as shown in Fig. 5.45, what confirms proper torque tracking<br />

performance in steady and dynamics state.<br />

108


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

Figure 5.45. <strong>Torque</strong> tracking performance for different sampling time T s<br />

= 50µ s , 100µ s ,<br />

200µ s , 400µ s ( f<br />

s<br />

= 20kHz, 10kHz, 5kHz,2.5kHz) using parameters <strong>of</strong> PI controller<br />

calculated individually (see Table 5.4.)<br />

109


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

5.4 Speed control loop for <strong>DTC</strong>-<strong>SVM</strong> structure<br />

The structure <strong>of</strong> speed control loop for cascade (a) and parallel (b) <strong>DTC</strong>-<strong>SVM</strong> scheme<br />

is shown in the Fig. 5.46.<br />

a)<br />

U DC<br />

Ψ s _ ref<br />

U<br />

s α _ ref<br />

S A<br />

S B<br />

Ω m _ ref<br />

M e _ ref<br />

e M<br />

∆δ Ψ<br />

U<br />

s β _ ref<br />

S C<br />

θ Ψs Ψs<br />

I s<br />

M e<br />

I s<br />

d<br />

dt<br />

γ m<br />

U DC<br />

b)<br />

Ψ s _ ref<br />

e Ψ<br />

ss<br />

U sx _ ref<br />

s<br />

_ ref<br />

S A<br />

S B<br />

Ω m _ ref<br />

M e _ ref<br />

e M<br />

U sy _ ref<br />

U α<br />

U<br />

s β _ ref<br />

S C<br />

θ Ψs<br />

Ψ<br />

I s<br />

M e<br />

d<br />

dt<br />

Figure 5.46. Speed control loop for: a) cascade structure <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> scheme, b) parallel<br />

structure <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> scheme.<br />

γ<br />

m<br />

The mechanical speed equation (2.66) for the PMSM is:<br />

dΩ<br />

M M J dt<br />

m<br />

e<br />

−<br />

L<br />

= (5.81)<br />

Taking Laplace transformation to equation (5.81) one obtains:<br />

M () s − M () s = JsΩ () s<br />

(5.82)<br />

e L m<br />

The transfer function between mechanical rotor speed<br />

M<br />

e<br />

can be expressed as:<br />

Ω<br />

m<br />

and electromagnetic torque<br />

110


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

Ωm<br />

1 1<br />

GΩ<br />

() s = = = (5.83)<br />

M Js Js<br />

e<br />

Continuous s-domain<br />

The block diagram <strong>of</strong> speed control loop is shown in Fig. 5.47, where CΩ ( s)<br />

is a<br />

transfer function <strong>of</strong> the PI speed controller given by:<br />

KiΩ<br />

KpΩ( s+<br />

)<br />

1<br />

K<br />

pΩ<br />

CΩ() s = KpΩ(1 + ) = (5.84)<br />

T s s<br />

iΩ<br />

and DΩ ( s)<br />

is approximated transfer function <strong>of</strong> closed torque control loop for cascade<br />

or parallel <strong>DTC</strong>-<strong>SVM</strong> structure.<br />

M L<br />

Ω m _ ref<br />

M e ref<br />

_<br />

CΩ () s<br />

D () s<br />

G () s<br />

Ω<br />

M e<br />

Ω<br />

Ω m<br />

Figure 5.47. Block diagram <strong>of</strong> speed control loop in s-domain.<br />

Discrete design<br />

The transfer function for PI controller in discrete system using backward difference<br />

method for discretization process is expressed as:<br />

K<br />

pΩ<br />

( z − )<br />

KpΩ<br />

+ KiΩ<br />

CΩ( z) = ( KpΩ + KiΩ)<br />

(5.85)<br />

( z −1)<br />

K<br />

pΩ<br />

Where: KiΩ<br />

= Ts- integration and K<br />

p Ω<br />

proportional gain <strong>of</strong> speed controller, T s<br />

-<br />

T<br />

sampling time.<br />

iΩ<br />

111


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

Ω m _ ref<br />

( z)<br />

C<br />

Ω<br />

( z)<br />

Me_ ref()<br />

z<br />

D ( z)<br />

Ω<br />

M ( ) e<br />

z<br />

M ( L<br />

z )<br />

GΩ<br />

( z)<br />

}<br />

ZOH<br />

1<br />

Js<br />

Ω ( z m<br />

)<br />

Figure 5.48. Block diagram <strong>of</strong> speed control loop in discrete domain.<br />

The closed torque control loop transfer function (see for cascade <strong>DTC</strong>-<strong>SVM</strong> -Table 5.2<br />

or parallel <strong>DTC</strong>-<strong>SVM</strong> - Table 5.4) is:<br />

0.466*0.37267 a<br />

DΩ ( z) = GM<br />

_ closed( z)<br />

≅ = (5.86)<br />

2 2<br />

z - 1.289z + 0.4633) z +bz + c<br />

The GΩ ( z)<br />

is discrete transfer function for torque-speed relationship <strong>with</strong> zero order<br />

hold (ZOH). The G ( z)<br />

can be calculated as:<br />

Ω<br />

G () s<br />

1<br />

G z z Z z Z<br />

s<br />

( Js)<br />

s<br />

−1 Ω<br />

−1<br />

Ω( ) = (1 − ) [ ] = (1 − ) [ ]<br />

−1<br />

1 1 1 Ts<br />

= (1 − z ) Z [ ] = J s<br />

2 J ( z −1)<br />

(5.87)<br />

Finally, it can be expressed as:<br />

AΩ<br />

d<br />

GΩ ( z)<br />

=<br />

( z −1)<br />

(5.88)<br />

Ts<br />

Where: A = Ω d<br />

J<br />

, and T<br />

s<br />

is sampling time.<br />

Using the sampling time T = 200µ<br />

s and motor parameters J = 0.0173 the GΩ ( z)<br />

can<br />

be calculated as:<br />

s<br />

G ( z)<br />

Ω<br />

=<br />

0.01156<br />

( z −1)<br />

Hence, the transfer function <strong>of</strong> closed speed control loop can be written:<br />

G<br />

Ω _ closed<br />

Ωm( z) CΩ( z) DΩ( z) GΩ( z)<br />

( z)<br />

= =<br />

Ω ( z) 1 + C ( z) D ( z) G ( z)<br />

m_<br />

ref<br />

Ω Ω Ω<br />

(5.89)<br />

(5.90)<br />

And finally<br />

112


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

G<br />

Ω _ closed<br />

( z)<br />

=<br />

K<br />

pΩ<br />

( K + K )( z−<br />

) aA<br />

K + K<br />

pΩ iΩ Ωd<br />

pΩ<br />

iΩ<br />

2<br />

pΩ<br />

( z−1)( z − bz+ c)( z− 1) + ( KpΩ + KiΩ)( z−<br />

) aAΩ<br />

d<br />

KpΩ<br />

+ KiΩ<br />

K<br />

(5.91)<br />

In many practical cases the digital filter is used in speed measurement loop (see Fig.<br />

5.49).<br />

Ω m _ ref<br />

( z)<br />

PI controller GΩ<br />

( z)<br />

C ( z)<br />

Ω<br />

Me_ ref()<br />

z<br />

D ( z)<br />

Ω<br />

M ( ) e<br />

z<br />

M ( ) L<br />

z<br />

ZOH<br />

<strong>Control</strong> Plant<br />

}<br />

1<br />

Js<br />

Ω ( z m<br />

)<br />

<strong>Torque</strong> control loop<br />

F ( z)<br />

Ω<br />

Digital Filter<br />

Figure 5.49. Block scheme <strong>of</strong> speed control <strong>with</strong> digital filter in speed measurement loop<br />

(discrete domain).<br />

The transfer function FΩ ( s)<br />

<strong>of</strong> first order low pass filter in s domain is expressed as:<br />

FΩ () s =<br />

1<br />

s<br />

+ 1<br />

ω<br />

i<br />

(5.92)<br />

2 ω<br />

Where ωc<br />

= 2π<br />

fc<br />

and f<br />

c<br />

is cut <strong>of</strong>f frequency and tan(<br />

cTs<br />

ω<br />

i<br />

= ) [2]. In practice f<br />

c<br />

T 2<br />

is selected in the range 20-250Hz<br />

2( z −1)<br />

Using the Tutsins’s approximation method s = for discretization process, the<br />

Ts<br />

( z+<br />

1)<br />

discrete transfer function <strong>of</strong> first order low pass filter can be expressed as:<br />

Tsωi<br />

2 −Tsωi<br />

Where: a1<br />

= , b1<br />

=<br />

2 + T ω 2 + T ω<br />

s<br />

i<br />

Tsωi<br />

( z + 1)<br />

2 + T ω a ( z+<br />

1)<br />

FΩ<br />

( z)<br />

= =<br />

z<br />

2 T ω<br />

s<br />

s<br />

s i<br />

1<br />

(5.93)<br />

2 −Tsωi<br />

−<br />

z−<br />

b1<br />

+<br />

i<br />

s<br />

i<br />

113


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

The discrete transfer function <strong>of</strong> digital filter for T = 200µ<br />

s and f = 25Hz<br />

can be<br />

calculated as:<br />

0.015466 (z+1)<br />

FΩ ( z)<br />

= (5.94)<br />

(z-0.9691)<br />

Hence, the transfer function speed control loop <strong>with</strong> digital filter:<br />

s<br />

c<br />

G<br />

Ω _ closed<br />

Ωm( z) CΩ( z) DΩ( z) GΩ( z)<br />

( z)<br />

= =<br />

Ω ( z) 1 + C ( z) D ( z) G ( z) F( c)<br />

m_<br />

ref<br />

Ω Ω Ω<br />

(5.95)<br />

And finally<br />

G<br />

=<br />

Ω _ closed<br />

( z)<br />

=<br />

K<br />

( K + K )( z−b)( z−<br />

) aA<br />

pΩ<br />

pΩ iΩ 1<br />

Ωd<br />

KpΩ<br />

+ KiΩ<br />

K<br />

z bz c z z z b K K z aA a z<br />

2<br />

pΩ<br />

( − + )( −1)( −1)( −<br />

1) + (<br />

pΩ +<br />

iΩ)( − )<br />

Ωd<br />

1( + 1)<br />

KpΩ<br />

+ KiΩ<br />

(5.96)<br />

Selecting<br />

K Ω<br />

, K Ω<br />

will influence poles placement <strong>of</strong> closed speed control loop and as<br />

p<br />

i<br />

a consequence also speed step responses can be selected.<br />

In order to select the best value <strong>of</strong> PI speed controller it is recommended to use the<br />

SISO tools from Matlab package to tune the parameter <strong>of</strong> PI speed controller.<br />

The speed response <strong>with</strong> digital filter simulated in SIMULINK is shown in Fig. 5.50<br />

and simulated in SABER in Fig. 5.51 is presented.<br />

114


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

Figure 5.50. Simulated (SIMULINK) speed response <strong>with</strong> digital filter<br />

Figure 5.51. Simulated (SABER) speed response <strong>with</strong> digital filter in feedback. From the top<br />

reference torque, measured speed.<br />

However, the speed respond is characterized by large overshoot. Therefore, the prefilter<br />

will be applied in order to reduce overshoot (see Fig. 5.52). The discrete transfer<br />

function <strong>of</strong> prefilter Pz ( ) can be expressed as”<br />

KK _ s 0.005<br />

Pz ( ) = =<br />

z−bb_ s z−0.995<br />

(5.97)<br />

where<br />

KK _ s = lim( z − 0.995)=0.005 is gain <strong>of</strong> the prefilter.<br />

z→1<br />

115


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

Ω m _ ref<br />

( z)<br />

P ( z)<br />

Ω<br />

C ( z)<br />

Ω<br />

M<br />

_<br />

() z<br />

e ref<br />

D ( z)<br />

Ω<br />

M ( ) e<br />

z<br />

M ( L<br />

z )<br />

GΩ<br />

( z)<br />

}<br />

ZOH<br />

1<br />

Js<br />

Ω ( z m<br />

)<br />

F ( z)<br />

Ω<br />

Figure 5.52. Speed response <strong>with</strong> digital filter in feedback FΩ ( z)<br />

and prefilter PΩ ( z)<br />

at the<br />

input.<br />

The speed response <strong>with</strong> and <strong>with</strong>out prefilter are shown in Fig. 5.53.<br />

Without prefilter<br />

With prefilter<br />

Figure 5.53. Speed response: blue signal <strong>with</strong>out prefilter and green signal <strong>with</strong> prefilter at the<br />

input.<br />

Design parameters <strong>of</strong> PI speed controller for sampling time T<br />

s<br />

= 50 µ s,<br />

100 µ s,<br />

200 µ s,<br />

400µ s are summarized in Table 5.5.<br />

116


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

117


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

Simulated results for speed tracking performance for different reference speed level in<br />

Fig. 5.54 are shown.<br />

Figure 5.54. Simulated speed tracking performance <strong>with</strong> prefilter at the input for 10%, 20%,<br />

50%, 100% <strong>of</strong> nominal speed. From the top actual speed, reference torque.<br />

Investigation for influence <strong>of</strong> load torque in Fig. 5.55 is presented.<br />

Figure 5.55. Simulated disturbance rejection performance <strong>of</strong> speed control loop for step load<br />

change 50% <strong>of</strong> nominal torque. From the top electromagnetic torque, measured speed<br />

Simulation results for speed control loop in Saber package for sampling time<br />

Ts<br />

= 200µ<br />

sand PI speed parameters controller: K Ω<br />

= 1.1940 , T Ω<br />

= 0.0398 (see Table<br />

5.5) in Fig. 5.56 is shown.<br />

p<br />

i<br />

118


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

Figure 5.56. Simulated speed tracking performance to step <strong>of</strong> speed from 0 to 1000rpm.<br />

The presented simulation results confirm well proper operation and design methodology<br />

for digital speed control loop.<br />

119


<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong><br />

5.5 Summary<br />

This Chapter presents design <strong>of</strong> discrete control loops for two <strong>DTC</strong>-<strong>SVM</strong> schemes:<br />

series (cascade) and parallel structures <strong>of</strong> flux and torque controllers, Fig. 5.2 and<br />

Fig. 5.25, respectively. The cascade structure operates <strong>with</strong> P-flux controller and PItorque<br />

controller whereas in parallel structure two PI controllers are used. In the first<br />

step <strong>of</strong> design calculation <strong>of</strong> discrete Z- transfer function from continuous s- domain<br />

transfer function using zero order hold (ZOH) method <strong>of</strong> discretization has been<br />

performed. The continuous PI controller transfer function has been discretized using<br />

backward difference approach. Secondly, a SISO tool from Matlab package for<br />

digital controller parameter calculation has been applied. The results <strong>of</strong> design were<br />

verified by Simulink (using simplified discrete transfer function) and Saber (using<br />

full motor and inverter model) simulation. Also, the influence <strong>of</strong> sampling time<br />

selection on controller parameters have been discussed. Finally, also the speed<br />

control loop was synthesized using similar methodology.<br />

120


Sensorless Speed <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>)<br />

Chapter 6<br />

DIRECT TORQUE CONTROL WITH SPACE VECTOR MODULATION (<strong>DTC</strong>-<br />

<strong>SVM</strong>) OF PMSM DRIVE WITHOUT MOTION SENSOR<br />

6.1 Introduction<br />

Many motion control applications, such as material handling, packaging and hydraulic<br />

or pneumatic cylinder replacement, require the use <strong>of</strong> a position transducer for speed or<br />

position feedback, such as an encoder or resolver. In addition, permanent magnet<br />

synchronous motors require position feedback to perform commutation. Some <strong>of</strong><br />

systems utilize velocity transducer as well. These sensor add cost, weight, and reduce<br />

the reliability <strong>of</strong> the system. Also, a special mechanical arrangement needs to be made<br />

for mounting the position sensors. An extra signal wires are required from the sensor to<br />

the controller. Additionally, some type <strong>of</strong> position sensors are temperature sensitive and<br />

their accuracy degrades, when the system temperature exceed the limits. Therefore, the<br />

research in the area <strong>of</strong> sensorless speed control <strong>of</strong> PMSM is beneficial because <strong>of</strong> the<br />

elimination <strong>of</strong> the feedback wiring, reduced cost, and improved reliability.<br />

Sensorless speed <strong>DTC</strong>-<strong>SVM</strong> control block scheme is presented in Fig. 6.1.<br />

Figure 6.1. Block scheme <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> for PMSM drive <strong>with</strong>out motion sensor.<br />

As we can see the operation <strong>of</strong> speed controlled PMSM drive <strong>with</strong>out mechanical<br />

motion sensor is based only on measurement <strong>of</strong> following signals, which are available<br />

in every PWM inverter-fed drive system as:<br />

• DC link voltage,<br />

• motor phase currents,<br />

121


Sensorless Speed <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>)<br />

Based on this signals, the state variable <strong>of</strong> the drive can be indirectly calculated or<br />

estimated what further allow to achieve the estimated (actual) rotor speed <strong>of</strong> PMSM.<br />

Two motion sensorless control schemes <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> for PMSM drive are presented in<br />

Fig. 6.2<br />

a)<br />

U DC<br />

Ω m _ ref<br />

Ψ s _ ref<br />

M e _ ref<br />

e Me<br />

∆δ<br />

U<br />

s α _ ref<br />

U<br />

s β _ ref<br />

S<br />

A<br />

S B<br />

S C<br />

θ Ψ s<br />

Ψ<br />

s<br />

I<br />

s<br />

U s<br />

M e<br />

I s<br />

Ω m _ est<br />

b)<br />

U DC<br />

Ω m _ ref<br />

Ψ s _ ref<br />

M e _ ref<br />

e Ψ s<br />

e M e<br />

U sx _ ref<br />

U sy _ ref<br />

U<br />

s α _ ref<br />

U<br />

s β _ ref<br />

S A<br />

S B<br />

S C<br />

Ψ<br />

s<br />

M<br />

e<br />

θ Ψs<br />

U s<br />

I<br />

s<br />

Ω m _ est<br />

Figure 6.2. The <strong>DTC</strong>-<strong>SVM</strong> block schemes <strong>of</strong> PMSM <strong>with</strong>out motion sensor: a) cascade<br />

structure and b) parallel structure.<br />

In motion sensorless PMSM drives, as shown in Fig.6.2, the position or speed<br />

transducer (see Fig. 5.52) is replaced by a speed estimation block, which generates the<br />

speed feedback signal Ω into the control systems and stator flux model.<br />

m_<br />

est<br />

122


Sensorless Speed <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>)<br />

The problem associated <strong>with</strong> speed sensorless operation <strong>of</strong> PMSM drive sourced from<br />

VSI is listed below:<br />

• initial rotor flux detection at start up <strong>of</strong> PMSM controlled drive,<br />

• stator flux estimation <strong>with</strong>out measured speed or position signal from sensor,<br />

• rotor speed estimation based on state variable <strong>of</strong> the PMSM, especially in low<br />

speed operation region.<br />

Therefore, in this Section the initial rotor position detection method <strong>of</strong> permanent<br />

magnets, as well as stator flux and rotor speed estimation techniques will be discussed.<br />

6.2 Initial rotor detection method<br />

In a PMSM drive the detection <strong>of</strong> initial flux position is an important task. The initial<br />

position <strong>of</strong> the rotor must be detected correctly in order to initialize the flux estimation<br />

procedure. In case <strong>of</strong> wrong detection the control algorithm has incorrect information<br />

and the rotor shaft can be rotated through few second in positive or negative direction.<br />

This situation is not acceptable in any drive system. Therefore, for the stable starting <strong>of</strong><br />

PMSM drive <strong>with</strong>out the temporary reversal rotation, the initial rotor position<br />

estimation is proposed.<br />

The simplest method to achieved the initial rotor flux position is based on the following<br />

rule. For short time the stator winding is supplied by the DC voltage. It impress the DC<br />

current, which generates the magnetic field. The permanent magnet <strong>of</strong> PMSM sets<br />

accordance <strong>with</strong> this field line. This position <strong>of</strong> PM flux is used to set initial values for<br />

the stator flux estimation algorithm.<br />

This method is very simple and not complicate. However, has disadvantage that during<br />

this process the rotor can be moved in unknown direction depending on:<br />

• position <strong>of</strong> PM before initial detection procedure,<br />

• direction <strong>of</strong> DC voltage supply into the motor phase.<br />

In order to make the initial rotor flux position correct <strong>with</strong>out any movement (at<br />

standstill) the following algorithms can be used in the literature [78,90,93,95,99,101]<br />

123


Sensorless Speed <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>)<br />

To estimate the initial rotor position before starting <strong>DTC</strong>-<strong>SVM</strong> control, two kind <strong>of</strong> the<br />

rectangular pulsewise voltages are applied from the inverter to the motor:<br />

• one is the short pulsewise voltage,<br />

• and another is the longer one.<br />

Short pulsewise voltage test<br />

This test based on general principle that the three-phase winding inductances <strong>of</strong> PMSM<br />

are a function <strong>of</strong> the mechanical rotor position. Therefore, from the line current<br />

responses in stator oriented coordinates α,<br />

β under the short pulse wise voltage (see<br />

Fig. 6.3) the position <strong>of</strong> PM can be estimated.<br />

Figure 6.3. Voltage pulse wise during short time voltage test.<br />

During the short time (100µ s ) the vector V 1<br />

=(100) and opposite V 4<br />

=(011) is<br />

generated by voltage source inverter. The achieved current responses in α,<br />

β system for<br />

two type <strong>of</strong> PMSM during this test in respect to mechanical rotor position are presented<br />

in Fig. 6.4.<br />

a) b)<br />

Figure 6.4. Current components in stator oriented coordinates α,<br />

β under supplied voltage<br />

vector to the motor for very short time : a) for Ld = Lq<br />

and b) for Lq > Ld.<br />

124


Sensorless Speed <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>)<br />

The stator current trajectory can be divided in to eight sectors (see Fig. 6.5a).<br />

7π<br />

8<br />

o<br />

157 .5<br />

Sector 4<br />

− +<br />

5π<br />

8<br />

o<br />

112 .5<br />

Sector 3<br />

− −<br />

3π<br />

8<br />

+ −<br />

o<br />

67 .5<br />

Sector 2<br />

I s<br />

o<br />

22.5<br />

π<br />

8<br />

+ +<br />

Sector 5<br />

o<br />

202 .5<br />

9π<br />

8<br />

Sector 6<br />

+ −<br />

o<br />

247 .5<br />

11π<br />

8<br />

− −<br />

Sector 7<br />

− +<br />

o<br />

292 .5<br />

13π<br />

8<br />

+ +<br />

Sector 8<br />

Sector 1<br />

15π<br />

8 π<br />

−<br />

8<br />

o<br />

337 .5<br />

Figure 6.5. Stator current trajectory.<br />

Based on the measured response <strong>of</strong> phase currents in α,<br />

β coordinates, the<br />

( ) sign I s<br />

+ and sign( I s<br />

)<br />

− is calculated from following formulas:<br />

sign( I )<br />

+ = I + I − I<br />

(6.1)<br />

s sα<br />

sβ<br />

s<br />

sign( I )<br />

− = I −I − I<br />

(6.2)<br />

s sα<br />

sβ<br />

s<br />

The possible combinations <strong>of</strong> sign( I s<br />

)<br />

+ and sign( I s<br />

)<br />

− are shown in Fig. 6.6:<br />

Figure 6.6. Possible combination <strong>of</strong> sign( I s<br />

)<br />

+<br />

sign( I s<br />

) + + − −<br />

−<br />

sign( I s<br />

) + − − +<br />

+ and sign( I s<br />

)<br />

− under short pulse supply.<br />

Let us assuming, for example, the case where the sign( I s<br />

)<br />

+ and sign( I s<br />

)<br />

− have positive<br />

π π<br />

sign. The position γ m<br />

exist in the domain <strong>of</strong> − ~ or 7 π 9 π<br />

~ and two estimated<br />

8 8 8 8<br />

position can be obtained.<br />

The mathematical analysis <strong>of</strong> I , I<br />

sα<br />

sβ waveforms leads to following equations:<br />

I = I +∆ I cos 2γ<br />

(6.3)<br />

sα<br />

s s m<br />

125


Sensorless Speed <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>)<br />

I<br />

=∆ I sin 2γ<br />

(6.4)<br />

sβ<br />

s m<br />

where I<br />

s<br />

is DC component in Is<br />

α<br />

and<br />

∆ Is<br />

amplitude <strong>of</strong> fluctuated component.<br />

The current components in α,<br />

β can be modeled as an average value <strong>of</strong> I<br />

s<br />

plus some<br />

<strong>of</strong>fset value<br />

∆<br />

Is<br />

, as a function <strong>of</strong> the mechanical position γ m<br />

. Fig. 6.4 shows the<br />

current components given in Eq. (6.1-2), which are function <strong>of</strong> the phase angle 2γ m<br />

.<br />

Solving those Eqs. (6.1 and 6.2) in respect to mechanical rotor position, two domains <strong>of</strong><br />

mechanical rotor position can be obtained as:<br />

Isβ<br />

γ<br />

m1<br />

=<br />

(6.5)<br />

2( I − I )<br />

sα<br />

s<br />

= Isβ<br />

γ<br />

m2<br />

2( I − I )<br />

+ π (6.6)<br />

sα<br />

s<br />

The estimated rotor positions for other combination <strong>of</strong> sign( I s<br />

)<br />

summarized in Table 6.7.<br />

+ and sign( I s<br />

)<br />

− are<br />

15π<br />

π<br />

−<br />

8 8<br />

7π<br />

9π<br />

−<br />

8 8<br />

π 3π<br />

−<br />

8 8<br />

9π<br />

11π<br />

−<br />

8 8<br />

3π<br />

5π<br />

−<br />

8 8<br />

11π<br />

13π<br />

−<br />

8 8<br />

5π<br />

7π<br />

−<br />

8 8<br />

13π<br />

15π<br />

−<br />

8 8<br />

I sα − I s + I sβ<br />

+ +<br />

+ −<br />

−<br />

−<br />

−<br />

I sα<br />

− I s − I sβ<br />

γ m<br />

+<br />

I sβ<br />

2(<br />

I sα<br />

− I s )<br />

I sβ<br />

+ π<br />

2(<br />

I sα<br />

− I s )<br />

I sα<br />

+ I s π<br />

− +<br />

2I<br />

sβ<br />

4<br />

I sα<br />

+ I s 5π<br />

− +<br />

2I<br />

sβ<br />

4<br />

I sβ<br />

π<br />

+<br />

2( I s α − I s ) 2<br />

I sβ<br />

3π<br />

+<br />

2( I s α − I s ) 2<br />

I sα<br />

+ I s 3π<br />

− +<br />

2I<br />

sβ<br />

4<br />

I sα<br />

+ I s 7π<br />

− +<br />

2I<br />

sβ<br />

4<br />

Table 6.7. Mechanical rotor position calculations.<br />

Long pulsewise voltage test<br />

This test help us to choose the proper estimated value <strong>of</strong> mechanical rotor position from<br />

two values calculated during the short pulse wise voltage test.<br />

126


Sensorless Speed <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>)<br />

This test is based on the saturation effect <strong>of</strong> magnetic circuit. If the long pulse will send<br />

in direction <strong>of</strong> north pole <strong>of</strong> PM, the current response will be slower, and, if current<br />

response will be faster that the previous, it means that it is south pole.<br />

6.3 Stator flux estimation methods<br />

Flux estimation is an important task in implementation <strong>of</strong> high-performance <strong>DTC</strong>-<strong>SVM</strong><br />

motor drives. <strong>Vector</strong> control method <strong>of</strong> PMSM drive needs knowledge about actual<br />

value <strong>of</strong> the stator flux magnitude and position as well electromagnetic torque. Also, the<br />

flux estimation is needed to calculate the actual rotor speed for sensorless operation.<br />

6.3.1 Overview<br />

Many different technique has been developed for PMSM flux estimation [107].<br />

Generally, they may be divided into two groups: open loop estimators and closed loop<br />

estimators/observers. Most <strong>of</strong> these method are based on so called “current model” or<br />

“voltage model” [110,113]. In fact closed loop estimators/observers are based on the<br />

current or voltage model <strong>with</strong> an error correction loop, which drives error between two<br />

flux models to zero in steady state. However, an observer has its own dynamics, is<br />

sensitive to parameter changes, and has to be carefully designed for individual drives.<br />

Therefore, for commercially manufactured drives is to complicated and impractical.<br />

This is the reason why in this Chapter only open loop flux estimators will be<br />

considered.<br />

6.3.2 Current model based flux estimator<br />

The block scheme <strong>of</strong> the current based flux model is presented in Fig. 6.7. It requires:<br />

• knowledge <strong>of</strong> PMSM machine inductance L , L ,<br />

• speed or position signal,<br />

• PMSM phase currents.<br />

d<br />

q<br />

This kind <strong>of</strong> flux estimator served in experimental test as a master (standard) to run the<br />

<strong>DTC</strong>-<strong>SVM</strong> scheme <strong>with</strong> speed sensor.<br />

127


Sensorless Speed <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>)<br />

Ψ PM<br />

I sA<br />

ABC<br />

Is<br />

α<br />

αβ<br />

Isd<br />

L d<br />

Ψ sd<br />

dq<br />

Ψ sα<br />

Ψ s<br />

I sB<br />

αβ<br />

Is<br />

β<br />

dq<br />

I sq<br />

L q<br />

Ψ sq<br />

αβ<br />

Ψ sβ<br />

θ Ψs<br />

γ m<br />

Figure 6.7. Current model based stator flux estimator.<br />

6.3.3 Voltage model based flux estimator <strong>with</strong> ideal integrator<br />

The stator flux linkage can be obtained by using terminal voltages and currents. It is the<br />

integral <strong>of</strong> terminal voltages minus the resistance voltage drop:<br />

dΨ sα<br />

= ( Us α − RsIs<br />

α<br />

)<br />

(6.7)<br />

dt<br />

dΨ sβ<br />

= ( Usβ<br />

− RsIsβ<br />

)<br />

(6.8)<br />

dt<br />

However, at low speed (frequencies) some problems arise, when this technique is<br />

applied, since the stator voltage becomes very small and the resistive voltage drops<br />

become dominant, requiring very accurate knowledge <strong>of</strong> the stator resistance R s<br />

and<br />

very accurate integration. The stator resistance can vary due to temperature changes.<br />

This effect can also be taken into consideration by using the thermal model <strong>of</strong> the<br />

machine. Drifts and <strong>of</strong>fsets can greatly influence the precision <strong>of</strong> integration. The<br />

overall accuracy <strong>of</strong> the estimated flux linkage vector will also depend on the accuracy<br />

<strong>of</strong> the monitored voltages and currents.<br />

The most know classical voltage model obtains the flux components in stator<br />

coordinates ( α,<br />

β ) by integrating the motor back electromotive force E , E<br />

sα<br />

sβ (see Fig.<br />

6.8). The method is sensitive for only one motor parameter, stator resistance R s<br />

.<br />

However, the application <strong>of</strong> pure integrator is difficult because <strong>of</strong> dc drift and initial<br />

value problems.<br />

128


Sensorless Speed <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>)<br />

I sA<br />

ABC<br />

Is<br />

α<br />

I sB<br />

αβ<br />

Is<br />

β<br />

R s<br />

U sA<br />

U sB<br />

ABC<br />

αβ<br />

U<br />

s α<br />

U<br />

s β<br />

R s<br />

Es<br />

α<br />

Es<br />

β<br />

∫<br />

∫<br />

Ψ sα<br />

Ψ sβ<br />

Ψ s<br />

θ Ψs<br />

Figure 6.8. Voltage model based estimator <strong>with</strong> ideal integrator.<br />

There are proposed many improvements <strong>of</strong> the classical voltage model. Some <strong>of</strong> them<br />

are presented bellow.<br />

6.3.4 Voltage model based flux estimator <strong>with</strong> low pas filter<br />

A common way to improve the stator flux voltage based model is to use a first-order<br />

low-pass filter (LP) instead <strong>of</strong> the pure integrator. The equations (6.7 and 6.8) are<br />

transferred to the form:<br />

dΨ sα<br />

= ( Us α − RsIs α)<br />

+ Fc Ψ<br />

sα<br />

(6.9)<br />

dt<br />

dΨ sβ<br />

= ( Usβ − RsIsβ)<br />

+ Fc Ψ<br />

sβ<br />

(6.10)<br />

dt<br />

The block diagram <strong>of</strong> the estimator is presented in Fig. 6.9. Discrete time<br />

implementation <strong>of</strong> the integrator becomes:<br />

zΨ ( z) =Ψ ( z) + ( U − R I ) T<br />

(6.11)<br />

sα sα sα s sα<br />

s<br />

zΨ ( z) =Ψ ( z) + ( U − R I ) T<br />

(6.12)<br />

sβ sβ sβ s sβ<br />

s<br />

A LP filter does not give high accuracy at frequencies lower than cut<strong>of</strong>f frequency<br />

ω = 2π F . There will be errors both in the magnitude and in the phase angle. As<br />

c<br />

c<br />

results, the proposed voltage estimator <strong>with</strong> LP filter can be used successfully only in a<br />

limited speed range above cut<strong>of</strong>f frequency<br />

129


Sensorless Speed <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>)<br />

I sA<br />

ABC<br />

Is<br />

α<br />

I sB<br />

αβ<br />

Is<br />

β<br />

R s<br />

F c<br />

U sA<br />

U sB<br />

ABC<br />

αβ<br />

U<br />

s α<br />

U<br />

s β<br />

-<br />

-<br />

R s<br />

Es<br />

α<br />

Es<br />

β<br />

-<br />

-<br />

∫<br />

∫<br />

Ψ sα<br />

Ψ sβ<br />

Cartesian<br />

To<br />

Polar<br />

Ψ s<br />

θ Ψs<br />

Stator flux estimator (improved voltage model)<br />

F c<br />

Figure 6.9. Voltage model based estimator <strong>with</strong> low-pass filter.<br />

Discrete time implementation <strong>of</strong> the LP filter becomes:<br />

zΨ ( z) =Ψ ( z) + ( U − R I ) T + FΨ ( z)<br />

(6.13)<br />

sα sα sα s sα s c sα<br />

zΨ ( z) =Ψ ( z) + ( U −R I ) T −FΨ ( z)<br />

(6.14)<br />

sβ sβ sβ s sβ s c sα<br />

6.3.5 Improved voltage model based flux estimator<br />

Many other methods were developed in order to eliminate dc-<strong>of</strong>fset and initial values<br />

problems [107]. In general, the output Y <strong>of</strong> these new integrators (Fig. 6.10) is<br />

expressed as:<br />

ωc<br />

1<br />

Y = X + Y<br />

s+ ω s+<br />

ω<br />

c<br />

c<br />

lim<br />

(6.15)<br />

Where X is the input and Y is output <strong>of</strong> the integrator respectively. The Y lim<br />

is a<br />

compensation signal used as a feedback and ω<br />

c<br />

is cut<strong>of</strong>f frequency.<br />

X<br />

1<br />

s + ω c<br />

Y<br />

Compensation<br />

signal<br />

ωc<br />

s +ω<br />

c<br />

Y lim<br />

− lim<br />

lim<br />

Figure 6.10. Improved integration method <strong>with</strong> saturation block.<br />

130


Sensorless Speed <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>)<br />

The first part <strong>of</strong> the equation represents a LP filter. The second part realizes a feedback,<br />

which is used to compensate the error in the output. The block diagram <strong>of</strong> new<br />

integration algorithm <strong>with</strong> saturation block is shown in see Fig. 6.11.<br />

I sA<br />

ABC<br />

Is<br />

α<br />

I sB<br />

αβ<br />

Is<br />

β<br />

R s<br />

ωc<br />

s +ω<br />

c<br />

Y lim<br />

− lim<br />

lim<br />

U sA<br />

U sB<br />

ABC<br />

αβ<br />

U<br />

s α<br />

U<br />

s β<br />

R s<br />

Es<br />

α<br />

Es<br />

β<br />

1<br />

s + ω c<br />

1<br />

s + ω c<br />

Ψ α _ comp<br />

Ψ s α<br />

Ψ sβ<br />

Ψ s<br />

θ Ψs<br />

Ψ<br />

s β _ comp<br />

ωc<br />

s +ω<br />

c<br />

Y lim<br />

− lim<br />

lim<br />

Figure 6.11. Full block diagram <strong>of</strong> voltage model based estimator <strong>with</strong> saturation block on the<br />

α,<br />

β components.<br />

The main task <strong>of</strong> saturation block is to stop the integration when the output signal<br />

Ψ<br />

sα<br />

or<br />

Ψ<br />

sβ<br />

exceed the reference value <strong>of</strong> stator flux amplitude. Please note that if the<br />

compensation signal is set to zero, the improved integrator represents a first-order LPfilter.<br />

If the compensation signal<br />

integrator operates as a pure integrator.<br />

Ψ or Ψ<br />

β _<br />

is not zero the improved<br />

sα<br />

_ comp<br />

s<br />

comp<br />

Discrete time implementation <strong>of</strong> the improved integrator becomes:<br />

ω<br />

zΨ ( z) =Ψ ( z) + ( U − R I ) T + ( Ψ ( z) −Ψ ( z))<br />

(6.16)<br />

c<br />

sα sα sα s sα s sα _lim<br />

sα<br />

Ts<br />

ω<br />

zΨ ( z) =Ψ ( z) + ( U − R I ) T + ( Ψ ( z) −Ψ ( z))<br />

(6.17)<br />

c<br />

sβ sβ sβ s sβ s sβ _lim<br />

sβ<br />

Ts<br />

The output <strong>of</strong> saturation block can be described as:<br />

⎧Ψ<br />

sαβ<br />

( z) if (Ψ<br />

sαβ(z))=lim<br />

sαβ<br />

131


Sensorless Speed <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>)<br />

Where lim is the limited value. Please note that lim should be set at reference stator<br />

flux amplitude<br />

lim =Ψ<br />

s _ ref<br />

equal Ψ<br />

PM<br />

.<br />

From Eq. 6.18 it can be observed that when one <strong>of</strong> the stator flux linkage components<br />

Ψ<br />

sα<br />

or Ψ<br />

sβ<br />

exceeds the limit, it causes in distortion <strong>of</strong> the output waveform.<br />

6.4 Electromagnetic torque estimation<br />

The PMSM motor output torque is calculated based on the equations (2.51), (2.52),<br />

(2.55), (2.58) presented in Chapter 2, which for stator oriented coordinate system can be<br />

written as follows:<br />

3<br />

M<br />

e<br />

= pb( Ψs αIsβ −Ψ<br />

sβIsα<br />

)<br />

(6.19)<br />

2<br />

It can be seen that calculated torque is dependent on the current measurement accuracy<br />

and stator flux estimation method. In practice current measurement is performed <strong>with</strong><br />

high accuracy ( ≤ 1% <strong>with</strong> 150kHz frequency bandwidth) using, for example, LEM<br />

sensors.<br />

6.5 Rotor speed estimation methods<br />

6.5.1 Overview<br />

High performance operation <strong>of</strong> motion sensorless PMSM drives depends mainly on<br />

accurate knowledge <strong>of</strong> rotor PM flux magnitude, position and speed. The rotor position<br />

estimation methods can be classified into two major groups:<br />

• motor model based,<br />

• rotor saliency based techniques.<br />

The rotor saliency based approach is suitable only for the Interior PMSM (see Fig. 1.2 c<br />

and d). Motor model based approach detect the back EMF vector, which includes<br />

information about position and speed, using either open loop models/estimators<br />

[81,85,86] or closed loop estimators/observers [70, 73,74,96,97,100]. Also adaptive<br />

observers [72,92,98,83] and Extended Kalman Filters (EKF) [67,73] have been<br />

proposed for motor position and speed estimation. These methods, however, are<br />

computationally intensive and require careful design and proper initialization.<br />

132


Sensorless Speed <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>)<br />

Therefore, for commercially manufactured drives are impractical and further a simple<br />

open loop based techniques will be considered.<br />

6.5.2 Back electromotive force (BEMF) technique<br />

This technique uses the back electromotive force to estimate the rotor speed [70]. The<br />

velocity signal could be integrated to generate a position estimate. However, this signal<br />

is sensitive to parameter variations and tends to drift and have <strong>of</strong>fset problem. Another<br />

problem <strong>with</strong> using BEMF to estimate position is that at zero speed the BEMF goes to<br />

zero and at low speed the signal to noise ratio can not be ignored.<br />

6.5.3 Stator flux based technique<br />

Generally, the calculation <strong>of</strong> rotor speed is based on the simple relationship:<br />

θ = θ − δ , (6.20)<br />

r Ψs Ψs<br />

where θ<br />

r<br />

is electrical position, θ Ψ s<br />

is stator flux position and δ Ψ s<br />

is torque angle.<br />

After differentiation equation (6.20) and taking into account that θ r<br />

= p b<br />

γ m<br />

the<br />

mechanical speed <strong>of</strong> PMSM rotor can be expressed as:<br />

⎛dθ<br />

⎜<br />

⎝ dt<br />

Ψs<br />

Ω<br />

m<br />

= −<br />

dδΨ<br />

⎞<br />

⎟/<br />

pb<br />

, (6.21)<br />

dt ⎠<br />

dθ s<br />

where Ω Ψ<br />

Ψs<br />

= is angular speed <strong>of</strong> stator flux vector and δ Ψ<br />

is torque angle.<br />

dt<br />

As we can observe form equation (6.21) in order to calculate the mechanical rotor speed<br />

it is necessary to calculate separately two components. One <strong>of</strong> them is angular speed <strong>of</strong><br />

stator flux vector Ω<br />

Ψs<br />

and the second one is change <strong>of</strong> the load angle d δ Ψ (see Fig.<br />

dt<br />

6.12).<br />

133


Sensorless Speed <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>)<br />

Is<br />

α<br />

Is<br />

β<br />

U<br />

s α<br />

Ψ<br />

sα<br />

Ψ sβ<br />

Ω =<br />

Ψs<br />

d<br />

dt<br />

θ Ψ s<br />

dδ Ψ<br />

dt<br />

−<br />

1<br />

p b<br />

Ω m<br />

U<br />

s β<br />

Figure 6.12. Block diagram <strong>of</strong> stator flux vector based rotor speed estimator.<br />

The synchronous speed<br />

Ω<br />

Ψs<br />

is calculated based on the stator flux estimator:<br />

θ Ψ s<br />

Ψ<br />

sβ<br />

= arctg( )<br />

Ψ<br />

sα<br />

(6.22)<br />

and the calculation <strong>of</strong><br />

ΩΨs<br />

can be done as:<br />

The estimation <strong>of</strong> the synchronous speed<br />

dθ θ<br />

s Ψs( k) −θ<br />

Ψ<br />

Ψs( k−1)<br />

Ω<br />

Ψs<br />

= = (6.23)<br />

dt T<br />

s<br />

Ω<br />

Ψs<br />

based on the derivative <strong>of</strong> the position <strong>of</strong><br />

stator flux space vector can be modified taking in to account equation (6.22), which<br />

finally gives [12]:<br />

Ψ<br />

Ω =<br />

Ψs<br />

sα<br />

dΨ<br />

sβ<br />

−Ψ<br />

dt<br />

Ψ +Ψ<br />

sβ<br />

2 2<br />

sα<br />

sβ<br />

dΨ<br />

dt<br />

sα<br />

(6.24)<br />

Digital implementation <strong>of</strong> equation (6.24) can be written as:<br />

Ψ Ψ −Ψ Ψ<br />

Ω =<br />

Ψ T<br />

sα( k−1) sβ( k) sβ( k−1) sα( k)<br />

Ψs( k) 2<br />

s<br />

s<br />

(6.25)<br />

Also from equation (2.27b) in stator flux coordinate system the synchronous speed<br />

can be obtained as:<br />

Ω<br />

Ψs<br />

U − R I E<br />

Ω<br />

Ψs<br />

= =<br />

Ψ Ψ<br />

sy s sy sy<br />

s<br />

s<br />

(6.26)<br />

134


Sensorless Speed <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>)<br />

Calculation <strong>of</strong> d δ Ψ<br />

dt<br />

Based on the flux-current (Eq. 2.29b) and torque (Eq. 2.58) equations in stator flux<br />

coordinates under consideration that for surface PMSM machine L d<br />

= L q<br />

and making<br />

assumption that for small changes <strong>of</strong> torque angleδ Ψ<br />

, the sinδ<br />

be written as:<br />

Ψ<br />

= δ , the equations can<br />

Ψ<br />

0= LI −Ψ sin<br />

(6.27)<br />

s sy PM<br />

δ Ψ<br />

3<br />

M<br />

e<br />

= pbΨ PMIsy<br />

(6.28)<br />

2<br />

the torque angle δ Ψ<br />

can be calculated as:<br />

LI<br />

s sy 2ML<br />

e s<br />

δ Ψ<br />

= =<br />

Ψ 3p<br />

Ψ Ψ<br />

PM b s PM<br />

(6.29)<br />

Further, the d dt<br />

δ Ψ is calculated as:<br />

( ) ( 1)<br />

δΨ<br />

Ψ k Ψ k−<br />

d<br />

dt<br />

δ −δ<br />

= (6.30)<br />

T<br />

s<br />

Figure 6.13. Simulated oscillograms <strong>of</strong> rotor speed estimation according to block scheme from<br />

Fig. 6.12 (in Saber). From the top: synchronous speed Ω<br />

Ψs<br />

, the d signal, the measured and<br />

dt<br />

estimated rotor speed, the measured and estimated electromagnetic torque.<br />

δ Ψ<br />

135


Sensorless Speed <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>)<br />

The speed estimation problem is still open, especially at low and zero speed operations.<br />

The accuracy <strong>of</strong> presented method depends on accuracy <strong>of</strong> applied stator flux estimation<br />

and differentiation algorithm. It allows, however, for robust start, closed loop operation<br />

above 10% <strong>of</strong> nominal speed, and braking the drive to zero speed.<br />

6.5 Summary<br />

The main problems associated <strong>with</strong> PMSM sensorless speed operation are presented<br />

in this Chapter. For robust start <strong>of</strong> PMSM <strong>with</strong>out the temporary rotor reversal a<br />

special initialization algorithm has been used. This algorithm performs two test:<br />

short and the longer voltage generated by the PWM inverter. The used speed<br />

estimation algorithm is based on stator flux vector and torque angle estimation and<br />

does not operate accurately around zero speed region. However, it allows robust<br />

start and closed speed operation in the speed range above 10% <strong>of</strong> nominal speed.<br />

The effectiveness <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> <strong>with</strong> and <strong>with</strong>out motion sensor has been proved by<br />

simulation and experimental results (see Chapter 7)<br />

136


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

Chapter 7 DSP IMPLEMENTATION OF <strong>DTC</strong>-<strong>SVM</strong> CONTROL<br />

7.1 Description <strong>of</strong> the laboratory test-stand<br />

The basic structure <strong>of</strong> laboratory setup is presented in Fig. 7.1 and the photo <strong>of</strong><br />

laboratory setup is shown in Fig. 7.2. The motor setup consists <strong>of</strong> 3kW permanent<br />

magnet synchronous motor and DC motor, which is used as a load. The PMSM machine<br />

is supplied by PWM inverter, which is controlled by digital signal processor (DSP)<br />

based on DS1103 board. The voltage inverter is supplied from three-phase diodes<br />

rectifier. The DSP interface is used in order to separate the high power from the low<br />

power circuit (computer part). Please note that the DS1103 is inserted inside the PC<br />

computer.<br />

Figure 7.1. Block scheme <strong>of</strong> laboratory setup.<br />

Figure 7.2. Laboratory setup. 1-voltage inverter, 2-control for DC motor, 3- PMSM machine, 4<br />

– DSP interface.<br />

137


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

The detailed power circuit <strong>of</strong> the laboratory setup is shown in Fig. 7.3.<br />

Rectifier<br />

Inverter<br />

three-phase<br />

supply<br />

network<br />

U A<br />

U B<br />

U C<br />

D 1<br />

D3<br />

D5<br />

CF<br />

T 1<br />

D 1<br />

T 3<br />

D 3<br />

T 5<br />

D 5<br />

current<br />

sensors<br />

A<br />

B<br />

C<br />

PMSM<br />

D 2<br />

D 4<br />

D 6<br />

T 2<br />

D 2<br />

T 4<br />

D 4<br />

T 6<br />

D 6<br />

speed or position<br />

sensor<br />

SA<br />

SA<br />

SB<br />

SB<br />

SC<br />

SC<br />

Reference speed<br />

<strong>DTC</strong>-<strong>SVM</strong><br />

DS1103<br />

microprocessor<br />

Figure 7.3. Power circuit <strong>of</strong> the laboratory setup.<br />

In presented system the actual two currents and DC link voltage are measured by LEM<br />

sensors and processed by A/D converter. The rotor position and speed are obtained <strong>with</strong><br />

an encoder <strong>of</strong> 2500 pulse per revolution. All internal data <strong>of</strong> DSP can be sent through a<br />

D/A converter and displayed in the scope. All data <strong>of</strong> the PMSM and inverter are given<br />

in the Appendices.<br />

The control algorithm for PMSM machine was written in C language and was<br />

implemented inside the processor. Also, a simple dead-time compensation method and<br />

voltage drop on the semiconductor elements are implemented.<br />

The phase voltage <strong>of</strong> the motor are reconstructed inside the processor using the<br />

measured DC-link voltage and duty cycles <strong>of</strong> PWM for each phases. Motor and PI<br />

controller model are given in Appendices.<br />

Various tests have been carried out in order to investigate the drive performance and to<br />

characterize the steady-state and transient behavior.<br />

138


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

<strong>Control</strong> Desk experiment s<strong>of</strong>tware run on the PC computer provides all function for<br />

controlling, monitoring and automation <strong>of</strong> real-time experiments and makes the<br />

development <strong>of</strong> controllers more effective. A <strong>Control</strong> Desk experiment layout for<br />

control the PMSM machine using <strong>DTC</strong>-<strong>SVM</strong> control method is presented in Fig. 7.4.<br />

Figure 7.4. Performed <strong>Control</strong> Desk experimental layout for control <strong>of</strong> PMSM drive.<br />

139


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

7.2 Steady state behaviour<br />

The experimental steady state no load operation at 25Hz stator frequency for<br />

conventional ST-<strong>DTC</strong> (Fig. 4.10) and <strong>DTC</strong>-<strong>SVM</strong> (Fig. 5.46b) control is presented in<br />

Fig. 7.5. The sampling time has been set at T = 200µ<br />

s for <strong>DTC</strong>-<strong>SVM</strong> and T = 25µ<br />

s for<br />

hysteresis based ST-<strong>DTC</strong> method, respectively.<br />

a)<br />

s<br />

s<br />

b)<br />

Figure 7.5. No load experimental steady state oscillograms at stator frequency 25Hz.<br />

(a) ST-<strong>DTC</strong> for T = 25µ<br />

s (b) <strong>DTC</strong>-<strong>SVM</strong> for T = 200µ<br />

s.<br />

s<br />

From the top: line to line voltage, phase current, amplitude <strong>of</strong> stator flux, motor torque.<br />

s<br />

As we can observed from Fig. 7.5a the motor phase current characterized by high<br />

current ripples. This is mainly because the inductances <strong>of</strong> the PMSM is smaller than an<br />

equivalent power induction motor IM. In order to reduce the current ripples the<br />

140


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

sampling time <strong>of</strong> microcontroller should be decrease. However there is hardware<br />

limitation. The loaded program to microprocessor can not run faster.<br />

Using the space vector modulation (<strong>SVM</strong>) based <strong>DTC</strong> much better results can be<br />

obtained. Note, that in spite <strong>of</strong> lower switching frequency <strong>DTC</strong>-<strong>SVM</strong> guarantees less<br />

current and torque ripple. This is mainly because contrary to hysteresis operation <strong>with</strong><br />

<strong>SVM</strong> operation, the inverter output voltage is unipolar (compare output voltage<br />

waveform in Fig. 7.5a <strong>with</strong> Fig. 7.5b). This also reduces semiconductor device voltage<br />

stress and instantaneous current reversal in DC link.<br />

The presented experimental results (Fig. 7.6-7.9) are measured in the system <strong>with</strong><br />

measured speed taken to the feedback. These investigations have been performed to<br />

show the behaviour <strong>of</strong> the <strong>DTC</strong>-<strong>SVM</strong> system <strong>with</strong>out influence <strong>of</strong> the speed estimation.<br />

In Fig. 7.6 and Fig. 7.7 steady state operation for different values <strong>of</strong> the mechanical<br />

speed and load torque are shown.<br />

Figure 7.6. Experimental steady state operation <strong>of</strong> PMSM controlled via <strong>DTC</strong>-<strong>SVM</strong> <strong>with</strong> the<br />

encoder speed signal taken to the feedback ( Ω = 300rpm<br />

, M = 0 ).<br />

m<br />

From the top: 1) amplitude <strong>of</strong> stator flux (0.25Wb/div), 2) electromagnetic torque (10Nm/div),<br />

3) line to line voltage (1000V/div), 4) stator phase current (10A/div).<br />

l<br />

141


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

Figure 7.7. Experimental steady state operation <strong>of</strong> PMSM controlled via <strong>DTC</strong>-<strong>SVM</strong> <strong>with</strong> the<br />

encoder speed signal taken to the feedback ( Ω = 300rpm<br />

, M = 10Nm<br />

-50% <strong>of</strong> nominal<br />

torque). From the top: 1) amplitude <strong>of</strong> stator flux (0.25Wb/div), 2) electromagnetic torque<br />

(10Nm/div), 3) line to line voltage (1000V/div), 4) stator phase current (10A/div).<br />

m<br />

l<br />

Figure 7.8. Experimental steady state operation <strong>of</strong> PMSM controlled via <strong>DTC</strong>-<strong>SVM</strong> <strong>with</strong> the<br />

encoder speed signal taken to the feedback ( Ω = 1500rpm<br />

, M = 0 ). From the top: 1)<br />

amplitude <strong>of</strong> stator flux (0.25Wb/div), 2) electromagnetic torque (10Nm/div), 3) line to line<br />

voltage (1000V/div), 4) stator phase current (10A/div).<br />

m<br />

l<br />

142


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

Figure 7.9. Experimental steady state operation <strong>of</strong> PMSM controlled via <strong>DTC</strong>-<strong>SVM</strong> <strong>with</strong> the<br />

encoder speed signal taken to the feedback ( Ω = 1500rpm<br />

, M = 10Nm<br />

-50% <strong>of</strong> nominal<br />

torque). From the top: 1) amplitude <strong>of</strong> stator flux (0.25Wb/div), 2) electromagnetic torque<br />

(10Nm/div), 3) line to line voltage (1000V/div), 4) stator phase current (10A/div).<br />

m<br />

l<br />

7.3 Dynamic behaviour<br />

The experimental results <strong>of</strong> flux and torque control loop obtained in dynamic states for<br />

PMSM machine controlled via two different <strong>DTC</strong>-<strong>SVM</strong> schemes are presented.<br />

7.3.1 Flux and torque control loop<br />

Cascade <strong>DTC</strong>–<strong>SVM</strong> control scheme (Fig. 5.46a)<br />

In order to show behaviour <strong>of</strong> the system the dynamic testes for the flux and torque<br />

controllers has been carried out for sampling time, T = 200µ<br />

s. It corresponds to<br />

switching frequency f = 5kHz<br />

. Please note that the flux digital controller parameters<br />

s<br />

were selected according to Table 5.1 and the torque digital controller parameters were<br />

selected from Table 5.2. (see Chapter 5.2).<br />

In Fig. 7.10 stator flux tracking performance is presented. This result is comparable<br />

<strong>with</strong> simulation results presented in Fig. 5.15.<br />

s<br />

143


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

Figure 7.10. Experimental flux tracking performance <strong>of</strong> PM synchronous motor for zero speed<br />

at sampling time Ts<br />

= 200µ<br />

s. Reference flux from70% to 100% <strong>of</strong> nominal value . From the<br />

top: 1-reference stator flux amplitude(0.05Wb/div), 2- stator flux amplitude (0.05Wb/div), 3-<br />

electromagnetic torque (2Nm/div), 4- motor phase current (10A/div).<br />

In Fig. 7.11 torque tracking performance is presented. The achieved result is<br />

comparable <strong>with</strong> simulation results presented in Fig. 5.21b.<br />

Figure 7.11. Experimental torque tracking performance <strong>of</strong> PM synchronous motor for zero<br />

speed at sampling time Ts<br />

= 200µ<br />

s. Reference torque from 0 to nominal value. From the top:1-<br />

reference torque (10Nm/div), 2 - estimated torque (10Nm/div), 3 - stator flux amplitude<br />

(0.1Wb/div), 4- motor phase current (10A/div)<br />

144


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

Figure 7.12. Experimental torque tracking performance <strong>of</strong> PM synchronous motor for zero<br />

speed (zoom) at sampling time Ts<br />

= 200µ<br />

s. Reference torque from 0 to nominal value. From<br />

the top:1- reference torque (10Nm/div), 2 - estimated torque (10Nm/div), 3 - stator flux<br />

amplitude (0.1Wb/div), 4- motor phase current (10A/div).<br />

All experimental results presented in Fig. 7.10-7.12 confirm very well proper and stable<br />

operation <strong>of</strong> flux and torque control loops for cascade <strong>DTC</strong>-<strong>SVM</strong> structure.<br />

145


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

Influence <strong>of</strong> sampling time for torque control loop in cascade <strong>DTC</strong>-<strong>SVM</strong><br />

The influence <strong>of</strong> sampling time on experimental torque tracking performance is<br />

illustrated in Fig. 7.13-7.15. The dynamic test has been carried out for the same<br />

condition ( Ω = 0rpm<br />

) as for simulation shown in Fig. 5.24. The controller parameters<br />

m<br />

has been set according to Table 5.2. In all oscilograms we may see proper operation <strong>of</strong><br />

the torque control loop for different sampling time used in practice.<br />

Figure 7.13. Experimental torque tracking performance <strong>of</strong> PM synchronous motor for zero<br />

speed at sampling time Ts<br />

= 100µ<br />

s( fs<br />

= 10kHz<br />

). Reference torque from 0 to nominal value.<br />

From the top:1- reference torque (4Nm/div), 2 - estimated torque (4Nm/div)<br />

Figure 7.14. Experimental torque tracking performance <strong>of</strong> PM synchronous motor for zero<br />

speed at sampling time Ts<br />

= 200µ<br />

s( fs<br />

= 5kHz<br />

). Reference torque from 0 to nominal value.<br />

From the top:1- reference torque (4Nm/div), 2 - estimated torque (4Nm/div)<br />

146


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

Figure 7.15. Experimental torque tracking performance <strong>of</strong> PM synchronous motor for zero<br />

speed at sampling time T = 400µ<br />

s( f = 2.5kHz<br />

). Reference torque from 0 to nominal value.<br />

s<br />

From the top:1- reference torque (4Nm/div), 2 - estimated torque (4Nm/div).<br />

s<br />

Parallel structure <strong>of</strong> <strong>DTC</strong>–<strong>SVM</strong> scheme (Fig. 5.46b)<br />

Dynamic testes for the flux and torque controllers were carried out for sampling time<br />

Ts<br />

= 200µ<br />

s, which corresponds to switching frequency f = 5kHz<br />

. Please not that the<br />

digital flux controller parameters were selected according to Table 5.3 and the digital<br />

torque controller parameters were selected from Table 5.4. (see Chapter 5.3).<br />

In Fig. 7.16 stator flux tracking performance is presented. This result is comparable<br />

<strong>with</strong> simulation results presented in Fig. 5.33 for flux and Fig. 5.45 for torque loop,<br />

respectively.<br />

s<br />

147


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

Figure 7.16. Experimental flux tracking performance <strong>of</strong> PM synchronous motor for zero speed<br />

at sampling time Ts<br />

= 200µ<br />

s. Reference flux from70% to 100% <strong>of</strong> nominal value . From the<br />

top: 1-reference stator flux amplitude(0.05Wb/div), 2- stator flux amplitude (0.05Wb/div), 3-<br />

electromagnetic torque (2Nm/div), 4- motor phase current (10A/div).<br />

It can be observed that achieved result is comparable <strong>with</strong> simulation results presented<br />

in Fig. 5.34.<br />

Figure 7.17. Experimental torque tracking performance <strong>of</strong> PM synchronous motor for zero<br />

speed at sampling time Ts<br />

= 200µ<br />

s. Reference torque from 0 to nominal value. From the top:1-<br />

reference torque (10Nm/div), 2 - estimated torque (10Nm/div), 3 - stator flux amplitude<br />

(0.1Wb/div), 4- motor phase current (10A/div)<br />

148


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

Figure 7.18. Experimental torque tracking performance <strong>of</strong> PM synchronous motor for zero<br />

speed (zoom) at sampling time Ts<br />

= 200µ<br />

s. Reference torque from 0 to nominal value. From<br />

the top:1- reference torque (10Nm/div), 2 - estimated torque (10Nm/div), 3 - stator flux<br />

amplitude (0.1Wb/div), 4- motor phase current (10A/div).<br />

The achieved result is comparable <strong>with</strong> simulation results presented in Fig. 5.42b.<br />

Experimental results presented in Fig. 7.16-7.18 confirm very well the effectiveness <strong>of</strong><br />

controller design and proper operation <strong>of</strong> flux and torque control loops for <strong>DTC</strong>-<strong>SVM</strong><br />

structure.<br />

149


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

Influence <strong>of</strong> sampling time for torque control loop in parallel <strong>DTC</strong>-<strong>SVM</strong><br />

The influence <strong>of</strong> sampling time on experimental torque tracking performance is<br />

illustrated in Fig. 7.19-7.21. The dynamic test has been carried out for the same<br />

condition ( Ω = 0rpm<br />

) as for simulation shown in Fig. 5.45. The controller parameters<br />

m<br />

has been set according to Table 5.4. In all oscilograms we may see proper operation <strong>of</strong><br />

the torque control loop for different sampling time used in practice.<br />

Figure 7.19. Experimental torque tracking performance <strong>of</strong> PM synchronous motor for zero<br />

speed at sampling time Ts<br />

= 100µ<br />

s. Reference torque from 0 to nominal value. From the top:1-<br />

reference torque (4Nm/div), 2 - estimated torque (4Nm/div)<br />

Figure 7.20. Experimental torque tracking performance <strong>of</strong> PM synchronous motor for zero<br />

speed at sampling time Ts<br />

= 200µ<br />

s. Reference torque from 0 to nominal value. From the top:1-<br />

reference torque (4Nm/div), 2 - estimated torque (4Nm/div).<br />

150


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

Figure 7.21. Experimental torque tracking performance <strong>of</strong> PM synchronous motor for<br />

zero speed at sampling time Ts<br />

= 400µ<br />

s. Reference torque from 0 to nominal value.<br />

From the top:1- reference torque (4Nm/div), 2 - estimated torque (4Nm/div).<br />

In this Chapter the results <strong>of</strong> experimental verification 3kW PMSM drive <strong>with</strong> two<br />

<strong>DTC</strong>-<strong>SVM</strong> schemes has been presented. As shown the drive performance confirms<br />

applied design methodology. The performance <strong>of</strong> both cascade and parallel <strong>DTC</strong>-<strong>SVM</strong><br />

control structure are similar. However, parallel structure has been selected for industrial<br />

manufacturing because <strong>of</strong> :<br />

• less noisy control algorithm (differentiation required in cascade structure –<br />

equation (5.6) is eliminated),<br />

• stator flux control in closed loop,<br />

• the same structure can be used for IM and PMSM control (universal control for<br />

AC motors).<br />

7.3.2 Speed control loop<br />

0peration <strong>with</strong> speed sensor<br />

Dynamic testes for the speed control loop were measured for sampling timeT<br />

= 200µ<br />

s.<br />

Please note that the digital speed controller parameters were selected according to Table<br />

5.5 (see Chapter 5.4). In Fig. 7.22-7.26 rotor speed tracking performance are presented.<br />

s<br />

151


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

Figure 7.22. Experimental start up and breaking to zero speed <strong>of</strong> PMSM motor controlled via<br />

<strong>DTC</strong>-<strong>SVM</strong> <strong>with</strong> the encoder speed signal taken to the feedback ( Ω<br />

m<br />

= 0rpm → 300rpm<br />

). From<br />

the top: 1- reference speed (180rpm/div), 2- measured speed (180rpm/div), 3 - electromagnetic<br />

torque (20Nm/div), 4- motor phase current (20A/div).<br />

Figure 7.23. Experimental speed reversal for PMSM motor controlled via <strong>DTC</strong>-<strong>SVM</strong> <strong>with</strong> the<br />

encoder speed signal taken to the feedback ( Ω<br />

m<br />

=−300rpm → 300rpm<br />

). From the top: 1-<br />

reference speed (360rpm/div), 2- measured speed (360rpm/div), 3 - electromagnetic torque<br />

(20Nm/div), 4- motor phase current (20A/div).<br />

152


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

Figure 7.24. Experimental start up and breaking to zero speed <strong>of</strong> PMSM motor controlled via<br />

<strong>DTC</strong>-<strong>SVM</strong> <strong>with</strong> the encoder speed signal taken to the feedback ( Ω<br />

m<br />

= 0rpm → 1500rpm<br />

).From<br />

the top: 1- reference speed (900rpm/div), 2- measured speed (900rpm/div), 3 - electromagnetic<br />

torque (20Nm/div), 4- motor phase current (20A/div).<br />

Figure 7.25. Experimental speed reversal for PMSM motor controlled via <strong>DTC</strong>-<strong>SVM</strong> <strong>with</strong> the<br />

encoder speed signal taken to the feedback ( Ω<br />

m<br />

=−1500rpm → 1500rpm<br />

). From the top: 1-<br />

reference speed (1800rpm/div), 2- measured speed (1800rpm/div), 3 - electromagnetic torque<br />

(20Nm/div), 4- motor phase current (20A/div).<br />

153


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

Figure 7.26. Experimental speed reversal for PMSM motor controlled via <strong>DTC</strong>-<strong>SVM</strong> <strong>with</strong> the<br />

encoder speed signal taken to the feedback ( Ω<br />

m<br />

=−1200rpm → 1200rpm<br />

). From the top: 1-<br />

stator flux component Ψ<br />

sα<br />

(0.25Wb/div), 2- measured speed (360rpm/div), 3 - electromagnetic<br />

torque (20Nm/div), 4- motor phase current (20A/div).<br />

Figure 7.27. Experimental response to load torque step change <strong>of</strong> PMSM motor controlled via<br />

<strong>DTC</strong>-<strong>SVM</strong> <strong>with</strong> the encoder speed signal taken to the feedback at Ω = 0rpm<br />

.<br />

( M = 0Nm → 10Nm<br />

) From the top: 1- reference speed (180rpm/div), 2- measured speed<br />

l<br />

(180rpm/div), 3 - electromagnetic torque (5Nm/div), 4- motor phase current (10A/div).<br />

m<br />

154


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

Figure 7.28. Experimental response to load torque step change <strong>of</strong> PMSM motor controlled via<br />

<strong>DTC</strong>-<strong>SVM</strong> <strong>with</strong> the encoder speed signal taken to the feedback at Ω = 300rpm<br />

.<br />

( M = 0Nm → 10Nm<br />

).From the top: 1- reference speed (180rpm/div), 2- measured speed<br />

l<br />

(180rpm/div), 3 - electromagnetic torque (5Nm/div), 4- motor phase current (10A/div).<br />

m<br />

Figure 7.29. Experimental response to load torque step change <strong>of</strong> PMSM motor controlled via<br />

<strong>DTC</strong>-<strong>SVM</strong> <strong>with</strong> the encoder speed signal taken to the feedback at Ω = 1500rpm<br />

.<br />

( M = 0Nm → 10Nm<br />

). From the top: 1- reference speed (900rpm/div), 2- measured speed<br />

l<br />

(900rpm/div), 3 - electromagnetic torque (5Nm/div), 4- motor phase current (10A/div).<br />

m<br />

155


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

Experimental results presented in Figures 7.22-7.29 well confirm the effectiveness <strong>of</strong><br />

developed controller synthesis methodology and proper operation <strong>of</strong> speed control loop<br />

for parallel <strong>DTC</strong>-<strong>SVM</strong> structure.<br />

Sensorless speed operation<br />

Dynamic tests for the speed control loop <strong>with</strong>out motion sensor were measured for<br />

sampling time T = 200µ<br />

s. Please note that the digital speed controller parameters were<br />

s<br />

selected exactly like for operation <strong>with</strong> speed sensor according to Table 5.5 (see<br />

Chapter 5.4).<br />

The results <strong>of</strong> speed estimator dynamic test are presented in Fig. 7.30. In this test speed<br />

controller operates <strong>with</strong> the encoder signal in feedback and speed estimator works in<br />

open loop fashion.<br />

Figure 7.30. Experimental dynamic test <strong>of</strong> the speed estimation. Speed reversal Ω =± 300rpm<br />

From the top: 1- reference speed (180rpm/div), 2- measured speed (180rpm/div), 3 - estimated<br />

speed (180rpm/div), 4- error <strong>of</strong> estimated speed (5%/div).<br />

m<br />

The typical dynamic performance tests <strong>of</strong> sensorless <strong>DTC</strong>-<strong>SVM</strong> drive has been<br />

illustrated in Fig. 7.31-7.36. Start up and breaking to zero speed for different speed level<br />

are shown in Fig. 7.31 and 7.33. Also, the speed reversal for low (Fig. 7.32) and<br />

nominal (Fig. 7.34) speed are presented. The half load torque step change tests are<br />

shown in Fig. 7.36 and 7.37.<br />

156


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

Figure 7.31. Experimental start up and breaking to zero speed <strong>of</strong> PMSM motor controlled via<br />

<strong>DTC</strong>-<strong>SVM</strong> <strong>with</strong> the estimated speed signal taken to the feedback ( Ω<br />

m<br />

= 0rpm → 300rpm<br />

).<br />

From the top: 1- reference speed (180rpm/div), 2- measured speed (180rpm/div), 3 -<br />

electromagnetic torque (20Nm/div), 4- motor phase current (20A/div).<br />

Figure 7.32. Experimental speed reversal for PMSM motor controlled via <strong>DTC</strong>-<strong>SVM</strong> <strong>with</strong> the<br />

estimated speed signal taken to the feedback ( Ω<br />

m<br />

=−300rpm → 300rpm<br />

). From the top: 1-<br />

reference speed (360rpm/div), 2- measured speed (360rpm/div), 3 - electromagnetic torque<br />

(20Nm/div), 4- motor phase current (20A/div).<br />

157


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

Figure 7.33. Experimental speed step response for PMSM motor controlled via <strong>DTC</strong>-<strong>SVM</strong> <strong>with</strong><br />

the estimated speed signal taken to the feedback ( Ω<br />

m<br />

= 0rpm → 1500rpm<br />

).From the top: 1-<br />

reference speed (900rpm/div), 2- measured speed (900rpm/div), 3 - electromagnetic torque<br />

(20Nm/div), 4- motor phase current (20A/div).<br />

Figure 7.34. Experimental speed reversal for PMSM motor controlled via <strong>DTC</strong>-<strong>SVM</strong> <strong>with</strong> the<br />

estimated speed signal taken to the feedback ( Ω<br />

m<br />

=−1500rpm → 1500rpm<br />

). From the top: 1-<br />

reference speed (1800rpm/div), 2- measured speed (1800rpm/div), 3 - electromagnetic torque<br />

(20Nm/div), 4- motor phase current (20A/div).<br />

158


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

Figure 7.35. Experimental speed reversal for PMSM motor controlled via <strong>DTC</strong>-<strong>SVM</strong> <strong>with</strong> the<br />

estimated speed signal taken to the feedback ( Ω<br />

m<br />

=−1200rpm → 1200rpm<br />

). From the top: 1-<br />

stator flux component Ψ<br />

sα<br />

(0.25Wb/div), 2- measured speed (360rpm/div), 3 - electromagnetic<br />

torque (20Nm/div), 4- motor phase current (20A/div).<br />

Figure 7.36. Experimental response to load torque step change <strong>of</strong> PMSM motor controlled via<br />

<strong>DTC</strong>-<strong>SVM</strong> <strong>with</strong> the estimated speed signal taken to the feedback at Ω = 300rpm<br />

.<br />

( M = 0Nm → 10Nm<br />

). From the top: 1- reference speed (180rpm/div), 2- measured speed<br />

l<br />

(180rpm/div), 3 - electromagnetic torque (10Nm/div), 4- motor phase current (10A/div).<br />

m<br />

159


DSP implementation <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> control<br />

Figure 7.37. Experimental response to load torque step change <strong>of</strong> PMSM motor controlled via<br />

<strong>DTC</strong>-<strong>SVM</strong> <strong>with</strong> the estimated speed signal taken to the feedback at Ω = 1500rpm<br />

.<br />

( M = 0Nm → 10Nm<br />

). From the top: 1- reference speed (900rpm/div), 2- measured speed<br />

l<br />

(900rpm/div), 3 - electromagnetic torque (10Nm/div), 4- motor phase current (10A/div).<br />

m<br />

Experimental results presented in Figures 7.31-7.37 confirm very well the effectiveness<br />

<strong>of</strong> speed estimation algorithm <strong>of</strong> speed control loop for parallel <strong>DTC</strong>-<strong>SVM</strong> structure.<br />

160


Summary and closing conclusion<br />

Chapter 8<br />

SUMMARY AND CLOSING CONCLUSIONS<br />

This thesis studied basic problems related to selection, investigation and implementation<br />

<strong>of</strong> the PWM inverter-fed permanent magnet synchronous motor (PMSM) drives suitable<br />

for serial manufacturing. The selected method should provide: robust start and operation<br />

in wide speed control range including zero speed, <strong>with</strong> and <strong>with</strong>out mechanical motion<br />

sensor; guarantee good and repeatable parameters <strong>of</strong> PMSM drive for wide power range<br />

(1-100kW). The control and protection algorithm should be implemented in simple and<br />

cheap microprocessor.<br />

To solve so formulated general task several related problems had to be solved. At first,<br />

the space vector based mathematical description and static characteristic <strong>of</strong> PMSM<br />

under different control modes were studied (Chapter 2). Secondly, the three phase<br />

voltage source inverter model including nonlinearities (dead time, semiconductor<br />

voltage drop, DC link pulsation) and pulse <strong>with</strong> modulation PWM techniques were<br />

presented (Chapter 3). Next, based on the study <strong>of</strong> most important high performance<br />

control methods as field oriented control (FOC), and direct torque control (<strong>DTC</strong>), the<br />

method called direct torque control <strong>with</strong> space vector modulator (<strong>DTC</strong>-<strong>SVM</strong>) has been<br />

selected for further consideration (Chapter 4). This methods combines main advantage<br />

<strong>of</strong> FOC (space vector modulator and fixed switching frequency) and <strong>DTC</strong> (simple<br />

structure, rotor parameter independent), as well as eliminates disadvantages like:<br />

coordinate transformation, the need <strong>of</strong> internal current control loops, high sampling<br />

time, high torque and current ripple at steady state operation, etc.<br />

Consequently the most important contribution <strong>of</strong> this thesis is included in the Chapter 5,<br />

where the two basic variants <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> schemes: series (cascade) and parallel<br />

structures <strong>of</strong> flux and torque controllers are presented (Fig. 5.2 and Fig. 5.25). Also, a<br />

systematic methodology <strong>of</strong> digital controller design for these both <strong>DTC</strong>-<strong>SVM</strong> variants<br />

have been given. This methodology has been verified by Simulink (using simplified<br />

discrete transfer function) and Saber (using full motor and inverter model) simulation<br />

studies. The influence <strong>of</strong> sampling time selection on controller design has also been<br />

discussed.<br />

The main problems associated <strong>with</strong> PMSM sensorless speed operation are presented in<br />

the Chapter 6. It should be noted that PMSM differ from IM drives mainly in:<br />

• PMSM parameters strongly depend on construction ,<br />

161


Summary and closing conclusion<br />

• position <strong>of</strong> PM flux has to be known prior to start up to achieve smooth<br />

operation.<br />

Therefore, for robust starting <strong>of</strong> PMSM <strong>with</strong>out the temporary rotor reversal a simple<br />

initialization algorithm has been used. This algorithm performs two tests: short and the<br />

longer voltage pulses generated by the PWM inverter. The used speed estimation<br />

algorithm is based on stator flux vector and torque angle estimation and does not<br />

operate accurately in zero speed region. However, it allows robust start and closed<br />

speed operation in the speed range above 10% <strong>of</strong> nominal speed. For application where<br />

high performance operation around zero speed are required, the <strong>DTC</strong>-<strong>SVM</strong> drive <strong>with</strong><br />

motion sensor (encoder) is recommended. The effectiveness <strong>of</strong> <strong>DTC</strong>-<strong>SVM</strong> scheme <strong>with</strong><br />

and <strong>with</strong>out motion sensor has been proved by simulation and experimental results<br />

(Chapter 7).<br />

Simulation study and experimental results have shown that from two variants <strong>of</strong> <strong>DTC</strong>-<br />

<strong>SVM</strong> schemes the parallel structure is more flexible in torque and flux controller<br />

design. Also, because <strong>of</strong> lack <strong>of</strong> the differentiation in the main control path (compare<br />

Fig. 5.2 and Fig. 5.25), it is less sensitive to noise which inherently associates signal<br />

processing in power electronic converters. Therefore, the parallel structure has been<br />

selected for industrial manufacturing and implemented on digital signal processor<br />

(DSP).<br />

Thanks to direct torque and flux control structure the described control is suitable to<br />

almost all – industrial applications including electrical vehicles (for example hybrid<br />

cars).<br />

Finally, it should be stressed that the developed system was brought into serial<br />

production. Presented algorithm <strong>DTC</strong>-<strong>SVM</strong> has been used in new generation <strong>of</strong> inverter<br />

drives produced by Polish company Power Electronic Manufacture – “TWERD”,<br />

Toruń.<br />

162


Appendices<br />

APPENDICES<br />

A1 Rotor and stator <strong>of</strong> PMSM machine<br />

A1.1. View <strong>of</strong> rotor (on the left side) and stator armature (on the right side) <strong>of</strong> PMSM.<br />

A2 Basic transformation<br />

y<br />

β<br />

K B<br />

K<br />

θ K<br />

K β<br />

K y<br />

K<br />

x<br />

K α<br />

θ K<br />

α<br />

x<br />

K A<br />

K C<br />

Fig. A2.1. <strong>Space</strong> vector representation in stationary α,<br />

β coordinates and synchronous<br />

rotating x,<br />

y coordinates.<br />

A, BC , ⇒ xy ,<br />

2<br />

Kx = [ KA cos θ<br />

K<br />

+ KB cos(2 π / 3 − θK ) + KC cos(2 π / 3 + θK<br />

)]<br />

3<br />

2<br />

Ky = [ KA cos( π /2 + θK ) + KB cos(2 π /3 − ( π /2 + θK )) + KC cos(2 π /3 + π /2 + θK<br />

)]<br />

3<br />

163


Appendices<br />

2<br />

Kx = [ UA cos θK + KB cos( θK − 2 π / 3) + KC cos( θK<br />

+ 2 π / 3)]<br />

3<br />

2<br />

Ky = [ −KA sin θK −KB sin( θK − 2 π / 3) − KC sin( θK<br />

+ 2 π / 3)]<br />

3<br />

⎡K<br />

A ⎤<br />

⎡K<br />

x ⎤ ⎡ cosθK cos( θK − 2 π / 3) cos( θK<br />

+ 2 π / 3) ⎤⎢<br />

K<br />

⎥<br />

⎢ B<br />

K<br />

⎥ = ⎢<br />

y sinθK sin( θK 2 π / 3) sin( θK<br />

2 π / 3)<br />

⎥⎢ ⎥<br />

⎣ ⎦ ⎣− − − − + ⎦<br />

⎢⎣<br />

K ⎥<br />

C ⎦<br />

ABC , , ⇒ α,<br />

β θ = 0<br />

2 1 1<br />

K = ( KA B C)<br />

3 − α<br />

K K<br />

2 − 2<br />

2 3 3 1<br />

K = ( KB C) (<br />

B C)<br />

3 2 − β<br />

K K K<br />

2 = 3<br />

−<br />

⎡ 1 1 ⎤<br />

1 − − ⎡ K<br />

A ⎤<br />

⎡Kα<br />

⎤ 2 ⎢ 2 2 ⎥<br />

⎢<br />

K<br />

⎥<br />

⎢ B<br />

K<br />

⎥ = ⎢<br />

⎥<br />

β 3 3 3 ⎢ ⎥<br />

⎣ ⎦ ⎢<br />

⎥<br />

⎢0<br />

− ⎢⎣K<br />

⎥<br />

C ⎦<br />

⎣ 2 2<br />

⎥<br />

⎦<br />

x, y⇒<br />

A, B,<br />

C<br />

K<br />

KA = Kxcosθ K<br />

+ Kycos( π / 2 + θK) = KxcosθK − KysinθK<br />

KB = Kxcos(2 π / 3 − θK) + Kycos(2 π / 3 − ( π / 2 + θK)) = Kxcos( θK −2 π / 3) −Kysin( θK<br />

−2 π / 3)<br />

K = K cos(2 π / 3 + θ ) + K cos(2 π / 3 + ( π / 2 + θ )) = K cos( θ + 2 π / 3) − K sin( θ + 2 π / 3)<br />

C x K y K x K y K<br />

⎡K<br />

A⎤ ⎡ cosθK −sinθK<br />

⎤<br />

⎢ K<br />

x<br />

K<br />

⎥ ⎢<br />

B<br />

cos( θK 2 π / 3) sin( θK<br />

2 π / 3)<br />

⎥ ⎡ ⎤<br />

⎢ ⎥<br />

=<br />

⎢<br />

− − −<br />

⎥⎢ K<br />

⎥<br />

y<br />

⎢K<br />

⎥ ⎢<br />

C<br />

cos( θK + 2 π / 3) − sin( θK<br />

+ 2 π / 3) ⎥ ⎣ ⎦<br />

⎣ ⎦ ⎣ ⎦<br />

α, β ⇒ ABC , , θ = 0<br />

K<br />

KA<br />

=<br />

K α<br />

1 3<br />

KB<br />

=− Kα<br />

+ K<br />

2 2<br />

1 3<br />

KC<br />

=− Kα<br />

− K<br />

2 2<br />

β<br />

β<br />

164


Appendices<br />

⎡ ⎤<br />

⎢ 1 0 ⎥<br />

⎡K<br />

A ⎤ ⎢ ⎥<br />

⎢ 1 3 K<br />

K<br />

⎥ ⎢ ⎥ ⎡ α ⎤<br />

⎢ B ⎥<br />

=<br />

⎢<br />

−<br />

2 2 ⎥⎢ K<br />

⎥<br />

β<br />

⎢K<br />

⎣ ⎦<br />

⎣ ⎥<br />

C ⎦ ⎢ ⎥<br />

⎢ 1 3⎥<br />

⎢<br />

− −<br />

⎣ 2 2 ⎥⎦<br />

xy , ⇒ α,<br />

β<br />

Kα = Kxcosθ K<br />

+ Kycos( π / 2 + θK) = KxcosθK − KysinθK<br />

Kβ = Kxcos( π / 2 − θK) + KycosθK = KxsinθK + KycosθK<br />

⎡Kα<br />

⎤ ⎡cosθK<br />

−sinθ<br />

K<br />

K⎤⎡ x ⎤<br />

⎢<br />

K<br />

⎥ = ⎢<br />

β sinθK<br />

cosθ<br />

⎥⎢ K<br />

⎥<br />

⎣ ⎦ ⎣<br />

K ⎦⎣ y ⎦<br />

α, β ⇒ x,<br />

y<br />

Kx = Kα cosθ K<br />

+ Kβ cos( π / 2 − θK) = Kα cosθK + Kβ<br />

sinθK<br />

Ky = Kα cos( π / 2 + θK) + Kβ cosθK =− Kα sinθK + Kβ<br />

cosθK<br />

⎡Kx ⎤ ⎡ cosθK sinθ<br />

K<br />

K ⎤⎡ α ⎤<br />

⎢<br />

K<br />

⎥ = ⎢<br />

y sinθK cosθ<br />

⎥⎢ K<br />

⎥<br />

⎣ ⎦ ⎣−<br />

K⎦⎣ β ⎦<br />

A3 Model <strong>of</strong> PM synchronous motor<br />

# This template models the permanent magnet synchronous motor(pmsm)<br />

# t1,t2 and t3 are motor input terminals<br />

# rotor speed (in rad/s) is the output<br />

# rs-Stator windings' resistence per phase(in Ohms)<br />

# ld-d_axis inductance(in H)<br />

# lq-q_axis inductance(in H)<br />

# pm-Rotor magnet flux(Wb)<br />

# j-Moment <strong>of</strong> inertia(in kgm2)<br />

# d- Damping constant (Nm/rad/s)<br />

# tl-motor load (Nm)<br />

# p-Number <strong>of</strong> pole pairs<br />

# power-The total input power (W)<br />

# Assumptions:No core losses,no saturation,thermal effects<br />

#(rs,ld,lq and pm values are constants)<br />

element template pmsm_dtc t1 t2 t3 t0 speed out_me out_psi out_thetam out_ualf out_ubet out_psia<br />

out_psib out_ialf out_ibet out_theta<br />

out_tl=rs,ld,lq,pm,d,tl,j,p,init_theta_m,init_omega_m,omega_m_const<br />

electrical t1,t2,t3,t0 # motor input terminals and stator neutral point<br />

output nu speed,<br />

out_me,out_psi,out_thetam,out_ualf,out_ubet,out_psia,out_psib,out_ialf,out_ibet,out_theta,out_tl<br />

number rs=0.692,ld=6m,lq=6m,pm=0.26379,d=0.002044,tl=0.0,j=0.003,p=3.0,<br />

init_theta_m=0.0,init_omega_m=0.0,omega_m_const=0.0<br />

{<br />

165


Appendices<br />


Appendices<br />

i(t1->t0)+=it1<br />

it1: it1=id*cos(p*theta_m)-iq*sin(p*theta_m)<br />

i(t2->t0)+=it2<br />

it2: it2=id*cos(p*theta_m-y)-iq*sin(p*theta_m-y)<br />

i(t3->t0)+=it3<br />

it3: it3=id*cos(p*theta_m+y)-iq*sin(p*theta_m+y)<br />

omega_m: (te-tl-d*omega_m)/j=d_by_dt(omega_m)<br />

#omega_m: omega_m=omega_m_const*math_pi/30.0<br />

theta_m: omega_m=d_by_dt(theta_m)<br />

speed: speed=omega_m<br />

out_me: out_me=te<br />

out_psi: out_psi=psi<br />

out_thetam:out_thetam=theta_m<br />

out_ualf: out_ualf=va<br />

out_ubet: out_ubet=vb<br />

out_psia: out_psia=psia<br />

out_psib: out_psib=psib<br />

out_ialf: out_ialf=ialf<br />

out_ibet: out_ibet=ibet<br />

out_theta: out_theta=theta<br />

out_tl: out_tl=tl<br />

}<br />

}<br />

A4 Motor parameters<br />

Surface type motor<br />

Power P 3kW<br />

Number <strong>of</strong> pole pairs p 3<br />

Phase current I(rms) 6.9A<br />

Phase voltage U(rms) 70V<br />

Magnetic flux-linkage Ψ 0.264 Wb<br />

Rotor speed<br />

Ω<br />

PM<br />

m<br />

3000rpm<br />

Nominal torque Me 20Nm<br />

Moment <strong>of</strong> the inertia J 0.0174kgm 2<br />

Stator winding resistance Rs 0.692 Ω<br />

Stator d-axis inductance Ld 6mH<br />

Stator d-axis inductance Lq 6mH<br />

Interior type motor<br />

Power P 2,2kW<br />

Number <strong>of</strong> pole pairs p 3<br />

Phase current I(rms) 4.1A<br />

Rated voltage U(rms) 380V<br />

Magnetic flux-linkage Ψ 0.4832 Wb<br />

Rotor speed<br />

Ω<br />

PM<br />

m<br />

1750rpm<br />

Nominal torque Me 12Nm<br />

Moment <strong>of</strong> the inertia J 0.010074kgm 2<br />

Stator winding resistance Rs 3.3 Ω<br />

Stator d-axis inductance Ld 41.59mH<br />

Stator d-axis inductance Lq 57.06mH<br />

167


Appendices<br />

A5 Voltage Source Inverter parameters<br />

Detailed date <strong>of</strong> IGBT transistors (module TOSHIBA M675Q2YS50):<br />

U = 1200V<br />

, I = 75A<br />

CE<br />

C<br />

UCEsat<br />

= 2.8 − 3.6V<br />

, forward diode voltage 2.4 − 3.5V<br />

Turn on time tON<br />

= 0.2µ<br />

s, Turn <strong>of</strong>f time tOFF<br />

= 0.6µ<br />

s<br />

Delay <strong>of</strong> IGBT drivers tONd<br />

= 0.5µ<br />

s TOFFd<br />

= 1µ<br />

s<br />

TON = tON + tONd<br />

= 0.7µ<br />

s total turn on time <strong>of</strong> IGBT<br />

TOFF = tOFF + tOFFd<br />

= 1.6µ<br />

s total turn <strong>of</strong>f time <strong>of</strong> IGBT<br />

Dead time T = 2.5µ<br />

s<br />

A6 PI speed controller<br />

d<br />

The commonly used in industrial application speed controller is a Proportional-Integral<br />

PI controller thanks to possibility to reduce the speed error between the reference (<br />

and actual rotor speed (<br />

X<br />

ref<br />

)<br />

X<br />

m<br />

) to zero (see Fig. A6.1). The output signal <strong>of</strong> controller is a<br />

reference torque, which has upper and lower limitation for this value equal the nominal<br />

torque or more than 130% <strong>of</strong> nominal torque. The output <strong>of</strong> the speed controller acts as<br />

a current reference command for the current controllers. This current command is<br />

limited to a nominal current <strong>of</strong> the motor.<br />

The speed controller demands produce proper electromagnetic torque.<br />

X_<br />

ref<br />

Reference signal<br />

−<br />

error signal<br />

Kp<br />

Y L<br />

<strong>Control</strong>ler output<br />

a)<br />

Feedback signal<br />

X_<br />

m<br />

Kp<br />

T<br />

i<br />

∫<br />

X _ ref<br />

Reference signal<br />

b)<br />

−<br />

Feedback signal<br />

error signal<br />

X _ m<br />

Kp<br />

Kp<br />

T<br />

i<br />

∫<br />

1<br />

Y NL<br />

lim_max<br />

lim_ min<br />

−<br />

Y E<br />

Y L<br />

<strong>Control</strong>ler output<br />

T i<br />

A.6.1. General structure <strong>of</strong> Proportional- Integral controller <strong>with</strong>out antiwindup (a) and <strong>with</strong><br />

antwindup (b).<br />

168


Appendices<br />

A7 PWM technique – six step mode<br />

Six-stepped-voltage waveforms are rich in harmonics. These time harmonics produce<br />

respective stator current harmonics, which in turn interact <strong>with</strong> fundamental air gap<br />

flux, generating harmonics torque pulsations. The torque pulsations are undesirable:<br />

they generate audible noise, speed pulsations, and losses. In case <strong>of</strong> supplied motor by<br />

using only active vectors (six step mode) we can observed non sinusoidal current, which<br />

generates torque ripples <strong>with</strong> frequency <strong>of</strong> six time fundamental frequency <strong>of</strong> supplied phase<br />

voltages.<br />

Fig. A7.1. Experimental operation in six step mode. From the left side stator voltages in α , β<br />

coordinates, From right side voltage trajectory.<br />

Fig. A7.2. Experimental operation in six step mode. From the left side stator currents in α , β<br />

coordinates, From right side stator current trajectory.<br />

169


Appendices<br />

Fig. A7.3. Experimental operation in six step mode. From the top: α stator voltage,<br />

electromagnetic torque in machine, phase current.<br />

170


List <strong>of</strong> symbol<br />

List <strong>of</strong> symbols<br />

Symbol (general)<br />

X - instantaneous value<br />

X<br />

N<br />

- normalized value<br />

X - vector<br />

X ∗ - conjugate vector<br />

X - amplitude <strong>of</strong> vector<br />

Re( X ) – real part <strong>of</strong> X<br />

Im( X ) – imaginary part <strong>of</strong> X<br />

Symbol (special)<br />

α,<br />

β - stator fixed system<br />

dq , - rotor reference system<br />

x,<br />

y - general reference system<br />

Ls<br />

- stator inductance<br />

Zs<br />

- stator impedance<br />

M<br />

s<br />

- mutual inductance<br />

Is<br />

- phase current value<br />

U<br />

s<br />

- phase voltage value<br />

Ψ<br />

s<br />

- phase flux value<br />

P -active power<br />

Q - reactive power<br />

S - apparent power<br />

Pe<br />

- electro-magnetic power<br />

Ωm<br />

- mechanical rotor speed<br />

Ωs<br />

- synchronous speed<br />

cosφ - power factor<br />

δ<br />

I<br />

, δ Ψ<br />

- torque angle<br />

φ - power angle<br />

R<br />

s<br />

- stator resistance<br />

L<br />

d<br />

, L<br />

d<br />

- direct and quadrature inductances<br />

θr<br />

- electrical rotor position<br />

γ<br />

m<br />

- mechanical rotor position<br />

pb<br />

- number <strong>of</strong> pole pairs<br />

Ψ<br />

PM<br />

- rotor flux <strong>of</strong> permanent magnets<br />

M<br />

e<br />

- electromagnetic torque<br />

M<br />

es<br />

- synchronous torque<br />

M<br />

er<br />

- reluctance torque<br />

M<br />

l<br />

- load torque (external load torque)<br />

M<br />

d<br />

- dynamic torque<br />

Jm<br />

- motor moment <strong>of</strong> inertia<br />

Jl<br />

- load moment <strong>of</strong> inertia<br />

J - moment <strong>of</strong> inertia <strong>of</strong> total system (sum <strong>of</strong> J<br />

m<br />

and J<br />

l<br />

)<br />

171


List <strong>of</strong> symbol<br />

Subscripts<br />

A , B , C - denote arbitrary phase quantities in a system <strong>of</strong> natural coordinate ABC. , ,<br />

d , q - arbitrary direct and quadrature components in a system <strong>of</strong> rotor coordinate dq. ,<br />

α , β - arbitrary alpha and beta components in a system <strong>of</strong> stator coordinate α,<br />

β .<br />

x , y - denote arbitrary components in a system <strong>of</strong> general coordinate x,<br />

y .<br />

.. r - denotes value <strong>of</strong> rotor<br />

.. s - denotes value <strong>of</strong> stator<br />

.. _max - maximum value<br />

.. _min – minimum value<br />

.. _ ref - reference value<br />

.. _ est - estimated value<br />

.. _ amp -amplitude value<br />

.. _ rms - root mean square value<br />

.. _ LL - line to line value<br />

* - reference value<br />

^ - estimated value<br />

Abbreviations<br />

RSM – reluctance synchronous motor<br />

BLDCM – blushless DC motor<br />

PMSM – permanent magnet synchronous motor<br />

IPMSM - interior permanent magnet synchronous motor<br />

SPMSM - surface permanent magnet synchronous motor<br />

EMF – electro-magnetic force<br />

VSI - voltage source inverter<br />

<strong>SVM</strong> – space vector modulator<br />

PWM – pulse width modulation<br />

PWM-VSI – voltage source inverter <strong>with</strong> PWM<br />

<strong>DTC</strong> - direct torque control<br />

<strong>DTC</strong>-<strong>SVM</strong> - direct torque control <strong>with</strong> space vector modulator<br />

RFOC - rotor field oriented control<br />

SFOC - stator field oriented control<br />

CTAC - constant torque angle control<br />

MTPAC - maximum torque per ampere control<br />

UPFC - unity power factor control<br />

CSFC - constant stator flux control<br />

172


References<br />

Books and PhD-Thesis<br />

1. I. Boldea, S.A. Nasar, “Electrical drives”, CRC Press, 1999.<br />

2. G. F. Franklin, J. D. Powell, “Feedback <strong>Control</strong> <strong>of</strong> Dynamic Systems”, Addison-<br />

Wesley Publishing Company, 1994.<br />

3. M. P. Kaźmierkowski, H. Tunia, “Automatic <strong>Control</strong> <strong>of</strong> Converter-Fed drives”,<br />

Elsevier, 1994.<br />

4. M. P. Kaźmierkowski, R. Krishnan, F. Blaabjerg, “<strong>Control</strong> in Power Electronics”,<br />

Academic Press, 2002.<br />

5. P.C. Krause, O. Wasynczuk, S.D. Sudh<strong>of</strong>f, “Analysis <strong>of</strong> Electric Machinery”, IEEE<br />

Press, 1995.<br />

6. R. Krishnan, “Electric Motor Drives – Modeling, Analysis, and <strong>Control</strong>”, Prentice Hall,<br />

2001.<br />

7. T. Kaczmarek, K. Zawirski, “Układy napędowe z silnikiem synchronicznym”,<br />

Politechnika Poznańska, 2000.<br />

8. M. Malinowski, "Sensorless <strong>Control</strong> Strategies for Three-Phase PWM Rectifiers", PhD<br />

Thesis, Warsaw University <strong>of</strong> Technology, 2001.<br />

9. Ch. Perera, “Sensorless <strong>Control</strong> <strong>of</strong> Permanent Magnet Synchronous Motor Drives”,<br />

PhD thesis, Institute <strong>of</strong> Energy Technology, 2002.<br />

10. K. Rajashekara, A. Kawamura, K. Matsuse, “Sensorless <strong>Control</strong> <strong>of</strong> AC Motor Drives”,<br />

IEEE Press,1996, USA.<br />

11. M. Zelechowski, “<strong>Space</strong> <strong>Vector</strong> Modulated – <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong>led (<strong>DTC</strong>-<strong>SVM</strong>)<br />

Inverter-Fed Induction Motor Drive”, Phd Thesis , Warsaw University <strong>of</strong> Technology,<br />

2005.<br />

12. P. Vas, “Sensorless <strong>Vector</strong> and <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong>”, Oxford University Press, 1998.<br />

Model <strong>of</strong> PMSM<br />

13. C. Bowen; Z. Jihua; R. Zhang; “Modeling and simulation <strong>of</strong> permanent magnet<br />

synchronous motor drives”, Electrical Machines and Systems, ICEMS 2001, Vol. 2, 18-<br />

20 Aug. 2001, pp. 905 – 908.<br />

14. P. Pillay, R. Krishnan, “Modeling <strong>of</strong> permanent magnet motor drives”, Transactions on<br />

Industrial Electronics, Vol. 35, Nov. 1988, pp. 537 – 541.<br />

15. N. Urasaki; T. Senjyu; K Uezato, “An Accurate Modeling for Permanent Magnet<br />

Synchronous Motor Drives”, Applied Power Electronics Conference and Exposition,<br />

APEC 2000, Vol. 1, 6-10 Feb. 2000, pp:387 – 392.<br />

16. A.H. Wijenayake, P.B. Schmidt, “Modeling and analysis <strong>of</strong> permanent magnet<br />

synchronous motor by taking saturation and core loss into account”, International<br />

Conference on Power Electronics and Drive Systems, Vol. 2, 26-29 May 1997, pp. 530<br />

– 534.<br />

Inverter and PWM <strong>Modulation</strong><br />

17. L. Ben-Brahim, "The analysis and compensation <strong>of</strong> dead-time effects in three phase<br />

PWM inverters", Industrial Electronics Society, 1998. IECON '98. Proceedings <strong>of</strong> the<br />

24th Annual Conference <strong>of</strong> the IEEE, Vol. 2d, 31 Aug.-4 Sept. 1998, pp.792-797.<br />

18. V. Blasko, "Analysis <strong>of</strong> a hybrid PWM based on modified space-vector and trianglecomparison<br />

methods", IEEE Transactions on Industry Applications, Vol. 33, Issue: 3,<br />

May-June 1997, pp.756-764.<br />

19. F.Blaabjerg, J.K. Pedersen, P. Thoegersen, “Improved modulation techniques for<br />

PWM-VSI drives”, IEEE Transactions on Industrial Electronics, Feb. 1997, Vol. 44, pp.<br />

87 – 95.<br />

20. J.W. Choi; S.K. Sul, "Inverter output voltage synthesis using novel dead time<br />

compensation", IEEE Transactions on Power Electronics, Vol. 11, Issue: 2, March<br />

1996, pp.221-227.<br />

173


References<br />

21. D.W. Chung, J. Kim, S.K. Sul, "Unified voltage modulation technique for real-time<br />

three-phase power conversion", IEEE Transactions on Industry Applications, Vol. 34,<br />

Issue: 2, March-April 1998, pp.374-380.<br />

22. A. Diaz, E.G. Strangas, "A novel wide range pulse width overmodulation method [for<br />

voltage source inverters]", Applied Power Electronics Conference and Exposition,<br />

APEC 2000.<br />

23. A.M. Hava, R.J. Kerkman, T.A. Lipo, "A high performance generalized discontinuous<br />

PWM algorithm", Applied Power Electronics Conference and Exposition, APEC '97<br />

Conference Proceedings 1997, Twelfth Annual, Vol. 2, 23-27 Feb. 1997, pp.886-894.<br />

24. D.G. Holmes, "The significance <strong>of</strong> zero space vector placement for carrier-based PWM<br />

schemes", IEEE Transactions on Industry Applications, Vol. 32, Issue: 5, Sept.-Oct.<br />

1996, pp.1122-1129.<br />

25. J. Holtz, "Pulsewidth modulation for electronic power conversion", Proceedings <strong>of</strong> the<br />

IEEE, Vol. 82, Issue: 8, Aug. 1994, pp.1194-1214.<br />

26. J. Holtz, W. Lotzkat, A.M. Khambadkone, "On continuous control <strong>of</strong> PWM inverters in<br />

the overmodulation range including the six-step mode", IEEE Transactions on Power<br />

Electronics, Vol. 8, Issue: 4, Oct. 1993, pp.546-553.<br />

27. J.L. Lin, "A new approach <strong>of</strong> dead-time compensation for PWM voltage inverters",<br />

IEEE Transactions on Circuits and Systems I: Fundamental Theory and Applications,<br />

Vol. 49, Issue: 4, April 2002, pp.476-483.<br />

28. D.C. Lee, G.M. Lee, "A novel overmodulation technique for space-vector PWM<br />

inverters", IEEE Transactions on Power Electronics, Vol. 13, Issue: 6, Nov. 1998,<br />

pp.1144-1151.<br />

29. M. Malinowski "Adaptive modulator for three-phase PWM rectifier/inverter", in proc.<br />

EPE-PEMC Conf., Kosice, 2000, pp.1.35-1.41.<br />

30. N. Urasaki, T. Senjyu, K. Uezato, T. Funabashi, “On-line dead-time compensation<br />

method for permanent magnet synchronous motor drive”, IEEE International<br />

Conference on Industrial Technology, CIT '02, 11-14 Dec. 2002, Vol. 1, pp. 268 – 273.<br />

FOC, <strong>DTC</strong> and <strong>DTC</strong>-<strong>SVM</strong><br />

31. F. Blaschke, "The principle <strong>of</strong> fields-orientation as applied to the Transvector closedloop<br />

control system for rotating-field machines", in Siemens Reviev 34, 1972, pp.217-<br />

220.<br />

32. G.S. Buja, M.P. Kazmierkowski, "<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>of</strong> PWM Inverter-Fed AC<br />

Motors-A Survey", IEEE Transactions on Industrial Electronics, Vol. 51, Aug. 2004,<br />

pp.744-757.<br />

33. D. Casadei, G. Serra, A. Tani, L. Zarri, F. Pr<strong>of</strong>umo, “Performance analysis <strong>of</strong> a speedsensorless<br />

induction motor drive based on a constant-switching-frequency <strong>DTC</strong><br />

scheme”; IEEE Transactions on Industry Applications, March-April 2003, Volume 39,<br />

pp. 476 – 484.<br />

34. D. Casadei, F. Pr<strong>of</strong>umo, G. Serra, A. Tani, “FOC and <strong>DTC</strong>: two viable schemes for<br />

induction motors torque control”; IEEE Transactions on Power Electronics, Sept. 2002,<br />

Vol. 17, pp. 779 – 787.<br />

35. T. Chun; W.Y. Hu, “Research on the direct torque control in electromagnetic<br />

synchronous motor drive”, Power Electronics and Motion <strong>Control</strong> Conference, PEMC<br />

2000, 15-18 Aug. 2000. Vol. 3, pp. 1262 – 1265.<br />

36. S.K. Chung, H.S. Kim, Ch. G. Kim, M.-J. Youn, “A New Instantaneous <strong>Torque</strong> <strong>Control</strong><br />

<strong>of</strong> PM Synchronous Motor for High-Performance <strong>Direct</strong>-Drive Applications” IEEE<br />

Transaction on Power Electronics, May 1998, Vol. 12, No. 3.<br />

37. Ch. French, P. Acarnley, “<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>of</strong> Permanent Magnet Drivers”, IEEE<br />

Transaction on Industrial Application, Vol. 32, No. 5, 1996, pp. 1080-1088.<br />

38. L.-H. Hoang, “Comparison <strong>of</strong> field-oriented control and direct torque control for<br />

induction motor drives”, Industry Applications Conference, 1999. Thirty-Fourth IAS<br />

Annual Meeting, 3-7 Oct. 1999, Vol. 2, pp.1245 – 1252.<br />

174


References<br />

39. J.-I. Itoh, N. Nomura, H. Ohsawa, “A comparison between V/f control and positionsensorless<br />

vector control for the permanent magnet synchronous motor”, Proceedings <strong>of</strong><br />

the Power Conversion Conference, PCC Osaka 2-5 April 2002, Vol. 3, pp. 1310 – 1315.<br />

40. T.M. Jahns, “Motion control <strong>with</strong> permanent-magnet AC machines”, Proceedings <strong>of</strong> the<br />

IEEE, Aug. 1994, Vol. 82, pp. 1241 – 1252.<br />

41. C. Lascu, I. Boldea, F. Blaabjerg, “Variable-structure direct torque control - a class <strong>of</strong><br />

fast and robust controllers for induction machine drives”, IEEE Transactions on<br />

Industrial Electronics, Aug. 2004,Vol. 51, pp. 785 – 792.<br />

42. T. Lixin; Z. Limin; M.F. Rahman, Y. Hu, “A novel direct torque control for interior<br />

permanent-magnet synchronous machine drive <strong>with</strong> low ripple in torque and flux-a<br />

speed-sensorless approach”, IEEE Transactions on Industry Applications, Nov.-Dec.<br />

2003,Vol. 39, pp. 1748 – 1756.<br />

43. J. Lee; Ch.-G. Kim; M.-J. Youn, “A dead-beat type digital controller for the direct<br />

torque control <strong>of</strong> an induction motor”, IEEE Transactions on Power Electronics, Sept.<br />

2002, Vol. 17, pp. 739 – 746.<br />

44. H. Rasmussen, P. Vadstrup, H. Børsting, "Adaptive sensorless field oriented control <strong>of</strong><br />

PM motors including zero speed", IEEE International Symposium on Industrial<br />

Electronics, Ajaccio, France, May 2004.<br />

45. M. F. Rahman, L. Zhong, K. W. Lim, “A direct torque controlled interior permanent<br />

magnet synchronous motor drive incorporating field weakening”, IEEE-IAS’97, 1997 ,<br />

p. 67 - 74.<br />

46. M.F. Rahman, L. Zhong, K.W. Lim, “An Investigation <strong>of</strong> <strong>Direct</strong> and Indirect <strong>Torque</strong><br />

Conttrollers for PM Synchronous Motor Drivers”, Conference PEDS’97, May,<br />

Singapure, pp. 519-523.<br />

47. M. F. Rahman, L. Zhong, “Comparison <strong>of</strong> torque response <strong>of</strong> the interior permanent<br />

magnet motor under PWM current and direct torque control”, IEEE, 1999 , p. 1464 -<br />

1470.<br />

48. J. Rodriguez, J. Pontt, C. Silva, S. Kouro, H. Miranda, “A novel direct torque control<br />

scheme for induction machines <strong>with</strong> space vector modulation”; Power Electronics<br />

Specialists Conference, PESC’04, 20-25 June 2004, Vol. 2, pp. 1392 – 1397.<br />

49. D. Telford, M.W. Dunnigan, B.W. Williams, “A comparison <strong>of</strong> vector control and<br />

direct torque control <strong>of</strong> an induction machine”, 2000 IEEE 31st Annual Power<br />

Electronics Specialists Conference, 2000. PESC 00, 18-23 June 2000, Vol. 1, pp. 421 –<br />

426.<br />

50. L. Takahashi, T. Noguchi, “A new quick response and high efficiency strategy <strong>of</strong><br />

induction motor”, IAS, 1985 , p. 495-502.<br />

51. L. Tang; L. Zhong; F. Rahman, “Modeling and experimental approach <strong>of</strong> a novel direct<br />

torque control scheme for interior permanent magnet synchronous machine drive”,<br />

IECON 02, 5-8 Nov. 2002, Vol. 1, pp. 235 – 240.<br />

52. M.N. Uddin, T.S. Radwan, G.H. George, M.A. Rahman, “Performance <strong>of</strong> current<br />

controllers for VSI-fed IPMSM drive”, IEEE Transactions on Industry Applications,<br />

Volume 36, Issue 6, Nov.-Dec. 2000 Page(s):1531 – 1538.<br />

53. L. Xu, M. Fu “A Novel Sensorless <strong>Control</strong> Technique for Permanent Magnet<br />

Synchronous Motors (PMSM) using Digital Signal Processor (DSP)”, NEACON’97,<br />

Dayton, Ohio, July 14-17, 1997.<br />

54. L. Xu, M. Fu “A Sensorless <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> Technique for Permanent Magnet<br />

Synchronous Motors”, IEEE Industrial Applications Conferences,1999, Vol. 1.<br />

55. Y. Xue, X. Xu, T.G. Habetler, D.M. Divan, “A low cost stator flux oriented voltage<br />

source variable speed drive”, Industry Applications Society Annual Meeting, 7-12 Oct.<br />

1990, Vol.1, pp. 410 – 415.<br />

56. H. Yuwen,T. Cun, G. Yikang,Y. Zhiqing, L.X. Tang, M.F. Rahman “In-depth Reserch<br />

on <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>of</strong> Permanent Magnet Synchronous Motor”, IEEE Conference<br />

IECON 2002 ,November, Sevilla, Spain.<br />

175


References<br />

57. L. Zhong, M.F. Rahman, W.Y. Hu, Lim, K.W., Rahman, M.A., “A direct torque<br />

controller for permanent magnet synchronous motor drives”, IEEE Transactions on<br />

Energy Conversion, Sept. 1999, Vol. 14, pp. 637 – 642.<br />

58. L. Zhong, M.F. Rahman, W.Y.Hu, K.W. Lim, “Analysis <strong>of</strong> <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> in<br />

Permanent Magnet Synchronous Motor Drives”, IEEE Transaction on Power<br />

Electronics, Vol. 12, No. 13, May, 1997.<br />

59. L. Zhong, M.F. Rahman, „Voltage Switching Tables for <strong>Control</strong>led Interior Permanent<br />

Magnet Motor“, IEEE Conference IECON’99, Vol. 3, pp. 1445-1451.<br />

60. L.Zhong, M. F. Rahman, “A direct torque controller for permanent magnet synchronous<br />

motor drives <strong>with</strong>out a speed sensor”, IEEE Trans. On Energy Conversion, Vol.14,<br />

No.3 ,September , 1999.<br />

61. M.R Zolghardi, D.Diallo, D.Roye, “<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> System for Synchronous<br />

Machine”, Europen Conference on Power Electronics and Application,<br />

EPE’97,Trondheim, Norway,Vol. 3, pp. 694-699.<br />

62. M.R. Zolghardi, E. M. Olasagasti, D. Roye, “Steady State <strong>Torque</strong> Correction <strong>of</strong> <strong>Direct</strong><br />

<strong>Torque</strong> <strong>Control</strong>led PM Synchronous Machine”, IEEE International Conference on<br />

Electric Machines Drivers, 1997.<br />

63. M.R Zolghardi, J.Guiraud, J.Davoine, D. Roye, “A DSP <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong>ler for<br />

Permanent Magnet Synchronous Motor Drivers”, IEEE Power Electronics Conference,<br />

PESC’98, Vol.2, pp. 2055-2061.<br />

64. M.R Zolghardi, D. Roye, “Sensorless <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>of</strong> Synchronous Motor<br />

Drive” Proceedings <strong>of</strong> International Conference on Electrical Machines ICEM’ 98, Vol.<br />

2, 1998, Istanbul, Turkey.<br />

65. M.R Zolghardi, D. Diallo, D. Roye, “Start up Strategies for <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong>led<br />

Synchronous Machine ”, Europen Conference on Power Electronics and Application,<br />

EPE’97,Trondheim, Norway,Vol. 3, pp. 689-693.<br />

66. M.R Zolghardi, C. Pelissu, D. Roye, “Start up <strong>of</strong> a global <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong><br />

System”, IEEE Power Electronic Conference PESC’96, Vol. 1, pp. 370-374.<br />

Sensorless <strong>Control</strong> and Initial Position Estimation<br />

67. S. Bolognani, R. Oboe, M. Zigliotto, “Sensorless full-digital PMSM drive <strong>with</strong> EKF<br />

estimation <strong>of</strong> speed and rotor position”, IEEE Transactions on Industrial Electronics,<br />

Vol. 46, Issue 1, Feb. 1999, pp. 184 – 191.<br />

68. S. Brock, J. Deskur, K. Zawirski, , “Robust speed and position control <strong>of</strong> PMSM”,<br />

Proceedings <strong>of</strong> the IEEE International Symposium on Industrial Electronics, 1999, ISIE<br />

'99. 12-16 July 1999, Vol. 2, pp. 667 – 672.<br />

69. M.J. Corley, R.D. Lorenz, “Rotor position and velocity estimation for a salient-pole<br />

permanent magnet synchronous machine at standstill and high speeds”, IEEE<br />

Transactions on Industry Applications, July-Aug. 1998, Vol. 34, pp. 784 – 789.<br />

70. Z. Chen, M. Tomita, S. Ichikawa, S. Doki, S. Okuma, “Sensorless control <strong>of</strong> interior<br />

permanent magnet synchronous motor by estimation <strong>of</strong> an extended electromotive<br />

force”; Industry Applications Conference, 2000, Vol. 3, 8-12 Oct. 2000 pp. 1814 –<br />

1819.<br />

71. Z. Chen, M. Tomita, S. Doki, S. Okuma, “An extended electromotive force model for<br />

sensorless control <strong>of</strong> interior permanent-magnet synchronous motors”, IEEE<br />

Transactions on Industrial Electronics, Vol. 50, Issue 2, April 2003, pp. 288 – 295.<br />

72. Z. Chen, M. Tomita, S. Doki, S. Okuma, “New adaptive sliding observers for<br />

position- and velocity-sensorless controls <strong>of</strong> brushless DC motors”, IEEE Transactions<br />

on Industrial Electronics, Vol. 47, Issue 3, June 2000, pp. 582 – 591.<br />

73. R. Dhaouadi, N. Mohan, L. Norum, “Design and implementation <strong>of</strong> an extended<br />

Kalman filter for the state estimation <strong>of</strong> a permanent magnet synchronous motor”, IEEE<br />

Transactions on Power Electronics, Vol. 6, Issue 3, July 1991, pp. 491 – 497.<br />

176


References<br />

74. N. Ertugrul, P. Acarnley, “A new algorithm for sensorless operation <strong>of</strong> permanent<br />

magnet motors”; IEEE Transactions on Industry Applications, Jan.-Feb. 1994, Vol. 30,<br />

pp. 126 – 133.<br />

75. C. French, P. Acarnley, I. Al-Bahadly, “Sensorless position control <strong>of</strong> permanent<br />

magnet drives”; Industry Applications Conference, IAS '95, 8-12 Oct. 1995, Vol. 1, pp.<br />

61 – 68.<br />

76. M. Fu, L. Xu, “A novel sensorless control technique for permanent magnet synchronous<br />

motor (PMSM) using digital signal processor (DSP)”, Proceedings <strong>of</strong> the IEEE 1997<br />

National Aerospace and Electronics Conference, 14-17 July 1997, NAECON 1997,Vol.<br />

1, pp. 403 – 408.<br />

77. J. Hu, B. Wu, “New integration algorithms for estimating motor flux over a wide speed<br />

range”, IEEE Transactions on Power Electronics, , Sept. 1998, Vol. 13, pp. 969 – 977.<br />

78. Y. Jeong, R.D. Lorenz, T.M. Jahns, S. Sul, “Initial rotor position estimation <strong>of</strong> an<br />

interior permanent magnet synchronous machine using carrier-frequency injection<br />

methods”, IEEE International Electric Machines and Drives Conference, IEMDC'03, 1-<br />

4 June 2003, Vol. 2,pp. 1218 – 1223.<br />

79. L.A. Jones, J.H. Lang, “A state observer for the permanent-magnet synchronous<br />

motor”, IEEE Transactions on Industrial Electronics, Vol. 36, Issue 3, Aug. 1989 pp.<br />

374 – 382.<br />

80. J.P. Johnson, M. Ehsani, Y. Guzelgunler, “Review <strong>of</strong> sensorless methods for<br />

brushless DC”, Industry Applications Conference, 1999. Thirty-Fourth IAS Annual<br />

Meeting. Conference Record <strong>of</strong> the 1999 IEEE, 3-7 Oct. 1999, Vol. 1, pp. 143 – 150.<br />

81. J. S. Kim; S. K. Sul, “New stand-still position detection strategy for PMSM drive<br />

<strong>with</strong>out rotational transducers”, Applied Power Electronics Conference and Exposition,<br />

APEC '94, 13-17 Feb. 1994, Vol.1, pp. 363 – 369.<br />

82. J.-S. Kim, S.-K. Sul, “New approach for high-performance PMSM drives <strong>with</strong>out<br />

rotational position sensors”, IEEE Transactions on Power Electronics, Sept. 1997, Vol.<br />

12, pp. 904 – 911.<br />

83. T.-S. Low, T.-H. Lee, K.-T. Chang, “A nonlinear speed observer for permanentmagnet<br />

synchronous motors”, IEEE Transactions on Industrial Electronics, Vol. 40,<br />

Issue 3, June 1993, pp. 307 – 316.<br />

84. M. Linke, R. Kennel, J. Holtz, “Sensorless position control <strong>of</strong> permanent magnet<br />

synchronous machines <strong>with</strong>out limitation at zero speed”, IECON 02, 5-8 Nov. 2002,<br />

Vol. 1, pp. 674 – 679.<br />

85. N. Matsui, “Sensorless PM brushless DC motor drives”, Transactions on Industrial<br />

Electronics, IEEE , Vol. 43, April 1996, pp. 300 – 308.<br />

86. N. Matsui, M. Shigyo, “Brushless DC motor control <strong>with</strong>out position and speed<br />

sensors”; IEEE Transactions on Industry Applications, Jan.-Feb. 1992, Vol. 28 pp. 120<br />

– 127.<br />

87. N. Matsui, “Sensorless PM brushless DC motor drives”, IEEE Transactions on<br />

Industrial Electronics, April 1996, Vol. 43, pp. 300 – 308.<br />

88. S. Nakashima, Y. Inagaki, I. Miki, “Sensorless initial rotor position estimation <strong>of</strong><br />

surface permanent-magnet synchronous motor”, IEEE Transactions on Industry<br />

Applications, Nov.-Dec. 2000, Vol. 36, pp. 1598 – 1603.<br />

89. T. Noguchi, K. Yamada, S. Kondo, I. Takahashi, “Rotor position estimation method <strong>of</strong><br />

sensorless PM motor at rest <strong>with</strong> no sensitivity to armature resistance”, Industrial<br />

Electronics, <strong>Control</strong>, and Instrumentation, Proceedings <strong>of</strong> the 1996 IEEE IECON 1996,<br />

5-10 Aug. 1996,Vol. 2, pp. 1171 – 1176.<br />

90. S. Ostlund, M. Brokemper, “Initial rotor position detections for an integrated PM<br />

synchronous motor drive”, Industry Applications Conference, IAS '95, 8-12 Oct. 1995,<br />

Vol. 1, pp. 741 – 747.<br />

91. G. Pesse, T. Paga, “A permanent magnet synchronous motor flux control scheme<br />

<strong>with</strong>out position sensor”, Conference proceedings on the EPE’97, Trondheim 1997,<br />

Vol. 4, pp. 553 – 556.<br />

177


References<br />

92. R.B. Sepe, J.H. Lang, “Real-time observer-based (adaptive) control <strong>of</strong> a permanentmagnet<br />

synchronous motor <strong>with</strong>out mechanical sensors”, IEEE Transactions on<br />

Industry Applications, Vol. 28, Issue 6, Nov.-Dec. 1992, pp. 1345 – 1352.<br />

93. P.B. Schmidt, M.L. Gasperi, G. Ray, A.H. Wijenayake, “Initial rotor angle detection<br />

<strong>of</strong> a nonsalient pole permanent magnet synchronous machine”, Industry Applications<br />

Conference, IAS '97., 5-9 Oct. 1997, Vol. 1, pp. 459 – 463.<br />

94. J.X. Shen, Z.Q. Zhu, D. Howe, “Improved speed estimation in sensorless PM<br />

brushless AC drives”, Transactions on Industry Applications, July-Aug. 2002, Vol. 38,<br />

pp. 1072 – 1080.<br />

95. Yen-Shin Lai; Fu-San Shyu; Shian Shau Tseng, “New initial position detection<br />

technique for three-phase brushless DC motor <strong>with</strong>out position and current sensors”,<br />

Transactions on Industry Applications, , March-April 2003, Vol. 39, pp. 485 – 491.<br />

96. J. Solsona, M.I. Valla, C. Muravchik, “A nonlinear reduced order observer for<br />

permanent magnet synchronous motors”, IEEE Transactions on Industrial Electronics,<br />

Aug. 1996, Vol. 43, pp. 492 - 497<br />

97. J.A. Solsona, M.I. Valla, “Disturbance and nonlinear Luenberger observers for<br />

estimating mechanical variables in permanent magnet synchronous motors under<br />

mechanical parameters uncertainties”, IEEE Transactions on Industrial Electronics, Vol.<br />

50, Issue 4, Aug. 2003 pp. 717 – 725.<br />

98. J. Solsona, M.I. Valla, C. Muravchik, “Nonlinear control <strong>of</strong> a permanent magnet<br />

synchronous motor <strong>with</strong> disturbance torque estimation”, International Conference on<br />

Industrial Electronics, <strong>Control</strong> and Instrumentation, 1997. IECON 97, Vol. 1, 9-14 Nov.<br />

1997, pp. 120 – 125.<br />

99. T. Takeshita, N. Matsui, “Sensorless control and initial position estimation <strong>of</strong> salientpole<br />

brushless DC motor”, Advanced Motion <strong>Control</strong>, 1996. AMC '96, 18-21 March<br />

1996, Vol. 1, pp. 18 – 23.<br />

100. M. Tomita, T. Senjyu, S. Doki, S. Okuma, “New sensorless control for brushless DC<br />

motors using disturbance observers and adaptive velocity estimations”, IEEE<br />

Transactions on Industrial Electronics, Vol. 45, Issue 2, April 1998, pp. 274 – 282.<br />

101. M. Tursini, R. Petrella, F. Parasiliti, “Initial rotor position estimation method for PM<br />

motors”; IEEE Transactions on Industry Applications, Nov.-Dec. 2003, Vol. 39, pp.<br />

1630 – 1640.<br />

102. R. Wu, G.R. Slemon, “A permanent magnet motor drive <strong>with</strong>out a shaft sensor”,<br />

Industry Applications Society Annual Meeting, 1990., Conference Record <strong>of</strong> the 1990<br />

IEEE, 7-12 Oct. 1990, Vol.1, pp. 553 – 558.<br />

103. D. Yousfi, M. Azizi, A. Saad, “Sensorless position and speed detection for permanent<br />

magnet synchronous motor”; Power Electronics and Motion <strong>Control</strong> Conference,<br />

PIEMC 2000, 15-18 Aug. 2000,Vol. 3, pp. 1224 – 1229.<br />

104. D. Yousfi, M. Azizi, A. Saad, “Robust position and speed estimation algorithm for<br />

permanent magnet synchronous drives”, Industry Applications Conference, 8-12 Oct.<br />

2000, Vol.3, pp. 1541 – 1546.<br />

Others<br />

105. H.-B. Shin, “New antiwindup PI controller for variable-speed motor drives”<br />

IEEE Transactions on Industrial Electronics, June 1998, Vol. 45, pp. 445 – 450.<br />

106. MathWorks, Inc, "Matlab® The Language <strong>of</strong> Technical Computing", Release 12,<br />

2000.<br />

Papers written during work on this thesis<br />

107. D. L. Sobczuk, D. Świerczyński "DSP implementation <strong>of</strong> vector controlled inverter<br />

fed induction motor <strong>with</strong> neural network speed estimator" SENE '99 tom. 2 str. 601-<br />

606.<br />

178


References<br />

108. D. Świerczyński, Marian P. Kaźmierkowski, Paweł Z. Grabowski "Nowe algorytmy<br />

estymacji strumienia stojana w silnikach indukcyjnych klatkowych zasilanych z<br />

falowników PWM" SENE '99 tom. 2 str. 649-654.<br />

109. M. Angielczyk, M. Malinowski, D. Świerczyński, „Stanowisko laboratoryjne oparte<br />

o kartę DSP do sterowania prostownikiem PWM oraz silnikiem PMSM zasilanym z<br />

falownika PWM”, Kamień Śląski 2000, Międzynarodowe XII Sympozjum<br />

Mikromaszyny i Serwonapędy, str. 480-487.<br />

110. D. Świerczyński, M. P. Kaźmierkowski.:„Bezpośrednie sterowanie momentu silnika<br />

synchronicznego o magnesach trwałych zasilanym z falownika PWM”, SENE 2001<br />

,tom II, str. 629<br />

111. D. Świerczyński, M. P. Kaźmierkowski."<strong>Direct</strong> torque control <strong>of</strong> permanent magnet<br />

synchronous motor (PMSM) using space vector modulation (<strong>DTC</strong>-<strong>SVM</strong>)", SME'2002,<br />

pp. 245<br />

112. D. Świerczyński, M. P. Kaźmierkowski, Frede Blaabjerg "DSP Based <strong>Direct</strong> <strong>Torque</strong><br />

<strong>Control</strong> <strong>of</strong> Permanent Magnet Synchronous Motor (PMSM) Using <strong>Space</strong> <strong>Vector</strong><br />

<strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>)", IEEE-ISIE 2002, L’Aquila, Italy, pp. 723-727.<br />

113. D. Świerczyński, M. Żelechowski "Studia symulacyjne nad universalną strukturą<br />

bezpośredniego sterowania momentem i strumieniem dla silników synchronicznych o<br />

magnesach trwałych oraz silników asynchronicznych ", Modelowania i Symulacja MiS<br />

2002, Zakopane.<br />

114. D. Świerczyński, M. P. Kaźmierkowski, Frede Blaabjerg "<strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong><br />

<strong>with</strong> <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>) for Permanent Magnet Synchronous<br />

Motor (PMSM)", EPE-PEMC 2002, Dubrovnik & Cavtat, Croatia.<br />

115. D. Świerczyński, M. Żelechowski "Universalną strukturą bezpośredniego sterowania<br />

momentem i strumieniem dla silników synchronicznych o magnesach trwałych oraz<br />

silników asynchronicznych ", Mikromaszyny i Serwonapędy MiS 2002, Krasiczyn.<br />

116. D. Świerczyński, M. P. Kaźmierkowski " <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> <strong>of</strong> Permanent<br />

Magnet Synchronous Motor (PMSM) Using <strong>Space</strong> <strong>Vector</strong> <strong>Modulation</strong> (<strong>DTC</strong>-<strong>SVM</strong>) –<br />

Simulation and Experimental Results", IECON 2002, Sevilla, Spain.<br />

117. D.Świerczyński, M. Żelechowski „Bezpośrednie sterowanie momentem i<br />

strumieniem dla silników synchronicznych i asynchronicznych”, II Krajowa<br />

Konferencja MiS-2 Modelowanie i Symulacja, Kościelisko, 24-28 czerwca 2002, pp.<br />

187-194.<br />

118. D. Świerczyński, M. Żelechowski ”Universal structure <strong>of</strong> direct torque control for<br />

AC motor drives”, III Summer Seminar on Nordick Network for Multi Disciplinary<br />

Optimised Electric Drives, June 21-23, 2003, Zegrze, Poland<br />

119. M. Jasiński D. Świerczyński, M. P. Kazmierkowski, M. Malinowski ”<strong>Direct</strong> <strong>Control</strong><br />

<strong>of</strong> an AC-DC-AC converter ”, III Summer Seminar on Nordick Network for Multi<br />

Disciplinary Optimised Electric Drives, June 21-23, 2003, Zegrze, Poland<br />

120. M. Żelechowski, D. Świerczyński, M. P. Kazmierkowski, J. Załęski „Uniwersalne<br />

napędy falownikowe ze sterowaniem mikroporcesorowym nowej generacji”,<br />

Elektro.info Nr 6 (17) 2003, str. 26-28<br />

121. D. Świerczyński, M. Żelechowski, „Universal Structure <strong>of</strong> <strong>Direct</strong> <strong>Torque</strong> <strong>Control</strong> for<br />

AC Motor Drives”, V Power Electronics Seminar, 7-8 July, 2003, Berlin, Niemcy<br />

122. M. Jasinski, D. Swierczynski, M. P. Kazmierkowski "Novel Sensorless <strong>Direct</strong> Power<br />

and <strong>Torque</strong> <strong>Control</strong> <strong>of</strong> <strong>Space</strong> <strong>Vector</strong> Modulated AC/DC/AC Converter", In Proc. <strong>of</strong> the<br />

ISIE 2004 Conf., 4-7 May, Ajaccio, France, pp.1147- 1152.<br />

123. D. Świerczyński, M. Żelechowski ”Universal structure <strong>of</strong> direct torque control for<br />

AC motor drives”, Przegląd Elektrotechniczny, May, 2004, Poland, pp. 489-492.<br />

124. M. P. Kaźmierkowski , M. Żelechowski, D. Świerczyński „Simple <strong>DTC</strong>-<strong>SVM</strong><br />

<strong>Control</strong> for Induction and PM Synchronous Motor”, In Proc. <strong>of</strong> the ICEM 2004 Conf.,<br />

5-8 September, Krakow, Poland, on CD.<br />

125. M. P. Kaźmierkowski , M. Żelechowski, D. Świerczyński „<strong>DTC</strong>-<strong>SVM</strong> an Efficient<br />

Method for <strong>Control</strong> Both Induction and PM Synchronous Motor”, In Proc. <strong>of</strong> the EPE-<br />

PEMC 2004 Conf., 2-4 September, Riga, Latvia, on CD.<br />

179


References<br />

126. B. W. Mehdawi, D. Świerczyński, M. P. Kaźmierkowski „Design and Investigation<br />

<strong>of</strong> <strong>Vector</strong> <strong>Control</strong> Scheme PWM Inverter-fed Nonsalient Permanent Magnet<br />

Synchronous Motor”, In Proc. <strong>of</strong> the PPEE 2005 Conf., 2-5 April, Poland, Wisla, on<br />

CD.<br />

180

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!